text
stringlengths
1
2.56M
id
stringlengths
40
40
metadata
dict
\section{Introduction}\label{Intro} The original 1970 conception of Mastermind by Mordechai Meirowitz was a sequence reconstruction game. One player (the \emph{codemaker}) would construct a hidden sequence of four pegs, each peg being one of six colors, and the other player (the \emph{codebreaker}) would make guesses of the same form, receiving feedback after each guess regarding how close they were to the hidden sequence. In 1963 Erd\H os and R\'enyi \cite{ER63} studied the two-color variant of this game, and after the release of Mastermind, Knuth showed that a minimax strategy guarantees guessing the hidden vector in no more than 5 turns \cite{DK76}. Many authors have since studied algorithms to minimize the number of guesses required in the worst case \cite{DS13,VC83,KT86,MG12,OS13,JP11,CC96,JP15, WG03, WG04}, and almost all of these results will be introduced and discussed at relevant points in this paper. We note here that the work of J{\"a}ger and Peczarski \cite{JP11,JP15} and Goddard \cite{WG03,WG04} deal with finding explicit optimal bounds for small numbers of colors and pegs, whereas we deal with asypmtotics when both of these quantities are large. The variants of Mastermind which we study are defined by the following parameters: \begin{enumerate}[label=(\roman*)] \item ($k$) \textit{Size of Alphabet} \item ($n$) \textit{Length of Sequence.} The hidden vector and all guess vectors will be elements of $[k]^n$. \item ($\Delta$) \textit{Distance Function.} $\Delta$ takes as inputs two vectors in $[k]^n$. The output may, for example, be a single integer, but this will not always be the case. Most research studies the two following distance functions: \begin{enumerate}[label=\alph*.] \item\textit{``Black-peg and white-peg.''} Informally, a black peg denotes ``the correct color in the correct spot,'' and a white peg denotes ``the correct color in an incorrect spot.'' For two vectors $Q_t$ and $H$, the black-peg and white-peg distance function is the ordered pair $\Delta(Q_t, H) := (b(Q_t, H), w(Q_t, H))$ where \begin{equation*}\label{blackHitsDefinition} b(Q_t, H) = \left|\{i\in [1,n] \mid q_i = h_i\}\right|, \end{equation*} and \begin{equation*}\label{whiteHitsDefinition} w(Q_t, H) = \max_{\sigma}~b(\sigma(Q_t), H) - b(Q_t, H), \end{equation*} where $\sigma$ iterates over all permutations of $Q_t$. This variant is the distance function used in the original game of Mastermind. \item\textit{``Black-peg-only.''} This is simply $\Delta(Q_t, H) := b(Q_t, H)$, where $b$ is defined as above. \end{enumerate} \item($R$) \textit{Repetition.} A commonly-studied variant of the game introduces the restriction that the guesses and vectors cannot have repeated components, i.e. they are vectors of the form $v \mid i \neq j \Rightarrow v_i \neq v_j$. \item($A$) \textit{Adaptiveness.} In the \textit{adaptive} variant of Mastermind, the codebreaker receives $\Delta(Q_t, H)$ after each guess $Q_t$, and may use this information to inform the selection of $Q_{t+1}$. In the \textit{non-adaptive} variant, the codebreaker submits any number of queries $Q_1, Q_2, \ldots, Q_m$ all at once (so the codebreaker chooses $m$). The codemaker then reports the distances $(\Delta(Q_1, H), \Delta(Q_2, H), \ldots, \Delta(Q_m, H))$, after which the codebreaker must determine $H$ without submitting any additional guesses. \end{enumerate} The adaptive variant with repetitions allowed is the most extensively studied in the literature \cite{DK76,VC83,MG12,CC96}. Doerr, Sp\"{o}hel, Thomas, and Winzen obtain both strong asymptotic lower bounds and an asymptotic improvement in algorithm performance \cite{DS13}, in part applying techniques from \cite{NB09,GK00}. We focus only on analyzing the worst-case performance of query strategies for these variants of Mastermind. That is, we always consider the number of queries necessary to guarantee identification of any hidden vector. As such, we will be only be focusing on deterministic strategies for the codebreaker. Our first main result concerns the \textit{Permutation Variant}, the black-peg, adaptive, no-repeats variant in which $n=k$. In this variant, the hidden sequence $H$ and all guesses $Q_t$ are permutations of $[n]$, hence the name. \begin{theorem}\label{permutationGameTheorem} For any strategy in the Permutation Variant with $n$ sufficiently large, the codebreaker must use at least $n - \log\log n$ guesses to determine $H$ in the worst case. \end{theorem} Explicit algorithms that take $O(n \log n)$ turns to solve this variant were developed by Ko and Teng \cite{KT86}, and El Ouali and Sauerland \cite{OS13}. Ko and Teng approach the problem with an algorithm akin to binary search. El Ouali and Sauerland improve this algorithm and extend it to handle variants with $k\ge n$, while also achieving an average factor of two reduction in the number of queries needed to identify $H$. Via a basic information-theoretic argument, one can show that the Permutation Game satisfies $f(n, n)\ge n - n/\log n + c$ for some constant $c>0$. We improve this lower bound to $f(n, n)\ge n- \log \log n$ for sufficiently large $n$. To our knowledge, this constitutes the first improvement over the trivial information-theoretic lower bound for the Permutation Game variant of Mastermind. Our second result concerns Knuth's Minimax algorithm for adaptive variants, which was first introduced in 1976: \begin{definition}[Knuth's Minimax Algorithm \cite{DK76}]\label{minimax def} At each turn, assign each query a score equal to the maximum across all responses of the number of possible values of $H$ that agree with that response. Guess the query with the minimum score. \end{definition} \begin{theorem}\label{minimaxTheorem} Knuth's Minimax algorithm identifies any hidden sequence $H$ in at most $nk$ queries. \end{theorem} Knuth's Minimax algorithm is empirically near-optimal for solving small games of Mastermind ($n$ and $k$ less than 10) in as few guesses as possible. However, it has proven difficult to analyze the asymptotic performance of the Minimax algorithm, primarily because its behavior is determined by the distribution remaining solutions after a series of guesses, which is difficult to analyze in general \cite{ KT86, OS13}. To our knowledge, this is the first upper bound on the worst-case performance of the minimax algorithm, but if it performs near-optimally for large $n$ and $k$, we would expect this bound to be much smaller. We know, for example, that algorithms exist that use only $O(n \log k)$ guesses when $k$ is not too large \cite{DK76, KT86}. Our third result concerns to non-adaptive variants of Mastermind. We extend the following theorem: \begin{theorem}[Doerr, Sp\"ohel, Thomas, and Winzen, \cite{DS13}]\label{DSTW nonadapt} In black-peg, non-adaptive Mastermind with repeats, $ \Omega\left( n\log(k)\right) $ guesses are required to identify $H$, \end{theorem} proving the result in the no-repeats case. \begin{theorem}\label{nonAdaptiveTheorem} In black-peg, non-adaptive Mastermind with no repeats, $ \Omega\left( n\log(k)\right) $ guesses are required to identify $H$. \end{theorem} For neither variant do the known upper bounds match these lower bounds when $n > k$; in the with-repeats case a corollary of a result in \cite{DS13} gives an upper bound of $O(k \log k)$ guesses, and for the no-repeats case no improvement over the $nk$ bound is known. On the other hand, the authors of \cite{DS13} are able to extend a result of Chv\'atal \cite{VC83} to provide tight bounds for $n \le k$ with in the with-repeats case. \subsection{Structure of the Paper} In Section 2 we prove Theorem 1, and the proof is found in 2.2. In Section 3 we prove Theorem 2, and in Section 4 we discuss Theorem 3, with the proof in 4.2. \section{Adaptive Variants of Mastermind} \subsection{The Permutation Game} We begin with notation necessary for the proof of Theorem 1. Recall that this is the adaptive, no-repeats variant with $n=k$. White pegs provide no information in this variant, so without loss of generality we assume black-peg responses. Throughout this section we will let $f(n)$ be the number of guesses required by an optimal strategy for the permutation game on $[n]^n$. We will bound $f(n)$ from below. We will use the derangement function $D(n)$, which counts the number of permutations in $S_n$ with no fixed points. It has the explicit form \begin{equation}\label{derangements} D(n) = n!\sum_{i=0}^n \frac{(-1)^i}{i!}, \end{equation} and is the nearest integer to $n!/e$. \subsubsection{Trivial Lower Bound} To motivate the proof of Theorem 1, we begin with the following simple result. \begin{proposition}[Trivial Lower Bound] \begin{equation*} f(n)\ge \log_{n}(n!) = n-\frac{n}{\ln(n)}+O(1). \end{equation*} \end{proposition} \begin{proof} In a deterministic strategy with $t$ queries, there are $n^t$ possible responses (specifically, $\{0,1,\ldots,n-2,n\}^t$). When $t < \log_{n}(n!)$ there are fewer possible responses than possible values of $H$, so by the pigeonhole principle at least two distinct values of $H$ will produce the same set of responses, and the codebreaker will be unable to distinguish between the two. \end{proof} \subsubsection{Solution subsets}\label{Buckets Definition} We now improve this trivial bound. Assume the codebreaker has some arbitrary fixed, deterministic guessing strategy. In $t$ turns of the game, the codebreaker has submitted queries $Q_1, \dots, Q_t$ and received responses $r_1,\dots,r_t$. The set of solution vectors $h$ satisfying $\Delta(Q_i,h) = r_i$ for all $i$ will be called the \textit{remaining solution set} $S_t$, and the codebreaker wins exactly when $|S_t| = 1$. Now we analyze the performance of the codebreaker's strategy. On turn $t$, the codebreaker makes the guess $Q_t$ according to some deterministic procedure. Each of the $n$ possible responses $\Delta(Q_t,H)$ produces a different remaining solution set $S_t$. Moreover, these different choices of $S_t$ partition $S_{t-1}$, as every vector in $S_{t-1}$ agrees with exactly one value of $\Delta(Q_t,H)$. We call the subset of $S_{t-1}$ corresponding to a response $r$ a \textit{solution subset}, formally defined as \begin{equation*} B_t(r) = \left\{h\in S_{t-1}\mid \Delta(Q_t, h) = r\right\}, \end{equation*} where $S_0$ is the set of all permutations of $[n]$. Following this notation, we see that $S_t = B_{t}(\Delta(Q_t,H))$. It will be useful to know the sizes of the sets $B(r,Q) := \{h | \Delta(h,Q) = r\}$. The number of elements in $B(r,Q)$ is equal to the number of ways to choose $r$ indices that are fixed points with respect to the query sequence $Q_1$ multiplied by the number of ways to permute the remaining $n-r$ colors without any fixed points. Hence we have \begin{equation*} |B(r,Q)| = \binom{n}{r}D(n-r). \end{equation*} So we may define $B(r) := B(r,Q)$ as the right-hand side is independent of $Q$. By definition, $B_t(r)$ for a fixed $Q_t$ is a subset of $B(r,Q_t)$, and so \begin{equation*} |B_t(r)|\le |B(r)| = \binom{n}{r}D(n-r). \end{equation*} \subsection{Proof of Theorem 1} Continuing, we make use of two technical lemmas. The first bounds the sums of sizes of the subsets defined above: \begin{lemma} For any positive integer $n$, we have: \begin{equation*} \sum_{i=x}^n\binom{n}{i}D(n-i) \le \frac{n!}{x!}. \end{equation*} \end{lemma} \begin{proof} We give a combinatorial proof. The left-hand side denotes the number of permutations of an $n$-element vector which have at least $x$ fixed points. The right-hand side denotes the number of ways to choose $x$ fixed points and simply permute the rest of the vector. This counts all vectors with at least $x$ fixed points at least once (and over-counts by some margin) so the inequality holds. \end{proof} This will allow us to prove a bound on the worst-case size of $|S_t|$. Note that for a fixed strategy and hidden vector $H$, every query $Q_t$, response $r_t$, and remaining solution subset $S_t$ are completely determined. \begin{lemma} For any fixed deterministic guessing strategy and $C_n < n$, there is at least one choice of hidden vector such that \begin{equation*} \frac{|S_t|}{n!}\ge \frac{C_{n}! - (H_{C_{n}+t} - H_{C_{n}})}{(C_n+t)!}, \end{equation*} for all $0\le t\le n-C_{n}$, where $H_n=\sum_{i = 1}^n\frac{1}{i}$ is the $n\nth$ harmonic number. \end{lemma} We prove Lemma 2 at the end of the section.\\ \begin{proof}[Proof of Theorem 1] Apply Lemma 2 for $t = n-C_n$. Then we have \begin{equation*} |S_{n-C_{n}}| \ge n!\left(\frac{C_{n}! - (H_n - H_{C_{n}})}{n!}\right) = C_{n}! - (H_n-H_{C_{n}}). \end{equation*} For any $C_{n}$ such that $C_{n}! - (H_n-H_{C_{n}}) > 1$, the above bound gives $|S_{n-C_n}| > 1$. This would mean that after $n-C_n$ guesses of any strategy, the remaining solution set is not necessarily reduced to a single element after $n-C_n$ guesses, allowing us to state $f(n) > n-C_n$. Noting that $H_n$ is asymptotic to $\log n$ and both grow to infinity, as long as $\log n = o(C_n!)$ we will eventually have that $C_n!-H_n > 1$. Since we have, for example, that $\log x = o((\log \log x)!)$, we have that with $C_n = \lceil \log \log n \rceil$ the above inequality will eventually be satisfied. In conclusion, when $n$ is sufficiently large, the minimum number of remaining possible solutions after $n - \lceil \log\log n \rceil$ guesses is at least \begin{equation*} S_{n - \lceil \log\log n \rceil} \ge (\log\log n)! - (H_n - H_{\log\log n}) > 1. \end{equation*} Thus there is no strategy that can identify any hidden sequence in fewer than $n-\log\log n$ turns, which concludes the proof. \end{proof} \subsection{Proof of Lemma 2} \begin{proof}[Proof] Recall that $S_t$ is the set of sequences that match the responses to the first $t$ questions of some fixed deterministic guessing strategy, given some hidden code $H$. Since all queries are possible when 0 questions have been asked, we have $|S_0| = n!$. From our definition of $B_t(r)$ above, the worst-case size of $S_t$ given $S_{t-1}$ is \begin{equation*} \max_{r\in\{0,1,\ldots,n\}}|B_{t}(r)|. \end{equation*} The minimum possible value of this maximum occurs when the subsets partition $S_{t-1}$ as evenly as possible. We know that $|B_t(r)| \leq |B(r)|$, which is easily seen to be decreasing in $r$. In the optimal distribution of this type, some subsets of higher index will have size equal to their upper bound $B(r)$, while the rest will be partially filled to some fixed amount. So with $|S_{t-1}|$ solutions remaining, there is an optimal $x$ such that completely filling subsets $x$ through $n$ and splitting the remaining solutions among subsets 0 through $x-1$ will give us this best distribution, and therefore a lower bound on $|S_t|$ in the worst case. Hence, for some optimal value of $x$, \begin{equation} |S_t| \ge \frac 1x \left(|S_{t-1}|-\sum_{i = x}^{n}B(r)\right) = \frac 1x \left(|S_{t-1}|-\sum_{i = x}^{n}\binom{n}{i}D(n-i)\right). \end{equation} In fact, $x$ is `optimal' precisely in that it maximizes the right-hand side of this inequality, and so the inequality in fact holds for all $x$. We apply Lemma 1 to bound the rightmost term and get \begin{equation}\label{recur} |S_t| \ge \frac{1}{x}\left(|S_{t-1}|-\frac{n!}{x!}\right). \end{equation} Given this recurrence, we proceed to prove Lemma 2 by induction. With $t=0$, we have $|S_0| = n!$. Then \begin{equation*} 1 =\frac{|S_0|}{n!}\ge \frac{C_{n}! - (H_{C_{n}}-H_{C_{n}})}{C_{n}!} = 1, \end{equation*} and the inequality is satisfied. \\ Now we move to the general case. Recalling that (\ref{recur}) holds for all $x$, we let $x=t+C_{n}$, giving \begin{equation*} \frac{|S_t|}{n!} \ge \frac{1}{C_{n}+t}\left(\frac{|S_{t-1}|}{n!} -\frac{1}{(C_{n}+t)!}\right). \end{equation*} Assuming the lemma inductively for $t-1$, we obtain \begin{align*} \frac{|S_t|}{n!} & \ge \frac{1}{C_{n}+t}\left(\frac{C_{n}!-(H_{C_{n}+t-1} - H_{C_{n}})} {(C_{n}+t-1)!} - \frac{1}{(C_{n}+t)!}\right)\\ & \ge \left(\frac{C_{n}! - (H_{C_{n}+t-1} - H_{C_{n}}) - \frac{1}{C_{n}+t}} {(C_{n}+t)!}\right)\\ & \ge \left(\frac{C_{n}! - (H_{C_{n}+t} - H_{C_{n}})}{(C_{n}+t)!}\right), \end{align*} which completes the induction. \end{proof} \textbf{Update:} El Ouali, Glazik, Sauerland, and Srivastav \cite{OS16} have announced an improvement of the lower bound in Theorem \ref{permutationGameTheorem} from $n-\log\log(n)$ to $n$, and an extension to $k > n$ with a lower bound of $k$. \section{Linear Algebra and the Minimax Algorithm} We now turn to a general upper bound on all Mastermind variants and note its application to Knuth's minimax algorithm in particular. We will represent an arbitrary query or hidden vector as a $(0,1)$-vector $A \in \R^{nk}$ in the following manner: $$ A_{in+j} = \begin{cases} 1 & \text{this guess/solution assigns the $i^\text{th}$ spot the $j^\text{th}$ color} \\ 0 & \text{otherwise}, \end{cases} $$ where $0 \leq i \leq n-1$ and $0 \leq j \leq k-1$. As such, each set of indices $A_{in},\dots,A_{in+k-1}$ will have exactly one 1, as the $i^\text{th}$ position is exactly one color. With this notation, the black-peg distance becomes the dot product of the guess and the hidden vector, as there will be a contribution to the dot product exactly when both vectors have a one in the same spot, i.e. there is the same color in the same spot of both vectors. The goal of Mastermind is then to find the unique valid $(0,1)$-vector such that its dot product with the hidden vector is $n$. By linearity of the dot product, once the codebreaker has queried any set of queries, it is possible to deduce the black-peg response to any linear combination of those queries as the corresponding linear combination of their responses. It follows that a winning strategy is simply to query a basis for the span of the set of valid queries, which, as a subspace of $\R^{nk}$, is of size at most $nk$. Note that this strategy does not make use of adaptive feedback, of white-peg responses, or the condition on repeated colors, and hence applies to every variant of Mastermind studied in this paper. This logic also allows us to bound the minimax algorithm (Definition \ref{minimax def}). \subsection{Proof of Theorem 2} The theorem can be easily reduced to the following lemma: \begin{lemma} At each turn, if the minimax algorithm has not determined the hidden vector, it guesses a vector that is linearly independent of its previous guesses. \end{lemma} \begin{proof} From the reasoning above, when the hidden vector is not known it must be linearly independent from the previous queries (and indeed, so must the entire remaining solution set). At this point, guessing a vector $q$ that is a linear combination of the previous guesses returns no new information, so the size of the remaining solution set remains unchanged and the guess $q$ receives a score of $|S_{t-1}|$. If we can find a new query $r$ that is guaranteed to eliminate at least one vector from the remaining solution set, the minimax algorithm will choose $r$ over $q$. Choosing any vector from the remaining solution set will do the trick; a black-peg response of $n$ will cut $S_t$ down from at least two vectors to the single vector $r$, whereas any other response will certainly eliminate $r$ from $S_t$. Thus, the minimax algorithm will never make a guess that is linearly dependent on its previous guesses. \end{proof} Theorem \ref{minimaxTheorem} follows as an immediate corollary from this lemma: Since these vectors are members of $\R^{nk}$, the minimax algorithm can make at most $nk$ linearly independent guesses. After that, the minimax cannot make a linearly independent guess and by the above lemma, the hidden vector must be uniquely determined. \section{Non-Adaptive Variants} This section follows closely the reasoning in \cite{DS13} as they analyze non-adaptive games. We perform an analysis of the black-peg, non-adaptive, no-repeats variant. We make use of \textit{(Shannon) Entropy}, defined for a random variable $X$ to be: \begin{equation*} H(X) := \sum_{x\in \text{Domain}(X)}\mathbb{P}[X=x]\cdot (-\log_2(\mathbb{P}[X=x])). \end{equation*} Entropy is \textit{subadditive}, that is, if $X_1, X_2, \ldots, X_n$ are random variables and $X = (X_1,\dots,X_n)$ is the random vector containing the $X_i$ as entries, then \begin{equation}\label{subadditivityDefinition} H(X)\le \sum_{i=1}^{n}H(X_i). \end{equation} \subsection{Proof of Theorem \ref{nonAdaptiveTheorem}} \begin{proof} Consider a set $\{Q_1, Q_2, \ldots, Q_s\}$ of $s$ query sequences such that any possible hidden sequence may be uniquely determined by the responses $\{\Delta(Q_t, H)\}$. That is, if two possible hidden vectors $H$ and $H'$ satisfy $\Delta(Q_i, H) = \Delta(Q_i, H')$ for all $i$, then $H = H'$. Let the hidden vector $Z$ be sampled uniformly at random from the set of $k!/(k-n)!$ possibilities. Then the responses $Y_i = b(Z, q_i)$ are now random variables, and the vector $Y = (Y_1, Y_2, \ldots, Y_s)$ is also a random variable. By our assumptions, $Y$ always uniquely determines, and is uniquely determined by, $Z$, and so $H(Z) = H(Y)$. Since $Z$ is a random variable with $k!/(k-n)!$ outcomes of equal probability, we compute \begin{equation}\label{equalEntropy} H(Y) = H(Z) = \log_2\left(\frac{k!}{(k-n)!}\right). \end{equation} Combining this with (\ref{subadditivityDefinition}) yields: \begin{equation}\label{entropySum} \log_2\left(\frac{k!}{(k-n)!}\right)\le\sum_{i=1}^s H(Y_i). \end{equation} Now we bound $H(Y_i)$ by some absolute constant. By definition, $Y_i$ is precisely the black-peg distance between $Q_i$ and $Z$. Then \begin{equation}\label{entropyDef} H(Y_i)=-\sum_{x=0}^n \mathbb{P}[Y_i=x]\cdot\log_2(\mathbb{P}[Y_{i}=x]). \end{equation} The quantity $\mathbb{P}[Y_i=x]$ is probability that $Z$ is a solution vector with $x$ fixed points with respect to the query $q_i$. Using the terminology from section \ref{Buckets Definition}, this is the probability that $Z$ is in solution subset $B(x)$. We use a simple bound on $|B(x)|$: this counts the number of permutations with exactly $x$ fixed points, which is fewer than the number of ways to choose $x$ fixed points and permute the other colors arbitrarily. Thus: \begin{align*} \mathbb{P}[Y_i = x] = \frac{|B(x)|}{\left(\frac{k!}{(k-n)!}\right)} \leq \frac{\binom{n}{x}\frac{(k-x)!}{(k-n)!}}{\left(\frac{k!}{(k-n)!}\right)} = \frac{1}{x!}\cdot\frac{n(n-1)\cdots(n-x+1)}{k(k-1)\cdots(k-x+1)} \leq \frac{1}{x!}. \end{align*} We can substitute this upper bound into (\ref{entropyDef}) for all $x<1/e$, because $-\alpha\log_2\alpha$ is an increasing function on that domain. For the first three values of $x$, we instead use the trivial upper bound $-\alpha\log_2\alpha \le 1/(e\log 2)$. Then \begin{align*} H(Y_i) &\leq \frac{3}{e\log 2}+\sum_{x=3}^n-\frac{1}{x!}\cdot\log_2\left(\frac{1}{x!}\right) \le \frac{3}{e\log 2}+\sum_{x=3}^\infty \frac{\log_2(x!)}{x!} <3. \end{align*} Combining this with (\ref{entropySum}) gives $H(Y) \leq 3s$, and since $H(Z) = H(Y)$ we have $H(Z) \leq 3s$. Substituting in (\ref{equalEntropy}) as a lower bound for $H(Z)$ and solving for $s$ gives \begin{equation*} s \geq \frac{1}{3}\log_2\left(\frac{k!}{(k-n)!}\right). \end{equation*} We can show that the right-hand side is $\Omega(n\log k)$. This is immediate for large $k$ relative to $n$ by bounding the ratio by $(k-n)^n$; for small $k$ (e.g. $k \leq 2n$) we bound the ratio by $n!$ and the claim follows from Stirling's approximation. So this gives us a lower bound of $\Omega(n \log k)$ turns for any non-adaptive strategy for Mastermind with no repeats and black-peg responses. \end{proof} \begin{comment} \subsubsection{Extension to white-peg responses} We extend the $R=T$ case of Theorem \ref{nonAdaptiveTheorem}, as proved in \cite{DS13}, to include $\Delta = bw$. We show that any black-white strategy must submit $\Omega(k)$ queries, and that we can convert a black and white-peg strategy into a black-peg-only strategy by adding $O(k)$ queries. Then changing a black-white strategy to a black-only strategy changes the number of queries by at most some absolute constant factor. So we have $f(\Delta = b) = O(f(\Delta = bw))$. Combining this with Theorem 3 implies the theorem for black-white responses as well. First we describe the conversion. Take a black-white strategy and append, for each color, a query where every spot is that color. As a black-peg only strategy, we can deduce from these queries exactly how many times each color appears in the hidden vector, from which we can determine the white-peg responses to any of the original queries. Therefore, if the old strategy had enough information to determine the hidden vector with black-white responses, this new strategy can determine the hidden vector with just black-peg responses. Since there are $k$ colors, we have appended $k$ queries in this conversion. Lastly, we show that any non-adaptive black-white strategy must submit $\Omega(k)$ queries. Assume that there are two colors $a$ and $b$ such that a non-adaptive strategy guesses neither $a$ nor $b$ in either spot 1 or 2. Then this strategy would not be able to distinguish between a hidden vector starting $a,b$ and one starting $b,a$. Then by contradiction any non-adaptive strategy must guess at least $k-1$ colors in these two spots. It can guess 2 colors in these spots per turn, for a total of at least $(k-1)/2 = \Omega(k)$ turns required to solve the game. Thus, by the above logic, we conclude $f(n,k) = \Omega(n \log k)$ in this case as well. \begin{problem} Can this argument be modified or extended to the case $R = F$, $\Delta = bw$ as well? \end{problem} \end{comment} \subsection{A Note on Non-Adaptive Variants with Repeats} Consider non-adaptive variants of Mastermind in which repetitions are allowed. When $k=n$, Theorem 13 of \cite{DS13} guarantees the existence of a set of $O(n \log n)$ queries which uniquely identify any hidden sequence $H$. When $k > n$, one can simply extend $H$ and all queries $Q_t$ by $k - n$ ``auxiliary'' positions. We fill these auxiliary positions with arbitrary colors, and adjust the codemaker's responses accordingly. Applying Theorem 13 of \cite{DS13} now guarantees the existence of a set of $O(k \log k)$ queries which will uniquely identify any hidden sequence. \section{Further Work} \begin{problem} In all adaptive variants of Mastermind, the best lower bounds are asymptotic to $n$. Can these lower bounds be improved? \end{problem} \begin{problem} In the variant of Mastermind with repeats, the best known strategy is $O(n\log\log k)$ \textnormal{\cite{DS13}}, whereas in the variant without repeats, it is $O(n \log k)$ \textnormal{\cite{OS13}}. Is it possible to extend or modify the first strategy to apply to the no-repeats game? \end{problem} \begin{problem} Theorem \ref{DSTW nonadapt} can be extended to white-peg responses as well by noting that the optimal white-peg strategy may be converted into a black-peg strategy losing a constant multiplicative factor by adding $k$ guesses, each composed entirely of pegs of a single color. The same construction does not apply to extend Theorem \ref{nonAdaptiveTheorem}. Can a lower bound be established in this case? \end{problem} \begin{problem} In non-adaptive variants, the best lower bounds are $O(n \log k)$, whereas upper bounds are either $O(k \log k)$ or $nk$. Can these bounds be brought closer together? \end{problem} \section{Acknowledgments} We would like to thank Daniel Montealegre for supervising and assisting our work throughout the summer, and Nathan Kaplan for both creating and guiding our project and for his invaluable assistance in editing this paper. We would also like to thank Sam Payne and the Summer Undergraduate Math Research at Yale program for organizing, funding, and supporting this project. SUMRY is supported in part by NSF grant CAREER DMS-1149054. \clearpage
179a692d315406be53debdaa72f6662154e3fc10
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Error Bounds}\label{Error Bounds} \subsection{The norm of a discretized Gaussian (Normalization of $\ket{\Psi(k_0)}$)} \subsubsection{Normalizing $\ket{\Psi(k_0)}$}\label{Normalizing the clock state} We calculate the norm of a state in the space $\Lambda_{\sigma,n_0}$: \begin{align} \ltwo{\ket{\Psi(k_0)}}^2 &= A^2 \sum_{k \in \mathcal{S}_d(k_0)} e^{-\frac{2\pi}{\sigma^2}(k-k_0)^2} \\ &= A^2 \left( \sum_{k\in \mathbb{Z}} e^{-\frac{2\pi}{\sigma^2}(k-k_0)^2} + \epsilon_1 \right), \\ \text{where} \abs{\epsilon_1} &= \sum_{k\in \mathbb{Z}/\mathcal{S}_d(k_0)} e^{-\frac{2\pi}{\sigma^2}(k-k_0)^2} < \frac{2 e^{-\frac{\pi d^2}{2\sigma^2}}}{1 - e^{-\frac{2\pi d}{\sigma^2}}}:=\bar\epsilon_1,\label{eq:norm bound ep bar 1} \end{align} using results from Sec. \ref{gaussianbounds}. Applying the Poissonian summation formula on the sum, \begin{align} \ltwo{\ket{\Psi(k_0)}}^2 &= A^2 \left( \frac{\sigma}{\sqrt{2}} \sum_{m \in \mathbb{Z}} e^{-\frac{\pi \sigma^2 (md)^2}{2d^2}} e^{-i 2\pi (md) k_0/d} + \epsilon_1 \right) \\ &= A^2 \left( \frac{\sigma}{\sqrt{2}} + \epsilon_2 + \epsilon_1 \right), \\ \text{where} \abs{\epsilon_2} &\leq \sum_{m\in\mathbb{Z}-\{0\}} e^{-\frac{\pi \sigma^2 m^2}{2}} < \frac{2 e^{-\frac{\pi \sigma^2}{2}}}{1 - e^{-\pi \sigma^2}}:=\bar\epsilon_2.\label{eq:norm bound ep bar 2} \end{align} Thus for $\ket{\Psi_\textup{nor}(k_0)}$ defined in Def. \ref{def:Gaussian clock states}, we have that $A$ given by Eq. \eqref{eq:A normalised} satisfies \begin{equation}\label{eq:up low bounds for A normalize} \left(\frac{2}{\sigma^2}\right)^{1/2}-\frac{\bar\epsilon_1+\bar\epsilon_2}{\frac{\sigma}{\sqrt{2}}\left(\frac{\sigma}{\sqrt{2}}+\bar\epsilon_1+\bar\epsilon_2 \right)}\leq A^2\leq \left(\frac{2}{\sigma^2}\right)^{1/2}+\frac{\bar\epsilon_1+\bar\epsilon_2}{\frac{\sigma}{\sqrt{2}}\left(\frac{\sigma}{\sqrt{2}}+\bar\epsilon_1+\bar\epsilon_2 \right)}, \end{equation} where $\bar\epsilon_1$, $\bar\epsilon_2$ are given by Eqs. \eqref{eq:norm bound ep bar 1} and \eqref{eq:norm bound ep bar 2} respectively. Of the two bounds $\bar\epsilon_1$ and $\bar\epsilon_2$, the first is the greater error for $\sigma>\sqrt{d}$, while the second dominates for $\sigma<\sqrt{d}$. At the midpoint, if $\sigma=\sqrt{d}$, \begin{equation} \bar\epsilon_1= \bar\epsilon_2 = \frac{2 e^{-\pi d/2}}{1 - e^{-2\pi}}. \end{equation} \subsubsection{Re-normalizing $\ket{\Psi(k_0)}$}\label{Re-normalizing the clock state} In some instances we will have to re-normalize the state $\ket{\Psi(k_0)}$. This will consist in upper bounding $A/A'$ where both $A$ and $A'$ satisfy Eq. \eqref{eq:A normalised} but for different values of $k_0$. We will thus want to upper bound \begin{equation} \epsilon_\textit{A}=\frac{A}{A'}-1. \end{equation} re-writing Eq. \eqref{eq:up low bounds for A normalize} using the short hand $a-\epsilon\leq A^2\leq a+\epsilon$ and noting that $\epsilon<a$, we can write \begin{equation} \epsilon_\textit{A}=\sqrt{\frac{A^2}{A^{\prime 2}}}-1\leq \sqrt{\frac{a+\epsilon}{a-\epsilon}}-1= \sqrt{1+\frac{2\epsilon}{a-\epsilon}}-1\leq \frac{2\epsilon}{a-\epsilon}=\frac{2\sqrt{2}}{\sigma}(\bar\epsilon_1+\bar\epsilon_2), \end{equation} where in the last line we have converted back to the notation of Eq. \eqref{eq:up low bounds for A normalize} and simplified the expression. Recall that $\bar\epsilon_1$, $\bar\epsilon_2$ are given by Eqs. \eqref{eq:norm bound ep bar 1} and \eqref{eq:norm bound ep bar 2} respectively. Similarly, we can lower bound $\epsilon_\textit{A}$. We find, \begin{equation} \epsilon_\textit{A}=\sqrt{\frac{A^2}{A^{\prime 2}}}-1\geq \sqrt{\frac{a-\epsilon}{a+\epsilon}}-1= \sqrt{1-\frac{2\epsilon}{a+\epsilon}}-1\geq -\frac{2\epsilon}{a+\epsilon}\geq -\frac{2\epsilon}{a-\epsilon}=- \frac{2\sqrt{2}}{\sigma}(\bar\epsilon_1+\bar\epsilon_2), \end{equation} \iffalse \subsection{Gaussian Error bounds} \twocolumngrid {\color{red} Remove the section since following sec is only relevant. check that proof refs r updated} \stepcounter{bound} \bigskip {\bf Bound E\arabic{bound}.} \text{ }If $f(x)$ is monotonically decreasing for $x\geq a-1$, where $a$ is an integer, then \begin{equation} \int_a^\infty f(x) dx < \sum_{n=a}^\infty f(n) < \int_{a-1}^\infty f(x) dx \end{equation} {\bf Proof.} From the monotonicity of $f(x)$, we have that $f(x) \leq f[x]< f(x-1)$. One can rewrite the sum as an integral, \begin{align} \sum_{n=a}^\infty f(n) &= \int_a^\infty f([x]) dx, \end{align} from which the result follows. \stepcounter{bound} \bigskip {\bf Bound E\arabic{bound}.} \text{ } $\forall a\geq X+\Delta$, \begin{equation} \left| \sum_{n=a}^\infty n f(n) \right| < \left( \Delta^2 + a + \frac{X e^{\frac{-(a-X)}{2\Delta^2}}}{1-e^{\frac{-(a-X)}{2\Delta^2}}} \right) e^{-\frac{(a-X)^2}{2\Delta^2}} \end{equation} {\bf Proof.} Since $a>X$, \begin{equation} \left| \sum_{n=a}^\infty n f(n) \right| \leq \sum_{n=a}^\infty (n-X) \left| f(n) \right| + X \sum_{n=a}^\infty \left| f(n) \right| \end{equation} Since $a\geq X+\Delta$, then $(x-X)f(x)$ is monotonically decreasing for $x\geq a$. Thus we can split the first sum into the $n=a$ term and apply Bound E3, followed by Bound E2 on the remainder of the first sum, \begin{equation} \sum_{n=a}^\infty (n-X) \left| f(n) \right| < (a-X) e^{-\frac{(a-X)^2}{2\Delta^2}} + \Delta^2 e^{-\frac{(a-X)^2}{2\Delta^2}}. \end{equation} For the second term we apply Bound E1 directly, \begin{equation} X \sum_{n=a}^\infty \left| f(n) \right| < X \frac{e^{-\frac{(a-X)^2}{2\Delta^2}}}{1 - e^{\frac{-(a-X)}{2\Delta^2}}}. \end{equation} Adding up, we obtain the result. \stepcounter{bound} \bigskip {\bf Bound E\arabic{bound}.} \text{ } \begin{align} \sum_{n=a}^\infty &(n-X)^2 f(n) \\ &< \left( (a-X)^2 + (a-X)\Delta^2 + \frac{\Delta^2}{1-e^{-\frac{a-X}{\Delta^2}}} \right) e^{-\frac{(a-X)^2}{2\Delta^2}} \nonumber \end{align} {\bf Proof.} We first split the sum (labelled $S$)into the $a=1$ term and the rest, and use the monotonicity of $(x-X)^2 f(x)$ to bound the rest by the integral \begin{align} S &< (a-X)^2 f(a) + \int_a^\infty (x-X)^2 e^{-\frac{(x-X)^2}{2\Delta^2}} dx \end{align} The integral can be evaluated by parts, \begin{align} \Delta^2(a-X)e^{-\frac{(a-X)^2}{2\Delta^2}} + \Delta^2 \int_a^\infty e^{-\frac{(x-X)^2}{2\Delta^2}} dx \end{align} The monotonicity of $f(x)$ may be used once more to bound the integral above by the sum, \begin{equation} \int_a^\infty e^{-\frac{(x-X)^2}{2\Delta^2}} dx < \sum_{n=a}^\infty e^{-\frac{(n-X)^2}{2\Delta^2}} \end{equation} Applying bound E1 and adding up the terms, we obtain the result. \stepcounter{bound} \bigskip {\bf Bound E\arabic{bound}.} \begin{equation} \Delta \sqrt{\frac{\pi}{2}} e^{-\frac{a^2}{2\Delta^2}} < \int_a^\infty e^{-\frac{x^2}{2\Delta^2}} dx < \Delta \sqrt{\frac{\pi}{2}} e^{-\frac{a^2}{4\Delta^2}} \end{equation} \emph{\bf Proof.} \begin{align} I = \int_a^\infty e^{-\frac{x^2}{2\Delta^2}} dx &= \sqrt{\int_a^\infty \int_a^\infty e^{-\frac{x^2+y^2}{2\Delta^2}} dx dy} \end{align} The region of integration is the postive quadrant of the $xy$ , minus the square bounded by $0\leq x,y\leq a$. Since the integrand is positive everywhere, we can bound the integral of the square region by the quadrant inscribed within the square ($\sqrt{x^2+y^2} \leq a$), and the quadrant within which the square is inscribed ($\sqrt{x^2+y^2} \leq \sqrt{2} a$). Converting $dxdy=rdrd\theta$, and evaluating the quadrant ($0\leq\theta\leq\pi/2$) integrals, \begin{equation} \sqrt{\frac{\pi}{2}} \sqrt{\int_{\sqrt{2}a}^\infty e^{-\frac{r^2}{2\Delta^2}} r dr} < I < \sqrt{\frac{\pi}{2}} \sqrt{\int_a^\infty e^{-\frac{r^2}{2\Delta^2}} r dr}, \end{equation} from which the result follows. \onecolumngrid The bound calculated so far in this section, allow us to bound $\epsilon_1,\epsilon_3$ for lemma \ref{lemm:1}. We now want to bound $\epsilon_2$ for lemma \ref{lemm:1}. \fi \onecolumngrid \subsection{Bounds on the tails of discrete Gaussians}\label{gaussianbounds} In this section we state some well known useful bounds which will be used throughout the proofs in appendix. For $\Delta\in{\mathbbm{R}}$, we have the following bounds on the summations over Gaussian tails. \begin{lemma}\label{G0} \begin{equation} \sum_{n=a}^\infty e^{-\frac{(n-X)^2}{\Delta^2}} < \frac{e^{-\frac{(a-X)^2}{\Delta^2}}}{1-e^{-\frac{2(a-X)}{\Delta^2}}},\quad \text{for } a>X\in{\mathbbm{R}} \end{equation} \end{lemma} \begin{proof} \begin{align} \sum_{n=a}^\infty e^{-\frac{(n-X)^2}{\Delta^2}} &= \sum_{m=0}^\infty e^{-\frac{(a-X+m)^2}{\Delta^2}}= e^{-\frac{(a-X)^2}{\Delta^2}} \sum_{m=0}^\infty e^{-\frac{2m(a-X)}{\Delta^2}} e^{-\frac{m^2}{\Delta^2}} \\ &< e^{-\frac{(a-X)^2}{\Delta^2}} \sum_{m=0}^\infty e^{-\frac{2m(a-X)}{\Delta^2}} = \frac{e^{-\frac{(a-X)^2}{\Delta^2}}}{1-e^{-\frac{2(a-X)}{\Delta^2}}} \end{align} \end{proof} \begin{lemma} \begin{equation} \int_b^\infty x e^{-\frac{x^2}{\Delta^2}} dx = \frac{\Delta^2}{2} e^{-\frac{b^2}{\Delta^2}} \end{equation} \end{lemma} \begin{proof} By direct integration. \end{proof} \begin{lemma}\label{G1} \begin{equation} \sum_{n=a}^\infty (n-X) e^{-\frac{(n-X)^2}{\Delta^2}} < \left( a - X + \frac{\Delta^2}{2} \right) e^{-\frac{(a-X)^2}{\Delta^2}},\quad \text{for } a>X+\Delta\in{\mathbbm{R}} \end{equation} \end{lemma} \begin{proof} \begin{align} \sum_{n=a}^\infty (n-X) e^{-\frac{(n-X)^2}{\Delta^2}} &= (a-X) e^{-\frac{(a-X)^2}{\Delta^2}} + \sum_{n=a+1}^\infty (n-X) e^{-\frac{(n-X)^2}{\Delta^2}} \end{align} Since $(x-X)e^{-\frac{(x-X)^2}{\Delta^2}}$ is monotonically decreasing for $x>X+\Delta$, (which $a$ satisfies), \begin{equation} \sum_{n=a+1}^\infty (n-X) e^{-\frac{(n-X)^2}{\Delta^2}} < \int_a^\infty (x-X) e^{-\frac{(x-X)^2}{\Delta^2}} dx = \frac{\Delta^2}{2} e^{-\frac{(a-X)^2}{\Delta^2}} \end{equation} \end{proof} \begin{lemma} \begin{align} \sum_{n=a}^\infty (n-X)^2 e^{-\frac{(n-X)^2}{\Delta^2}} < \left( (a-X)^2 + \frac{\Delta^2}{2} \left( a - X + \frac{1}{1 - e^{-\frac{2(a-X)}{\Delta^2}}} \right) \right) e^{-\frac{(a-X)^2}{\Delta^2}},\quad \text{for } a>X+\sqrt{2} \Delta\in{\mathbbm{R}} \end{align} \end{lemma} \begin{proof} \begin{equation} \sum_{n=a}^\infty (n-X)^2 e^{-\frac{(n-X)^2}{\Delta^2}} = (a-X)^2 e^{-\frac{(a-X)^2}{\Delta^2}} + \sum_{n=a+1}^\infty (n-X)^2 e^{-\frac{(n-X)^2}{\Delta^2}} \end{equation} Since $(x-X)^2e^{-\frac{(x-X)^2}{\Delta^2}}$ is monotonically decreasing for $x>X+\sqrt{2}\Delta$, (which $a$ satisfies), \begin{align} \sum_{n=a+1}^\infty (n-X)^2 e^{-\frac{(n-X)^2}{\Delta^2}} &< \int_{a}^\infty (x-X)^2 e^{-\frac{(x-X)^2}{\Delta^2}} dx \\ &= \frac{\Delta^2}{2} \left( (a-X) e^{-\frac{(a-X)^2}{\Delta^2}} + \int_a^\infty e^{-\frac{(x-X)^2}{\Delta^2}} dx \right) \end{align} We use the monotonicity again, together with lemma \ref{G0}, \begin{align} \int_a^\infty e^{-\frac{(x-X)^2}{\Delta^2}} dx &<\sum_{n=a}^\infty e^{-\frac{(n-X)^2}{\Delta^2}} < \frac{e^{-\frac{(a-X)^2}{\Delta^2}}}{1-e^{-\frac{2(a-X)}{\Delta^2}}} \end{align} \end{proof} \subsection{Commutator section} Here we provide some supplementary details to the proofs of some of the bounds used in the proof of Theorem \ref{Quasi-Canonical commutation}.\\ \centerline{\textit{1. Bounding} $\quad\epsilon_1^{co}$} \begin{align} \ltwo{\ket{\epsilon_1^{co}}} &< T_0 \sum_{k \in \mathcal{S}_d(k_0), k > \frac{d}{2}} \abs{\psi(k_0;k)} \\ &< T_0 \sum_{k=\frac{d+1}{2}}^\infty \abs{\psi(k_0;k)} & & \text{increasing the range of the sums} \\ &< T_0 \sum_{k=\frac{d}{2}}^\infty \abs{\psi(k_0;k)} & & \text{monotonicity (decreasing) of $\psi(k_0;k)$} \\ &< T_0 \sum_{k=\frac{d}{2}}^\infty e^{-\frac{\pi d^2}{\sigma^2}(k-k_0)^2} \\ &< T_0 A \frac{e^{-\frac{\pi}{\sigma^2}\left( \frac{d}{2} - k_0 \right)^2}}{1 - e^{-\frac{2\pi}{\sigma^2}\left( \frac{d}{2} - k_0 \right)}} \\ &= \begin{cases}\displaystyle 2 T_0 A \frac{e^{-\frac{\pi d}{4} (1-\beta)^2}}{1 - e^{-\pi(1-\beta)}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 2 T_0 A \frac{e^{-\frac{\pi d^2}{4 \sigma^2}(1-\beta)^2}}{1 - e^{-\frac{\pi d}{\sigma^2}(1-\beta)}} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{2. Bounding} $\quad\epsilon_2^{co}$} \begin{align} \ltwo{\ket{\epsilon_2^{co}}} &= \frac{T_0}{d\sqrt{d}} \sum_{n=0}^{d-1} \sum_{k \in \mathbb{Z}/\mathcal{S}_d(k_0)} \abs{k\psi(k_0;k)} \\ &< \frac{T_0}{\sqrt{d}} \sum_{k \in \mathbb{Z}/\mathcal{S}_d(k_0)} \abs{k \psi(k_0;k)} & & \text{trivial sum w.r.t. $n$} \\ &< \frac{2 T_0 A}{\sqrt{d}} \sum_{k=k_0+\frac{d}{2}} k e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} & & \text{picking the larger error (right side)} \\ &< \frac{2 T_0 A}{\sqrt{d}} \sum_{k=k_0+\frac{d}{2}} \left[ (k-k_0) + k_0 \right] e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} & & \text{splitting into two sums} \\ &< \frac{2 T_0 A}{\sqrt{d}} \left[ \frac{d}{2} + \frac{\sigma^2}{2\pi} + \frac{k_0}{1 - e^{-\frac{2\pi}{\sigma^2}\frac{d}{2}}} \right] e^{-\frac{\pi}{\sigma^2} \left(\frac{d}{2}\right)^2} \\ &< \begin{cases}\displaystyle T_0 A \sqrt{d} \left( 1 + \frac{1}{\pi} + \frac{\beta}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle T_0 A \sqrt{d} \left( 1 + \frac{\sigma^2}{\pi d} + \frac{\beta}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{3. Bounding} $\quad\epsilon_3^{co}$} \begin{align} \ltwo{\ket{\epsilon_3^{co}}} &= \sqrt{d} \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=0}^{d-1} \sum_{m\in\mathbb{Z}} \frac{A\sigma}{\sqrt{d}} \abs{m \left( -\frac{2\pi\sigma^2}{d^2}(p-n_0) - i \frac{2\pi k_0}{d} \right) e^{-\frac{\pi\sigma^2}{d^2}(p-n_0)^2} }_{p=n+md} \ket{\theta_l}. \end{align} Trivial sum w.r.t. $l$, bound the error by twice the right side error (w.r.t. $m$), choose $n=-d/2$ for worst case scenario, then trivial sum over $n$, and express $n_0$ in terms of $\alpha$, \begin{align} \ltwo{\ket{\epsilon_3^{co}}} &= 2 A d^2 \sigma 2\pi \sum_{m=1}^\infty \left[ \left( m - \frac{1+\alpha}{2} \right) + \frac{1+\alpha}{2} \right] \left( \frac{\sigma^2}{d} \left( m - \frac{1+\alpha}{2} \right) + \frac{\beta}{2} \right) e^{-\pi\sigma^2 \left( m - \frac{1+\alpha}{2} \right)^2}. \end{align} Applying all of the bounds on the tails of Gaussians on each of the 4 terms, \begin{align} \frac{\ltwo{\ket{\epsilon_3^{co}}}}{2\pi A d^2} &= 2 \sigma \cdot \frac{\sigma^2}{d} \left[ \left( \frac{1-\alpha}{2} \right)^2 + \frac{1}{2\pi\sigma^2} \left( \frac{1-\alpha}{2} + 1 + \frac{1}{1 - e^{-\pi\sigma^2(1-\alpha)}} \right) \right] e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} \\ &+ 2 \sigma \cdot \frac{\sigma^2}{d} \left( \frac{1+\alpha}{2} \right) \left( \frac{1-\alpha}{2} + \frac{1}{2\pi\sigma^2} \right) e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} \\ &+ 2 \sigma \cdot \frac{\beta}{2} \left( \frac{1-\alpha}{2} + \frac{1}{2\pi\sigma^2} \right) e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} \\ &+ 2 \sigma \cdot \left( \frac{1+\alpha}{2} \right) \frac{\beta}{2} \left( \frac{1}{1 - e^{-\pi\sigma^2(1-\alpha)}} \right) e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} \\ =& \begin{cases}\displaystyle \sqrt{d} \left( (1-\alpha) + \frac{1}{\pi d} \left( 2 + \frac{1}{1 - e^{-\pi d (1-\alpha)}} \right) + \frac{\beta}{2} \left( 1 - \alpha + \frac{1}{\pi d} + \frac{1+\alpha}{1 - e^{-\pi d(1-\alpha)}} \right) \right) e^{-\frac{\pi d}{4}(1-\alpha)^2} &\mbox{if } \sigma=\sqrt{d}\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad \\[10pt]\displaystyle \sigma \left( \frac{\sigma^2}{d} (1-\alpha) + \frac{1}{\pi d} \left( 2 + \frac{1}{1 - e^{-\pi\sigma^2(1-\alpha)}} \right) + \frac{\beta}{2} \left( 1 - \alpha + \frac{1}{\pi\sigma^2} + \frac{1+\alpha}{1 - e^{-\pi\sigma^2(1-\alpha)}} \right) \right) e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} &\mbox{otherwise}\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad \end{cases} \end{align} \centerline{\textit{4. Bounding} $\quad\epsilon_4^{co}$} \begin{align} \ltwo{\ket{\epsilon_4^{co}}} &= \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m\in\mathbb{Z}/\{0\}} \abs{ \frac{d}{dx} \left( x\psi(k_0;x) \right) }_{x=l+md} = \sum_{k\in\mathbb{Z}/\mathcal{S}_d(k_0)} \abs{ \frac{d}{dx} \left( x\psi(k_0;x) \right) }_{x=k}\\ &= \sum_{k\in\mathbb{Z}/\mathcal{S}_d(k_0)} \abs{ \left( 1 - \frac{2\pi}{\sigma^2}x(x-k_0) + i \frac{2\pi n_0}{d} x \right) e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} }_{x=k} \end{align} Once again, bounding the error by twice the right side error (w.r.t. $m$), and expressing $k_0$ in terms of $\beta$, \begin{align} \ltwo{\ket{\epsilon_4^{co}}} &< 2d A \sum_{m=1}^\infty \abs{ 1 - \frac{2\pi}{\sigma^2} (k-k_0)^2 -\frac{2\pi}{\sigma^2} k_0(k-k_0) + i \pi\alpha (k-k_0+k_0) } e^{-\frac{\pi}{\sigma^2} (k-k_0)^2} \end{align} Bounding each sum as a Gaussian tail, (first term is ignored as it will only make the total smaller, being of the opposite sign) \begin{align} \ltwo{\ket{\epsilon_4^{co}}} &< 2d A \frac{2\pi}{\sigma^2} \left( \frac{d^2}{4} + \frac{\sigma^2}{2\pi} \left( \frac{d}{2} + 1 + \frac{1}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) \right) e^{-\frac{\pi d^2}{4\sigma^2}} \\ &+ 2dA \frac{2\pi}{\sigma^2} \frac{\beta d}{2} \left( \frac{d}{2} + \frac{\sigma^2}{2\pi} \right) e^{-\frac{\pi d^2}{\sigma^2}} \\ &+ 2dA \pi \alpha \left( \frac{d}{2} + \frac{\sigma^2}{2\pi} \right) e^{-\frac{\pi d^2}{\sigma^2}} \\ &+ 2dA \pi \alpha \frac{\beta}{2} \frac{1}{1 - e^{-\frac{\pi d}{\sigma^2}}} e^{-\frac{\pi d^2}{\sigma^2}} \\ &< \begin{cases}\displaystyle dA \left( d (\pi+1) (1+\beta) + 2 + \frac{2}{1 - e^{-\pi}} + \alpha \left( d (\pi+1) + \frac{\pi\beta}{1 - e^{-\pi}} \right) \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle dA \left( \left( \frac{\pi d^2}{\sigma^2} + d \right)(1+\beta) + 2 + \frac{2}{1 - e^{-\frac{\pi d}{\sigma^2}}} + \alpha \left( \pi d + \sigma^2 + \frac{\pi\beta}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) \right) e^{-\frac{\pi d^2}{4\sigma^2}} &\mbox{otherwise} \end{cases} \end{align} \textit{ALTERNATE BOUND} \begin{align} \ltwo{\ket{\epsilon_4^{co}}} &= \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m\in\mathbb{Z}/\{0\}} \abs{ \frac{d}{dx} \left( x\psi(k_0;x) \right) }_{x=l+md} \\ &= \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m\in\mathbb{Z}/\{0\}} \abs{ \left( 1 - \frac{2\pi}{\sigma^2}x(x-k_0) + i \frac{2\pi n_0}{d} x \right) e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} }_{x=l+md} \end{align} Once again, bounding the error by twice the right side error (w.r.t. $m$), and replacing $l$ by its worst case, and trivializing the sum w.r.t. $l$, and finally expressing $k_0$ in terms of $\beta$, \begin{align} \ltwo{\ket{\epsilon_4^{co}}} &< 2d A \sum_{m=1}^\infty \abs{ \left[ 1 - \frac{2\pi d^2}{\sigma^2} \left( m - \frac{1}{2} \right)^2 -\frac{2\pi d^2}{\sigma^2} \frac{\beta}{2}\left( m - \frac{1}{2} \right) + i 2\pi d \frac{\alpha}{2} \left( m - \frac{1}{2} + \frac{\beta}{2} \right) \right] } e^{-\frac{\pi d^2}{\sigma^2} \left( m - \frac{1}{2} \right)^2} \end{align} Once again, bounding each sum as a Gaussian tail, \begin{align} \ltwo{\ket{\epsilon_4^{co}}} &< 2d A \frac{2\pi d^2}{\sigma^2} \left( \frac{1}{4} + \frac{\sigma^2}{2\pi d^2} \left( \frac{1}{2} + 1 + \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} \right) \right) e^{-\frac{\pi d^2}{4\sigma^2}} \\ &+ 2dA \frac{2\pi d^2}{\sigma^2} \frac{\beta}{2} \left( \frac{1}{2} + \frac{\sigma^2}{2\pi d^2} \right) e^{-\frac{\pi d^2}{\sigma^2}} \\ &+ 2dA \pi d \alpha \left( \frac{1}{2} + \frac{\sigma^2}{2\pi d^2} \right) e^{-\frac{\pi d^2}{\sigma^2}} \\ &+ 2dA \pi d \alpha \frac{\beta}{2} \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} e^{-\frac{\pi d^2}{\sigma^2}}. \end{align} \centerline{\textit{5. Bounding} $\quad\epsilon_5^{co}$} \begin{align} \ltwo{\ket{\epsilon_5^{co}}} &= \frac{1}{\sqrt{d}} \sum_{n=0}^{d-1} \sum_{k \in \mathbb{Z}/\mathcal{S}_d(k_0)} \abs{\psi(k_0;k)}, \\ &=\sqrt{d} \sum_{k \in \mathbb{Z}/\mathcal{S}_d(k_0)} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} & & \text{trivial sum w.r.t $n$} \\ &< 2\sqrt{d} A \sum_{k-k_0=\frac{d}{2}}^\infty e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} & & \text{bound error by right side} \\ &< \begin{cases}\displaystyle 2\sqrt{d} A \frac{ e^{-\frac{\pi d}{4}}}{1 - e^{-\pi}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 2\sqrt{d} A \frac{ e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{6. Bounding} $\quad\epsilon_6^{co}$} \begin{align} \ltwo{\ket{\epsilon_6^{co}}} &< \frac{2 \pi}{T_0} \sqrt{d} \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=0}^{d-1} \sum_{m \in \mathbb{Z}} \abs{m \tilde{\psi}(k_0;n+md)} \end{align} Trivial sum w.r.t. $l$, bounding the error by the right side (w.r.t. $m$) and replacing $n$ by the worst case $n=-\frac{d}{2}$, \begin{align} \ltwo{\ket{\epsilon_6^{co}}} &= \frac{2 \pi}{T_0} d^2 \sigma A \sum_{m=1}^\infty \left( \left( m - \frac{1+\alpha}{2} \right) + \frac{1+\alpha}{2} \right) e^{-\pi^2\sigma^2 \left( m - \frac{1+\alpha}{2} \right)^2} \\ &< \begin{cases}\displaystyle \frac{2\pi A d^2 \sqrt{d}}{T_0} \left( \frac{1-\alpha}{2} + \frac{1}{2\pi d} + \left( \frac{1+\alpha}{2} \right) \frac{1}{1 - e^{-\pi d(1-\alpha)}} \right) e^{-\frac{\pi d}{4}(1-\alpha)^2} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle \frac{2\pi A d^2 \sigma}{T_0} \left( \frac{1-\alpha}{2} + \frac{1}{2\pi\sigma^2} + \left( \frac{1+\alpha}{2} \right) \frac{1}{1 - e^{-\pi\sigma^2(1-\alpha)}} \right) e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{7. Bounding} $\quad\epsilon_7^{co}$} \begin{align} \ltwo{\ket{\epsilon_7^{co}}} &= \frac{d}{T_0} \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m \in \mathbb{Z}/\{0\}} \abs{\frac{d}{dx} \psi(k_0;x)}_{x=l+md} = \frac{dA}{T_0} \sum_{k\in\mathbb{Z}/\mathcal{S}_d(k_0)} \abs{\frac{d}{dx} \psi(k_0;x)}_{x=k}\\ &< \frac{2d}{T_0} \sum_{k-k_0=\frac{d}{2}}^\infty \abs{ \frac{2\pi}{\sigma^2}(k-k_0) + i \frac{2\pi n_0}{d} } e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} \\ &< \begin{cases}\displaystyle \frac{2\pi}{T_0} d A \left( 1 + \frac{1}{\pi} + \frac{\alpha}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle \frac{2\pi}{T_0} d A \left( \frac{d}{\sigma^2} + \frac{1}{\pi} + \frac{\alpha}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{8. Bounding} $\quad\epsilon_8^{co}$} \begin{align} \ltwo{\ket{\epsilon_8^{co}}} &< d A \sum_{k=\frac{d}{2}}^\infty \abs{ \frac{2\pi}{\sigma^2}(k-k_0) + i \frac{2\pi n_0}{d} } e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} \\ &< \begin{cases}\displaystyle \pi d A \left( 1 -\beta + \frac{1}{\pi} + \frac{\alpha}{1 - e^{-\pi(1-\beta)}} \right) e^{-\frac{\pi d}{4}(1-\beta)^2} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle \pi d A \left( \frac{d}{\sigma^2}(1-\beta) + \frac{1}{\pi} + \frac{\alpha}{1 - e^{-\frac{\pi d}{\sigma^2}(1-\beta)}} \right) e^{-\frac{\pi d^2}{4\sigma^2}(1-\beta)^2} &\mbox{otherwise} \end{cases} \end{align} \iffalse \subsection{Continuity no potential section} \centerline{\textit{1. Bounding} $\quad\epsilon_1$} \begin{align} \left| \epsilon_1 \right| &\leq \frac{2\pi}{d^2} \left( \sum_{k \in\mathbb{Z}/ \mathcal{S}_d(k_0)} \left| \psi(k_0;k) \right| \right) \left( \sum_{n=0}^{d-1} n \right) < \pi A \cdot 2 \sum_{k-k_0>\frac{d}{2}} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} \\ &< \begin{cases}\displaystyle 2\pi A \frac{e^{-\frac{\pi d}{4}}}{1 - e^{-\pi}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 2\pi A \frac{e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{2. Bounding} $\quad\epsilon_2$} \begin{align} \abs{\epsilon_2} &\leq \frac{2\pi}{\sqrt{d}} \sum_{n=0}^{d-1} \sum_{m \in \mathbb{Z}/ \{0\}} \abs{m} \abs{\tilde{\psi}(k_0;n+md)} < \frac{2\pi}{\sqrt{d}} \cdot d \cdot 2 \frac{A\sigma}{\sqrt{d}} \sum_{m=1}^{\infty} \left( m - \frac{n_0}{d} + \frac{n_0}{d} \right) e^{-\pi \sigma^2 \left( m - \frac{n_0}{d} \right)^2} \\ &< \begin{cases}\displaystyle 4\pi A \sqrt{d} \left( \frac{1-\alpha}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi d(1-\alpha)}} \right) e^{-\frac{\pi d}{4}(1-\alpha)^2} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 4\pi A \sigma \left( \frac{1-\alpha}{2} + \frac{1}{2\pi\sigma^2} + \frac{1}{1 - e^{-\pi\sigma^2(1-\alpha)}} \right) e^{-\frac{\pi\sigma^2}{4}(1-\alpha)^2} &\mbox{otherwise} \end{cases} \end{align} \centerline{\textit{3. Bounding} $\quad\epsilon_3$} \begin{align} \abs{\epsilon_3} &\leq \sum_{m\in\mathbb{Z}/\{0\}} \abs{ \frac{\partial}{\partial y} \psi\left( k_0 + y ; l \right) \bigg|_{y=md}} < 2A \sum_{m=1}^\infty e^{-\frac{\pi d^2}{\sigma^2} \left( m - \frac{1}{2} \right)^2} \abs{ \frac{2\pi d}{\sigma^2} \left( m - \frac{1}{2} \right) + i \frac{2\pi n_0}{d} } \\ &< 2A \left( \frac{2\pi d}{\sigma^2} \left( \frac{1}{2} + \frac{\sigma^2}{2\pi d^2} \right) + \frac{2\pi}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} \\ &< \begin{cases}\displaystyle 4\pi A \left( \frac{1}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi d}} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 4\pi A \left( \frac{d}{2\sigma^2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} &\mbox{otherwise} \end{cases} \end{align} \subsection{Miscellaneous} \begin{align} \sum_{k-k_0>\frac{d}{2}} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} &< \frac{ e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}}, \\ \sum_{k_0-k \geq \frac{d}{2}} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} &< \frac{ e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}}. \end{align} \begin{align} \sum_{k-k_0>\frac{d}{2}} \abs{k} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} &= \sum_{k-k_0>\frac{d}{2}} (k-k_0) e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} + k_0 \sum_{k-k_0>\frac{d}{2}} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} \\ &< \sum_{k-k_0>\frac{d}{2}+1} (k-k_0) e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} + \frac{d}{2} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{k_0 e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \\ &< \int_{x-k_0=\frac{d}{2}}^\infty (x-k_0) e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} dx + \frac{d}{2} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{k_0 e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \\ &< \frac{\sigma^2}{2\pi} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{d}{2} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{k_0 e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \\ &= \left( \frac{\sigma^2}{2\pi} + \frac{d}{2} + \frac{k_0}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}}. \end{align} \begin{align} \sum_{k_0-k \geq \frac{d}{2}} \abs{k} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} &= \sum_{k_0-k \geq \frac{d}{2}} (k_0-k) e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} + k_0 \sum_{k_0-k \geq \frac{d}{2}} e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} \\ &< \sum_{k_0-k>\frac{d}{2}+1} (k_0-k) e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} + \frac{d}{2} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{k_0 e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \\ &< \int_{-\infty}^{k_0-x=\frac{d}{2}} (k_0-x) e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} dx + \frac{d}{2} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{k_0 e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \\ &< \frac{\sigma^2}{2\pi} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{d}{2} e^{-\frac{\pi d^2}{4\sigma^2}} + \frac{k_0 e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \\ &= \left( \frac{\sigma^2}{2\pi} + \frac{d}{2} + \frac{k_0}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}}. \end{align} \begin{align} \frac{d}{dp} \tilde{\psi}(k_0;p) &= A \frac{\sigma}{\sqrt{d}} e^{-\frac{\pi\sigma^2}{d^2}(p-n_0)^2} e^{-i 2\pi pk_0/d} \left( -\frac{2\pi\sigma^2}{d^2}(p-n_0) - i \frac{2\pi k_0}{d} \right). \end{align} \begin{align} \sum_{m \in \mathbb{Z}} \abs{ m \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} } = 2 \pi A \frac{\sigma}{\sqrt{d}} \sum_{m \in \mathbb{Z}} e^{-\frac{\pi \sigma^2}{d^2}(n-n_0+md)^2} \abs{m} \abs{ \frac{\sigma^2}{d^2} (n-n_0+md) + i \frac{k_0}{d} }. \end{align} For $m \geq 1$, \begin{align} \abs{md+n-n_0} e^{-\frac{\pi \sigma^2}{d^2}(md+n-n_0)^2} < (md - n_0) e^{-\frac{\pi\sigma^2}{d^2}(md-n_0)^2}, \end{align} \begin{align} \therefore \sum_{m=1}^\infty \frac{\sigma^2}{d^2} \abs{m} \abs{md+n-n_0} e^{-\pi \sigma^2 \left( m - \frac{n-n_0}{d} \right)^2} &< \frac{\sigma^2}{d} \sum_{m=1}^\infty \left( \left( m - \frac{n_0}{d} \right)^2 + \frac{n_0}{d} \left( m - \frac{n_0}{d} \right) \right) e^{-\pi\sigma^2 \left( m - \frac{n_0}{d} \right)^2} \\ &< \frac{\sigma^2}{d} \left( \alpha + \frac{1}{2\pi\sigma^2} \left( 2 + \frac{1}{1 - e^{-2\pi\sigma^2\alpha}} \right) \right) e^{-\pi\sigma^2\alpha^2}. \end{align} \begin{align} \sum_{m=1}^\infty \abs{m} \frac{k_0}{d} e^{-\frac{\pi \sigma^2}{d^2}(md+n-n_0)^2} &= \sum_{m=1}^\infty \left( \left( m-\frac{n_0}{d} \right) + \frac{n_0}{d} \right) \frac{k_0}{d} e^{-\frac{\pi \sigma^2}{d^2}(md-n_0)^2} \\ &< \frac{k_0}{d} \left( \alpha + \frac{1}{2\pi\sigma^2} + \frac{1-\alpha}{1 - e^{-2\pi\sigma^2\alpha}} \right) e^{-\pi\sigma^2\alpha^2}. \end{align} \begin{align} \sum_{m \in \mathbb{Z}} \abs{ m \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} } &< 2\pi A \frac{\sigma}{\sqrt{d}} \left( \left( \sigma^2 + \frac{k_0}{d} \right) \alpha + \frac{1}{\pi} + \frac{k_0}{2\pi\sigma^2 d} + \frac{ \frac{1}{2\pi} + \frac{k_0}{d}(1-\alpha) }{1 - e^{-2\pi\sigma^2\alpha}} \right) e^{-\pi\sigma^2\alpha^2}. \end{align} \begin{align} \sum_{m \in \mathbb{Z}} \abs{ \tilde{\psi}(k_0;n+md)} &= \sum_{m \in \mathbb{Z}} A \frac{\sigma}{\sqrt{d}} e^{-\frac{\pi\sigma^2}{d^2}(n+md-n_0)^2}, \\ &= \sum_{m \in \mathbb{Z}} A \frac{\sigma}{\sqrt{d}} e^{-\pi\sigma^2 \left( m + \frac{n-n_0}{d} \right)^2} \\ &< \frac{A\sigma}{\sqrt{d}} \frac{2 e^{-\pi \sigma^2 \alpha^2}}{1 - e^{-2\pi \sigma^2 \alpha}}. \end{align} \begin{align} \sum_{m \in \mathbb{Z}/\{0\}} \abs{ \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l+md} } &= 2\pi A \sum_{m \in \mathbb{Z}/\{0\}} e^{-\frac{\pi d^2}{\sigma^2} \left( m + \frac{l-k_0}{d} \right)^2} \abs{ \frac{l+md-k_0}{\sigma^2} + i \frac{n_0}{d} } \\ &= 4\pi A \left( \frac{1}{2\pi d} + \frac{\frac{n_0}{d}}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}}. \end{align} \begin{align} \sum_{l \geq d} \abs{ \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l} } &= 2\pi A \sum_{l \geq d} e^{-\frac{\pi}{\sigma^2}(l-k_0)^2} \abs{ \frac{l-k_0}{\sigma^2} + i \frac{n_0}{d} } \\ &= 4\pi A \left( \frac{d}{2\pi} + \frac{\frac{n_0}{d}}{1 - e^{-\frac{\pi (d-k_0)}{\sigma^2}}} \right) e^{-\frac{\pi (d-k_0)^2}{4\sigma^2}} \end{align} \begin{align} \sum_{l \in \mathcal{S}_d(k_0)} &\sum_{m\in\mathbb{Z}/\{0\}} \psi(k_0;l+md) \ket{\theta_l} + \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m\in\mathbb{Z}/\{0\}} (l+md) \frac{d}{dx} \psi(k_0;x)\bigg|_{x=l+md} \ket{\theta_l} \\ &= \left( 1 - 2\pi \left( \frac{(x-k_0)^2}{\sigma^2} + \frac{k_0}{\sigma^2}(x-k_0) - i \frac{n_0}{d} (x-k_0) - i \frac{n_0}{d}k_0 \right) \right) e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} \bigg|_{x=l+md}. \end{align} \begin{align} \ket{\epsilon_6^{co}} &= -d \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=0}^{d-1} \sum_{m \in \mathbb{Z}} \frac{2\pi}{T_0} m \tilde{\psi}(k_0;n+md) \frac{e^{+i2\pi nl/d}}{\sqrt{d}} \ket{\theta_l}, \\ &= -d \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=0}^{d-1} \sum_{m \in \mathbb{Z}} \frac{2\pi}{T_0} m A \frac{\sigma}{\sqrt{d}} e^{-\frac{\pi \sigma^2}{d^2}(n+md-n_0)^2} phase \frac{e^{+i2\pi nl/d}}{\sqrt{d}} \ket{\theta_l} \\ \ltwo{\ket{\epsilon_6^{co}}} &< \frac{2\pi A \sigma d^2}{T_0} 2 \sum_{m=1}^\infty m e^{-\pi \sigma^2 \left( m - \frac{1}{2} - n_0/d \right)^2} \end{align} \fi \section{Introduction} \input{introduction.v8 \section{Informal overview of results and historical context}\label{summery} \subsection{The idealized quantum clock and control}\label{sec:ideal} We begin by reviewing the understanding of a time operator in quantum mechanics. Wolfgang Pauli \cite{pauli1,pauli2} argued that if there exists an ideal clock with Hamiltonian $\hat H$ and ideal observable of time $\hat{t}$; both self-adjoint on some suitably defined domains, then in the Heisenberg picture the pair must obey \begin{equation}\label{eq:div hat t} \frac{d}{dt} \hat t(t)= {\mathbbm{1}},\quad \forall\, t \in {\mathbbm{R}}, \end{equation} which in turn implies the canonical commutation relation $-\mathrm{i} [\hat t(t), \hat H]={\mathbbm{1}}, \quad \forall t\in {\mathbbm{R}}$ (we use units so that $\hbar =1$) on some suitably defined domain. Pauli further argued that the only pair of such operators (up to unitary equivalence) are $\hat t=\hat x$ and $\hat H=\hat p$, where $\hat x$, $\hat p$ are the canonically conjugate position-momentum operators of a free particle in one dimension. However, all such representations of $\hat p$ for which Eq. \eqref{eq:div hat t} is satisfied have spectra unbounded from below. Pauli thus concluded that no perfect time operator exists in quantum mechanics since such clock Hamiltonians would require infinite energy to construct due to the lack of a ground state. Later, other authors pointed out that Pauli in his analysis had only considered generators of Weyl pairs \cite{wong}. Alternatively, one could demand the weaker condition that the time operator measures time within a finite time interval only, thus replacing $\forall t \in {\mathbbm{R}}$ with $\forall\, t \in [0,t_\textup{max}]$ in Eq. \eqref{eq:div hat t}. Under such conditions, $\hat t$ and $\hat H$ can form Heisenberg pairs and a time operator $\hat t$ with spectrum $[0,t_\textup{max}]$ with a bounded from below Hamiltonian $\hat H$ exists, however the restriction of the spectrum of $\hat t$ to a finite interval requires \textit{confinement}, that is, the probability of finding the wave-function at the boundaries $0$, $t_\textup{max}$ must vanish \cite{wong}. While mathematically such a construction exists, it is conceivable that such confinement can only physically be achieved when one has an infinite potential at the boundaries of the interval $[0,t_\textup{max}]$ in order to prevent quantum mechanical tunnelling, which again, would require infinite energy and is therefore unphysical (e.g. particle in a box \cite{reed1975methods}). The question of whether a physically realizable perfect time operator exists in quantum mechanics is still a highly contentious issue. We will not dwell on this issue here, but rather; since we wish to construct a finite dimensional quantum clock that is capable of autonomous control, requiring finite energy, that imitates these ideal clocks, review here the relevant properties of these clocks that render them ideal as a quantum control system. For the simple example in which $\hat t=\hat x_c$, $\hat H=\hat p_c$ are position and momentum operators respectively, of a free particle in one dimension with domain $D_0$ of infinitely differentiable functions of compact support on $L^2({\mathbbm{R}})$, the dynamics are easily solvable. We refer to it as the \textit{idealised clock} and assume domain $D_0$ for all operators in this Section. The idealised clock allows for precise timing of events; something which is essential for our main objective, quantum autonomous control. The relevant properties of the idealised clock are: \begin{itemize} \item [1)] a distinguishable basis of time states, \item [2)] continuity in time evolution, and \item [3)] continuous autonomous control. \end{itemize} To be more precise, given that the Hamiltonian of the clock is $\hat{H} = \hat{p}_c$, the generalised eigenvectors of the position operator $\ket x$ are a \emph{distinguishable basis of time states} by which we mean $\braket{x|x^\prime} = \delta(x-x^\prime)$, and that given any initial generalized eigenvector of the position operator $\ket{x}$, the natural evolution of the clock will run through all of the generalised states in $\{\ket{x'} |\, x'>x\}$, \begin{equation} e^{-i\hat{H}_c t} \ket{x} = \ket{x+t}. \end{equation} Importantly, the property of shifting regularly w.r.t. time is true for any state of the clock, \begin{equation}\label{idealregularity_main} \braket{x|e^{-i\hat{H}_c t}| \Psi} = \braket{x-t|\Psi} \end{equation} The fact that this statement holds for all $x\in{\mathbbm{R}}$ and arbitrarily small $t\in{\mathbbm{R}}$ is the second property of the ideal clock, its \emph{continuity}. It implies that the clock's ability to tell time, for instance via a measurement of its position, remains the same for arbitrary intervals. The third property is the clock's ability to control an interaction on an external system. If one adds a position-dependent potential to the clock, it still retains its continuity, and only changes by a local phase that depends on the potential, \begin{equation}\label{idealpotential} \braket{x | \mathrm{e}^{-\mathrm{i} \left(\hat{H}_c + V(\hat{x}_c)\right) t} | \Psi} = \mathrm{e}^{-\mathrm{i}\int_{x-t}^x V(x^\prime) dx^\prime} \braket{x-t | \Psi},\qua x,t\in{\mathbbm{R}}, \,V\in D_0. \end{equation} A detailed proof of the above statements may be found in Section \ref{idealizedclock}. From Eq. \eqref{idealpotential} it follows that one can use the clock to autonomously control an additional quantum system. This is why this property is so important. It will be discussed in more detail in Section \ref{sec:Consequences of Quasi-Autonomous control}. \subsection{The finite clock}\label{sec:The finite clock main text} To construct a finite clock, as in \cite{SaleckerWigner,Peres}, we use a quantum system of dimension $d$ whose energy levels, denoted by $\{\ket{E_n}\}_{n=0}^{d-1}$, are equally spaced, \begin{equation}\label{finiteHamiltonian} \hat{H}_c = \sum_{n=0}^{d-1} n \omega \ketbra{E_n}{E_n}. \end{equation} The frequency $\omega$ determines both the energy spacing as well as the time of recurrence of the clock, $T_0 = 2\pi/\omega$, as $e^{-i \hat{H}_c T_0} = {\mathbbm{1}}_c$. This system possesses a distinguishable basis of time states $\left\{\ket{\theta_k}\right\}_{k=0}^{d-1}$ , mutually unbiased w.r.t. the energy states, \begin{align}\label{finitetimestates_main} \ket{\theta_k} &= \frac{1}{\sqrt{d}} \sum_{n=0}^{d-1} e^{-i2\pi n k/d} \ket{E_n}. \end{align} It will also be useful to extend the range of $k$ to ${\mathbbm{Z}}$. Extending the range of $k$ in Eq. \eqref{finitetimestates_main} it follows $\ket{\theta_k}=\ket{\theta_{k \textup{ mod. } d}}$ for $k\in{\mathbbm{Z}}$. The time states rotate into each other in regular time intervals of $T_0/d$, i.e. $e^{-i \hat{H}_c T_0/d} \ket{\theta_k} = \ket{\theta_{k+1}}$. Since they also form an orthonormal basis, this property is true for any state $\ket{\Psi}$ of the clock Hilbert space, \begin{equation}\label{finiteregularity_main} \braket{\theta_k | e^{-i \hat{H}_c m\, T_0 /d} |\Psi} = \braket{\theta_{k-m} | \Psi},\quad k,m\in{\mathbbm{Z}}. \end{equation} This is reflective of the idealised case Eq. \eqref{idealregularity_main}, but the key difference is that for the finite clock, Eq. \eqref{finiteregularity_main} only holds at \emph{regular} intervals ($m \in \mathbb{Z}$). Thus while every state of the clock is regular with respect to time, it is not continuous [property 2)]. In fact the time-states $\ket{\theta_k}$ themselves are considerably discontinuous. As a time-state evolves, it spreads out considerably in the time basis for non-integer intervals \cite{Peres,sergedft}. Additionally, the time-states fail to approximate the ideal commutator relationship between a (suitably defined) time operator and Hamiltonian in any limit \cite{Peres}. See Section \ref{SWPclock} for details on the behaviour of the time-states $\ket{\theta_k}$. For our work, rather than the time-state, we use a Gaussian superposition of time-states, \begin{align}\label{gaussianclock_main} \ket{\Psi_\textup{nor}(k_0)} &= \sum_{\mathclap{\substack{k\in \mathcal{S}_d(k_0)}}} \psi_\textup{nor}(k_0;k) \ket{\theta_k}, \\ \text{where} \quad \psi_\textup{nor}(k_0;x) &= A e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} e^{i 2\pi n_0(x-k_0)/d}, \end{align} with $A$ a normalisation constant and $k_0\in{\mathbbm{R}}$. The definition and basic properties of the state are detailed in the Section \ref{sec:Definition of Gaussian clock}. Briefly, $\sigma$ denotes the width of the state in the time basis, ranging from $\sigma\approx 0$ (approximately a time-eigenstate) to $\sigma \approx d$ (almost an energy eigenstate). $\omega \,n_0$ corresponds to the mean energy of the state, and ranges between $0$ and $\omega(d-1)$. $\mathcal{S}_d(k_0)$ is a set of $d$ consecutive integers centered about $k_0$, such that the set $\{ \ket{\theta_k}: k\in \mathcal{S}_d(k_0) \}$ forms a complete orthonormal basis for all states in the $d$ dimensional Hilbert space of the clock \subsection{RESULTS (Informal Overview)}\label{sec:RESULTS} We now briefly state our main results, reserving the full theorems for later. For simplicity we state all the results for the simplest case, $\sigma = \sqrt{d}$, $n_0 = (d-1)/2$, use $poly(d)$ to refer to polynomial in $d$ and use $\mathcal{O}$ for Big-O notation. We leave the more general case discussed in Section \ref{sec:discussion} for later Sections. \subsubsection{Quasi-Continuity} Our first result is to recover the \emph{continuity} of a quantum clock for the class of Gaussian superpositions of time-states introduced above. \begin{result}[See Theorem \ref{gaussiancontinuity} for the most general version] If the clock begins in a Gaussian superposition, $\braket{\theta_k|\Psi_\textup{nor}(k_0)} = \psi_\textup{nor} (k_0;k)$, then for all $k\in\mathcal{S}_d(k_0+t \,d/T_0)$ and $t\in{\mathbbm{R}}$, \begin{align}\label{eq:result 1 main text} \begin{split} \braket{\theta_k|e^{-\mathrm{i} t \hat H_c}|\Psi_\textup{nor}(k_0)} &= \psi_\textup{nor}(k_0 - d \frac{t}{T_0} ; k) + \varepsilon_c(t,d),\\ \varepsilon_c(t,d) &= \mathcal{O} \left( t\; poly(d)\; e^{-\frac{\pi}{4} d} \right) \,\textup{in the } d\rightarrow \infty \textup{ limit}. \end{split} \end{align} In other words, the evolution of the clock state, which is composed of the discrete coefficients $\braket{\theta_k|e^{-\mathrm{i} t \hat H_c}|\Psi_\textup{nor}(k_0)}$, may be approximated by the continuous movement of the background Gaussian function $\psi_\textup{nor}(k_0;x)$. [see a) Fig. \ref{fig:clocks main text}]. This is exactly as in the case of the idealized clock Eq. \eqref{idealregularity_main}, and the error in the approximation grows linearly in time, but is exponentially small with clock dimension. This is in marked contrast to using the time states, which only behave as good clock states at a finite number of intervals. \end{result} \subsubsection{Autonomous Quasi-Control} For the clock to serve as a quantum control unit, it must be continuous not only under its own evolution, but also in the presence of an appropriate potential, as in the case of the idealized clock \eqref{idealpotential}. Our second major result is to prove the same for the finite clock. \begin{result}[See Theorem \ref{movig through finite time} for the most general version] Let $V_0: {\mathbbm{R}}\rightarrow {\mathbbm{R}}$ be an infinitely differentiable, periodic function of one variable with period $2\pi$, normalized so that its integral over a period is $\Omega$, with $-\pi\leq \Omega<\pi$. For the finite clock of dimension $d$, construct a potential from $V_0$ as follows, \begin{align} \hat{V}_d &= \frac{d}{T_0} \sum_{k=0}^{d-1} {V}_d(k) \ketbra{\theta_k}{\theta_k}, \quad V_d(x) &= \frac{2\pi}{d} V_0 \left( \frac{2\pi x}{d} \right). \end{align} Then under the Hamiltonian $\hat H_c + \hat V_d$, if the clock begins in a Gaussian superposition $\braket{\theta_k|\Psi_\textup{nor}(k_0)} = \psi_\textup{nor} (k_0;k)$, then for all $k\in\mathcal{S}_d(k_0+t \,d/T_0)$ and $t\in{\mathbbm{R}}$, \begin{align}\label{eq:result 2 main text} \begin{split} \braket{\theta_k|e^{-\mathrm{i} t ( \hat H_c + \hat V_d )}|\Psi_\textup{nor}(k_0)} =&\,\mathrm{e}^{-\mathrm{i} \int_{k-d\,t/T_0}^k V_d(x)dx}\, \psi_\textup{nor}(k_0 - d \frac{t}{T_0};k) + \varepsilon_v(t,d), \\\;\;\;\;\; \varepsilon_v(t,d) =&\, \mathcal{O}\left( t\; poly(d)\; e^{-\frac{\pi}{4} \frac{d}{\zeta}} \right)\,\textup{in the } d\rightarrow \infty \textup{ limit}, \end{split} \end{align} where $\zeta \geq 1$ is a measure of the size of the derivatives of $V_0(x)$, \begin{align}\label{eq:b zeta main text} \begin{split} \zeta &=\left( 1+\frac{0.792\, \pi}{\ln(\pi d)}b \right)^2,\quad\text{for any } \\ b &\geq\; \sup_{k\in{\mathbbm{N}}^+}\left(2\max_{x\in[0,2\pi]} \left| V_0^{(k-1)}(x) \right|\,\right)^{1/k}, \end{split} \end{align} where $ V_0^{(k)}(x)$ is the $k^\textup{th}$ derivative with respect to $x$ of $ V_0(x)$, so roughly speaking - the larger the derivatives, the larger is $\zeta$. [see b) Fig. \ref{fig:clocks main text}]. We further require the derivatives $V_0$ to be such that the lower bound on $b$ in Eq. \eqref{eq:b zeta main text} is finite in order for Eq. \eqref{eq:result 2 main text} to be non-trivial. As in the case of the idealized clock, this result allows one to implement a unitary on an external system, or any time-dependent interaction that commutes with the Hamiltonian of the system. We discuss this in greater detail in a following Section, and bound the errors in the system and the clock at the end of such processes (in comparison to the ideal case). In particular, we will show that the state of the clock is disturbed by only an exponentially small amount in the dimension of the clock. In further work \cite{RMRenatoetal}, it is demonstrated that the result above also encapsulates continuous weak measurements on the state of the clock, that may be used to extract temporal information from the clock (i.e. to measure time). \begin{figure} \begin{center} \includegraphics[scale=0.2]{clockAndalarm2.jpg \caption{ \textbf{Resemblance of a quantum clock to a classical clock.} \textbf{a)} Illustration of main result 1. A quantum clock state $\ket{\Psi_\textup{nor}(12)}$ initially centered at ``12 O'clock" time evolved to $t=3 T_0/(2 d)$. This clock state will move continuously around the clock face, maintaining localization as a wave packet. It's this maintenance of localization which means it is well distinguishable from other states at all times. \textbf{b)} Illustration of main result 2. Now a potential peaked around ``6 O'Clock" is introduced. The clock dynamics will be, up to a small error the same as in a) until a time just before ``6 O'Clock", at which point the clock will start to acquire a phase due to its passage over the potential. At a time just after ``6 O'Clock", the dynamics will start to resemble those of a) again, but now with a global phase added. A direct consequence of main result 2, is that one can use this time dependent phase to control an auxiliary quantum system at specified time, resembling an alarm clock. }\label{fig:clocks main text} \end{center} \end{figure} \end{result} \subsubsection{Quasi-Canonical Commutation}\label{sec:Quasi-Canonical commutator} So far we have demonstrated that the finite dimensional Gaussian clock state, Eq. \eqref{gaussianclock_main}, satisfies properties 1) to 3) up to a small error. Recall that these desired properties were a direct consequence of Eq. \eqref{eq:div hat t} which is satisfied if $[\hat t,\hat H_c]=\mathrm{i} $ on some appropriately defined domain. We will now turn to the issue of the commutator. For the finite clock, since the $\ket{\theta_k}$ rotate into each other in regular time intervals, an intuitive definition of a time operator is in the eigenbasis of $\ket{\theta_k}$. Therefore, in analogy to the energy operator, $\hat H_c$, we define the time operato \begin{equation}\label{eq:t op main text def} \hat{t}_c = \sum_{k=0}^{d-1} k \frac{T_0}{d} \ketbra{\theta_k}{\theta_k}. \end{equation} However, as noted by Peres \cite{Peres}, this operator cannot obey $[\hat t_c, \hat H_c]= \mathrm{i}$ for any time state $\ket{\theta_k}$. In fact, \begin{equation} \bra{\theta_k} [\hat t_c, \hat H_c] \ket{\theta_k} = 0 \quad\text{for all } k\in{\mathbbm{Z}} \text{ and } d\in{\mathbbm{N}}^+. \end{equation} This is intrinsically related to the time-states not being good clocks themselves. On the contrary, for a Gaussian superposition of clock states, \begin{result}[See Theorem \ref{Quasi-Canonical commutation} for the most general version] \begin{equation}\label{eq:main result 3 amin text} \begin{split} [\hat t_c,\hat H_c]\ket{\Psi_\textup{nor}(k_0)}&=\mathrm{i} \ket{\Psi_\textup{nor}(k_0)} +\ket{\varepsilon_\textup{comm}},\\ \|\ket{\varepsilon_\textup{comm}}\|_2 &=\mathcal{O}\left(poly(d)\; \mathrm{e}^{-\frac{\pi}{4}d} \right)\,\textup{in the } d\rightarrow \infty \textup{ limit}. \end{split} \end{equation} The implication of Eq. \eqref{eq:main result 3 amin text} is that our Gaussian clock states $\ket{\Psi(k_0)}$ can achieve the canonical commutation relation up to a exponentially small error in clock dimension. In addition, since the l.h.s. of Eq. \eqref{eq:main result 3 amin text} is $T_0$ independent, so is the error $\|\ket{\varepsilon_\textup{comm}}\|_2$ \end{result} \subsubsection{Consequences of Autonomous Quasi-control}\label{sec:Consequences of Quasi-Autonomous control} We now discuss the important consequences that Result 2 has for quantum control. Consider a quantum system of dimension $d_s$ for which we wish to implement a unitary $\tilde U$ over a time interval $t\in[t_1,t_2]$ for some initial state $\rho_s\in\mathcal{S}(\mathcal{H}_s)$, \begin{equation}\label{eq:rho initial final} \rho_s(t)= \begin{cases} \rho_s &\mbox{ if } 0\leq t< t_1\\ \tilde U \rho_s \tilde U^\dag &\mbox{ if } t> t_2. \end{cases} \end{equation} Such an operation could represent, for example, the implementation of a quantum gate in a quantum computer. This can be implemented via the time dependent Hamiltonian $\hat H(t)=\hat H_s^{int} g(t)$ acting on $\mathcal{H}_s,$ where $\tilde U=\mathrm{e}^{\mathrm{i} \hat H^{int}_s},$ and $g\in L({\mathbbm{R}}: {\mathbbm{R}}_{\geq 0})$ is a normalised pulse, $\int_{t_1}^{t_2} dx\, g(x)=1,$ with support $[t_1,t_2]$. Indeed, the Schr\" odinger equation has the solution $\rho_s(t)=U(t)\rho_s U^\dag(t)$, where \begin{equation}\label{eq:t dependent ham} U(t)=\mathrm{e}^{-\mathrm{i} \hat H_s^{int} \int_0^t dx g(x)}, \end{equation} thus implementing Eq. \eqref{eq:rho initial final}. However, since $\hat H(t)$ is a time dependent Hamiltonian, it will require external control for its implementation, and thus is not autonomous. We now show that a direct consequence of the autonomous quasi-control result, is that the unitary $\tilde U$ can be implemented autonomously with the aid of the clock, while only incurring a small error due to the finite nature of the clock. To achieve this, we show in Section \ref{sec:Implementing Energy preserving unitaries with the finite} by explicit construction, that there exists a $V_0$ such that for all unitaries $\tilde U$, initial states $\rho_s$, and time intervals $t\in[0,t_1]\cup [t_2,T_0]$, for all $0<t_1<t_2<T_0$, the evolution of the initial state $\rho_{sc}'(0)=\rho_s\otimes\ketbra{\Psi_\textup{nor}(0)}{\Psi_\textup{nor}(0)}$ under the time \textit{independent} Hamiltonian $\hat H_{sc} = {\mathbbm{1}}_s \otimes \hat{H}_c + \hat H^{int}_s \otimes \hat{V}_d,$ denoted by $\rho_{sc}'(t)=\mathrm{e}^{-\mathrm{i} t \hat H_{sc}}\, \rho_{sc}'(0) \,\mathrm{e}^{\mathrm{i} t \hat H_{sc}}$ satisfies \begin{equation}\label{eq:q-auto contrl trace bound} \| \rho_s(t)-\rho_s'(t)\|_1\leq \sqrt{d_s\textup{tr}[\rho_s^2]} \,\left(\varepsilon_s(t,d)+2|\varepsilon_v(t,d)|(1+|\varepsilon_v(t,d)|)\right), \end{equation} where $\rho_s'(t)=\textup{tr}_c[\rho_{sc}'(t)]$ denotes the partial trace over the clock, $\varepsilon_s\geq 0$ is $\rho_s$ and $d_s$ independent. The explicit form of $\varepsilon_s$ specifying how it depends on the potential in general, is provided in the Section \ref{sec:Clocks as Quantum control}. Depending on the choice of the potential $V_0,$ $\varepsilon_s$ will have different decay rates with $d$. Moreover, such a choice will also influence how disturbed the motion of the clock is, due to the potential's back reaction on the clock. As shown in Section \ref{sec:Automation via the idealised clock}, the idealised clock is never disturbed due to back action of the potential, and the unitary $\tilde U$, is always implemented perfectly. Before discussing Eq. \eqref{eq:q-auto contrl trace bound} in more detail, we will first introduce a dynamical measure of disturbance for the clock. Since the clock state $\rho_c'(t)=\textup{tr}_c[\rho_{sc}'(t)]$ undergoes periodic dynamics with periodicity $T_0$ when no unitary is implemented, i.e. when $V_0=0$ for all $x\in{\mathbbm{R}}$, the difference in trace distance between the initial and final state after one period is exactly zero. Moreover, when $V_0\neq 0$ any difference between the two states is \textit{solely} due to the back reaction caused by the potential implementing the unitary on the system. A simple application of Result 2, allows for a direct characterization of this disturbance, \begin{equation}\label{eq:clock finite trace diff} \frac{1}{2}\|\rho_c'(0)-\rho_c'(T_0)\|_1\leq |\varepsilon_v(T_0,d)| \end{equation} where $\rho_c'(t)=\textup{tr}_s[\rho_{sc}'(t)]$ Eqs. \eqref{eq:q-auto contrl trace bound}, \eqref{eq:clock finite trace diff}, represent a trade-off as exemplified in Fig. \ref{fig:tradeoff}. Depending on how one parametrizes the potential $V_0$ with $d$, the decay rates of $\varepsilon_v$ and $\varepsilon_s$ will be different. We will highlight here two extremal cases. First case is that of minimal clock disturbance. This corresponds to when we allow for an arbitrarily small tolerance error $\varepsilon_s$ which is $d$ independent. This choice corresponds to cases in which $V_0$ is chosen to be $d$ independent from which it follows that $b$ is constant and $\lim_{d\rightarrow \infty}\zeta=1$, meaning that the clock disturbance $\varepsilon_v$ is minimal and asymptotically equal to that given by the quasi-continuity bound $\varepsilon_c$. More generally, we show that one can even achieve a parametrisation such that $\lim_{d\rightarrow \infty}\varepsilon_s=0$ while still maintaining $\lim_{d\rightarrow \infty} \zeta=1$. Alternatively, one can choose a potential with the aim of minimizing the r.h.s. of Eq. \eqref{eq:q-auto contrl trace bound}. This leads to both $\|\rho_s(t)-\rho_s'(t)\|_1$ and $\|\rho_c'(0)-\rho_c'(T_0)\|_1$ to be of the same order and decays faster than any power of $d$, specifically, exponential decay in $d^{1/4}\sqrt{\ln d}$. An alternative interpretation of our results is that the clock Hamiltonian is that of a quantum Harmonic oscillator with energy spacing $\omega$, and we are working in a subspace of it. In this interpretation, in the error terms $\varepsilon_c$, $\varepsilon_v$, $\varepsilon_s$, one could make the substitution $d\mapsto \frac{2}{\omega} \langle \hat H_c \rangle$, where $\langle \hat H_c \rangle$ is the mean energy of the clock's initial state, $\ket{\Psi_\textup{nor}(k_0)}$. In this case, we can interpret the clock as living in an infinite dimensional Hilbert space and with error terms $\varepsilon_c$, $\varepsilon_v$ which are exponentially small in mean clock energy. \begin{figure}[!htb] \includegraphics[scale=0.32 {PotsV6.png \caption{Plot of the potential $V_0(x)=A_c \cos^{2n}\left(\frac{x-\pi}{2} \right)$ over one period $[0,2\pi]$, with $A_c$ a normalisation constant. Blue $n=10$. Orange $n=100$. The blue potential is ``flatter" than the orange potential, and as such, $b$ in Eq. \eqref{eq:b zeta main text} is smaller than in the case of the orange potential. The disturbance $\varepsilon_v$ in the clock's dynamics caused by the implementation of the unitary (see Eq. \eqref{eq:clock finite trace diff}), will be smaller for the blue potential than for the orange potential. However, the error term $\varepsilon_s$ involved in implementing the system unitary (see Eq. \eqref{eq:q-auto contrl trace bound}), will be larger for the blue potential than for the orange potential. Generally speaking, whenever $b$ is larger, $\varepsilon_s$ can be made smaller at the expense of a larger disturbance to the clock's dynamics (i.e. larger $\varepsilon_v$). This trade off is reminiscent of the information gain disturbance principle \cite{info_disturbace_Paul}, but here rather than gaining information, the unitary is implemented more accurately (i.e. smaller r.h.s. in Eq. \eqref{eq:q-auto contrl trace bound}) }\label{fig:tradeoff} \end{figure} \subsection{Discussion}\label{sec:discussion} So far, the results stated have only been for when the initial Gaussian superposition of time states $\ket{\Psi_\textup{nor}(k_0)}$ has a width $\sigma$ equal to $\sqrt{d}$ and mean energy centered in the middle of the spectrum of $\hat H_c$. This particular choice of the width which we call \textit{symmetric}, has a particular physical significance. The uncertainty in both the energy and time basis, denoted $\Delta E$ and $\Delta t$ respectively, are equal with $\Delta E\Delta t=1/2$.\footnote{up to an additive correction term which decays exponentially in $d$.\label{refnote}} The last equality is true regardless of the width $\sigma$, implying that all our Gaussian clock states are minimum uncertainty states.\footref{refnote} However, whenever $\sigma >\sqrt{d}$, we have $\Delta E < \Delta t$ implying less uncertainty in energy. We call these \textit{energy squeezed} states in analogy with quantum optics terminology. Meanwhile, we call states for which $\sigma <\sqrt{d}$, \textit{time squeezed} since $\Delta E > \Delta t$ in this case. It is likely that only time squeezed states can achieve the Heisenberg limit \cite{HesenbergLim}. Yet, from the viewpoint of quantum control, as we shall now see, our results suggest that squeezed states (either in the time or energy basis) have larger errors $\varepsilon_c$ and $\varepsilon_v$, thus rendering them more fragile to back-reactions from the potential $V_0$. Full details are left to later sections, but qualitatively the behaviour is as follows. Whenever the initial clock state is energy or time squeezed, the error terms $\varepsilon_c$, $\varepsilon_v$ still maintain their linear scaling with time, but lose their exponential decay in clock dimension. Moreover, it is now exponential decay in $d^\eta$, $0<\eta<1$, with decreasing $\eta$ as squeezing in time or energy is increased. We conjecture that the exponential decay of $\varepsilon_c$ in the clock dimension is the best possible scaling with clock dimension which holds for all times in one time period $[0,T_0]$. Such a fundamental limitation, will have direct implications for how much work can be embezzled from the clock when it implements a unitary in the framework of \cite{second}. The second question of interest is how does the mean energy of the initial state of the clock effect the errors induced by finite size? From the energy-time uncertainty relation \cite{time_energy_PBusch}, and indeed the idealised clock, one might expect the larger the mean energy of the state, the better it performs. Contrarily, we find that errors maintain the same exponential decay in $d^\eta$ with $\eta=1$ reserved for symmetric states, but now with a smaller prefactor, -- one has to replace the $\pi/4$ in Eq. \eqref{eq:result 1 main text} with a factor which approaches zero as the mean energy of the initial clock state approaches either end of the spectrum of $\hat H_c$. This suggests that, when the dimension is finite, it is the \textit{dimension} itself, rather than the energy of the state, as suggested by the energy-time uncertainty relation \cite{time_energy_PBusch}, which is the resource for improving the accuracy of a clock.\\ On another level, by demonstrating that up to small errors, the three core paradigms in quantum control, -- the unitary (Eq. \eqref{eq:rho initial final}), the time dependent Hamiltonian (Eq. \eqref{eq:t dependent ham}), and time independent control (Eq. \eqref{eq:q-auto contrl trace bound}), are all equivalent up to small errors, our results represent a unification of the three paradigms. The implications of this are not only of a foundational nature, but also practical, since it implies that one can numerically simulate a time dependent Hamiltonian on $\mathcal{H}_s$, while being re-assured that in actual fact, the results of the simulation are equivalent to simulating a time independent Hamiltonian on the larger Hilbert space $\mathcal{H}_s\otimes\mathcal{H}_c$, which would be numerically intractable due to the increased dimension.\\ In the case that $\rho_s$ has a non-trivial Hamiltonian $\hat H_s$ and thus is not stationary in time, the described protocol for implementing a unitary $\tilde U$, now only applies to the case of energy preserving unitaries, namely $[\tilde U, \hat H_s]=0$. For non communing unitaries, one would additionally need a source of energy in addition to the source of timing provided by the clock. The energy source can be modeled in our setup explicitly by including a quantum battery system $\mathcal{H}_b$ into the setup or taken to be the clock itself. In the former case, the clock would perform an energy preserving unitary over the $\mathcal{H}_s\otimes\mathcal{H}_b$ system. The battery Hamiltonian and initial state would be such that locally, on $\mathcal{H}_s$, an arbitrary unitary would have been performed. Such a battery Hamiltonian and initial battery state has been studied in \cite{Aberg}, and will be applied to the setup in this paper in an upcoming paper. \subsection{Conclusions}\label{sec:conclusions} We solve the dynamics of a finite dimensional clock when the initial state is a coherent superposition of time states -- a basis which is mutually unbiased with respect to the energy eigenstates of the finite dimensional Hamiltonian. This can be used to measure the passage of time. We show that such superpositions evolve in time in ways which mimic idealised, infinite dimensional and energy clocks up to errors which decay exponentially fast in clock dimension. This represented a challenging problem due to its non-perturbative nature. We envisage that the techniques for solving this problem may be useful in solving other dynamical problems in many-body physics. We then demonstrate the consequences our results have for autonomous quantum control. We show that the clock can implement a timed unitary on the system via a joint clock-system time independent Hamiltonian, with an error which decays faster than any polynomial in the clock dimension; or equivalently, faster than any polynomial in the clock's mean energy. The implementation of the unitary induces a back-reaction onto the clock's dynamics. We discuss the trade-off between smaller clock disturbance and better temporal localization of the unitary's implementation, which our bounds address quantitatively. Our results single out states of equal uncertainty in time and energy, with a mean energy at the mid point of the energy spectrum, as being the most robust, and thus incurring the least disturbance due to its implementation of the unitary on the system. Such results are important, since the implementation of unitaries is a ubiquitous operation encountered in almost every research field of theoretical quantum mechanics; perhaps the most prominent example being in the field of quantum computation. Yet very little was known about how well such control, normally accounted for by a classical field, can actually be implemented by a fully quantum autonomous setup. Our results provide analytical and insightful bounds on exactly how well this can be achieved, venturing into the fundamental issue of time in quantum mechanics in the process. \\ \iffalse Finite dimensional clocks have shown interest in recent years in three areas of research: investigations on the detectability of the quantum nature of spacetime on length scales far larger than the Planck length (e.g. [26, 27]); studies of tunnelling times (e.g. [28]) and superluminal photon propagation through evanescent media [29]; and quantum information approaches to optimising quantum clock resolution [30] and synchronisation via nonlocal entangled systems (e.g. [31]). All of these questions and proposals are subject to ongoing controversial scrutiny, so that it is too early to attempt an assessment. -application to study a fully quantum version of a quantum quench?, cellular automota, quantum computation, time of arrival problem \fi \acknowledgments \vspace{-0.5cm} The authors would like to thank Sandu Popescu, Michael Berry, Michal Horodecki, Sandra R. Stupar, William D. Matthews and Renato Renner, Anthony Short, Artur Malabarba for helpful discussions. MW and JO acknowledge support from the EPSRC, Royal Society and the COST network (Action MP1209). RS acknowledges support from the Swiss National Science Foundation (grant PP00P2 138917 and QSIT). \iffalse --------\textbf{END}---------------------------------- .\\ .\\ .\\ .\\ -----------------------------------------------Random stuff beyond this point---------------------------------- By \emph{continuity}, For example, a particle with Hamiltonian $\hat{P}$ (momentum). The $\hat{x}$ (position) of the particle is a perfect indication of time, as in the Heisenberg picture, $d\hat{x}/dt = 1$. At the discrete time intervals $t=m*2\pi/(\omega d)$, the angle states are rotated perfectly, $\ket{\theta_k}\rightarrow\ket{\theta_{k+m}}$, but for intermediate times they spread out and their distinguishability drops considerably. Furthermore, one may verify that they are unable to approximate the ideal commutation relation Eq. \eqref{idealcommutator} for any dimension, as \begin{equation} \bra{\theta_k} [\hat{H},\hat{T}] \ket{\theta_k} = 0 \;\forall k. \end{equation} \subsection{Example: Landau errasure} \subsection{Punchlines} - A clock is used either to tell you when something happened, or to tell you to when to do something. - do local unitaries by local interaction terms - dismantle the "church of the higher hilbertspace dilema": a sequence of unitaries is normally thought of as being timed by a classical clock (i.e. the "clock on your wall int he lab") and being implemented by a classical field. This breaks the Quantum-classical divide, the so-called "church of the larger hilbertspace dilema". Our clock together with its ability to perform unitaries, proves a fundamental upper bound to the error to which this divide is not necessary. - it is not coincidence that the idealised clock also gives rise to perfect control \subsection{Introduction fragments} Attempts to define ``time of arrival operators"\cite{timeofarrival} suffer from the same problem. \subsection{Brainstorming ideas} On the other hand, point out the extensive work put into deriving the role of time (as an observable/operator) in quantum theory, and the conclusion that in fact, the ideal version is always like $\{x,p\}$, and thus unphysical (because unbounded -time of arrival derives this, check if others do the same as well). The idealized time operator (refs) and the Next point out that all of them suffer from either the use of the ideal unrealistic clock, or from vague arguments on clocks of finite dimension. - - Clocks and quantum control assume temporal information, that like every other form of information, requires work (Landauer's principle). - Interesting physics is lost in the infinite dimensional limit. - Many forays into quantum thermodynamics are an extension of classical laws into the quantum regime. For clocks however, there is a definite difference, since classical clocks are by definition perfect (non-interacting), and continuous. Recent work indicates that: - universe clocks seem to not care about coherence, but this is in the infinite d limit. - (thermal operations) possible embezzling is bounded by dimension - (alternate ticks) the amount of information about time is bounded by the dimension of the system - (autonomous clocks) coherent clock states outperform incoherent states of the same dimension - (autonomous clocks) there exists tradeoffs between power, accuracy and dimension - justification that the P Hamiltonian is the only ''perfect clock'' - 1) commutators imply that clock and time observable are x and p ''up to regularity/irregularity'' - 2) this is unbounded (this was observed by Pauli etc ref) - 3) therefore we want a clock that 1) mimics perfect clock 2) bounded from below (physical) [gap in reasoning: why not infinite dim but bounded from below?] whats more, this unphysical idealisation, hides lots of the interesting physics whic can only be captured by finite dimensinal clocks such at ... -4) now justify OUR clock by following the lines that Wigner guy does. -5) justification of Gaussian state distribution rather than initial clock state being only one angle state. (i.e. not Perez clock). The Perez clock is often referred to as a good approximation to ideal clocks. This is highly misleading, as although it presents a basis of orthogonal states through which it evolves regularly, in between these fixed time periods, the state is highly spread among ALL states, and thus the clock has absolutely no continuity, in the sense that a measurement of time at a random point would give a bad result, and any attempt to couple one angle state to an interaction in order to localize the interaction in time would fail. \section{Different levels of unitary description} \emph{The unitary picture. ---} Here one need only consider the state of the system, $\rho_s$, and an arbitary unitary operation $U$. The final state of the system is $\rho^\prime_s = U \rho U^\dagger$. The cost of the unitary is encoded in the system state. Thus if the system has a non-trivial Hamiltonian $H_s$, the energy cost of the unitary is $W = \Tr(H_s (\rho^\prime_s - \rho_s))$ (and similarly for the entropy change). \emph{The timed interaction picture. ---} Here the unitary is the result of an interaction Hamiltonian that is present only during a finite time interval. Since in general, the interaction Hamiltonian need not commute with the system Hamiltonian, the unitary is expressed as a time-ordered integral involving both Hamiltonians, \begin{equation}\label{timeorderinteraction} U = \mathcal{T} \left\{ e^{-i \int_{t_i}^{t_f} (H_s + H_{int}(t)) dt } \right\}, \end{equation} where $[t_i,t_f]$ is the time interval within which the interaction Hamiltonian is non-zero. The finite nature of the interaction allows one to account for the changes in energy and entropy via the state of the system outside the interaction interval, i.e. before $t_i$ and after $t_f$. \emph{The fully quantum/closed system/clock picture. ---} In order to eliminate the need for an external observer during the implementation of the unitary, one replaces the time-dependent interaction by a interaction Hamiltonian between the system in question and an additional quantum system, labeled a quantum clock. The clock also possesses its own Hamiltonian, specially chosen so that one of the degrees of freedom of the clock mimics the role of time, as seen from the following example of the ideal clock. \fi \iffalse \section{Definition of Gaussian clock states and properties}\label{sec:Definition of Gaussian clock} In this section we will introduce the class of clock states, that are Gaussian superpositions of the time-states, and review some of their properties. In the following, we call the Hilbert space of the clock, $\mathcal{H}_c$, the Hilbert space formed by the span of the time basis $\{\ket{\theta_k}\}_{k=0}^{d-1}$, or equivalently the energy basis $\{\ket{E_n}\}_{n=0}^{d-1}$. \begin{definition}\label{def:Gaussian clock states}\emph{(Gaussian clock states)}. Let $\Lambda_{\sigma,n_0}$ be the following space of states in the Hilbert space of the $d$ dimensional clock, \begin{equation}\label{eq:lambda set def} \Lambda_{\sigma,n_0}=\bigg\{ \ket{\Psi(k_0)}\in\mathcal{H}_{c},\quad k_0\in{\mathbbm{R}}\bigg\}, \end{equation} where \begin{align}\label{gaussianclock} \ket{\Psi(k_0)} = \sum_{\mathclap{\substack{k\in \mathcal{S}_d(k_0)}}} A e^{-\frac{\pi}{\sigma^2}(k-k_0)^2} e^{i 2\pi n_0(k-k_0)/d} \ket{\theta_k}, \end{align} with $\sigma \in (0,d)$, $n_0 \in (0,d-1)$, $A \in {\mathbbm{R}}^+$, and $\mathcal{S}_d(k_0)$ is the set of $d$ integers closest to $k_0$, defined as: \begin{align}\label{eq: mathcal S def} \mathcal{S}_d(k_0) = \left\{ k \; : \; k\in \mathbb{Z} \text{ and } -\frac{d}{2} \leq k_0-k < \frac{d}{2} \right\}. \end{align} In the special case that $\ket{\Psi(k_0)}$ is normalized, it will be denoted by \begin{equation}\label{eq:Psi ket normalised} \ket{\Psi_\textup{nor}(k_0)}=\ket{\Psi(k_0)}, \end{equation} and $A$ will take on the specific value \begin{equation}\label{eq:A normalised} A=A(\sigma;k_0)=\frac{1}{\sqrt{\sum_{k\in \mathcal{S}_d(k_0)} \mathrm{e}^{-\frac{2\pi}{\sigma^2}(k-k_0)^2}}}, \end{equation} s.t. $\braket{\Psi_\textup{nor}(k_0)|\Psi_\textup{nor}(k_0)}=1$. Bounds for $A$ can be found in Section \ref{Normalizing the clock state}. \end{definition} \begin{remark}[Technicality] Sometimes we will use big O notation $\mathcal{O}$, and $poly(x)$ to denote a generic polynomial in $x\in{\mathbbm{R}}$ of constant degree. When doing so, for simplicity, it will be assumed that $\sigma\in(0,d)$ is bounded away from $0$ for all $d\in {\mathbbm{N}}^+$, e.g. $1/\sigma$ is upper bounded for all $d$. \end{remark} \begin{remark}[Time, uncertainty, and energy of clock states] The parameters $k_0$ and $\sigma$ may be identified with the mean and variance respectively in the basis of time-states. $n_0$ may be identified with the mean energy of the clock, as we shall discuss shortly in Remark \ref{rem:symme of clock} which motivates Def. \ref{def:clock stat classes}. \end{remark} \begin{remark}[Keeping the clock state centered]\label{clockcentered} The choice of the set $\mathcal{S}_d(k_0)$ is to ensure that the Gaussian is always centered in the chosen basis of angle states. The reason it is possible to do this in the first place is because the basis of clock states is invariant under a translation of $d$, i.e. $\ket{\theta_k} = \ket{\theta_{k+d}}$ \eqref{finitetimestates}. Instead of the set $\{0,1,...,d-1\}$ we can choose to express the state w.r.t. to any set of $d$ consecutive integers, and we choose the specific set in which the state above is centered. If $d$ is even, then $\mathcal{S}_d(k_0)$ changes at integer $k_0$, if $d$ is odd, then it changes at half-integer values of $k_0$. Equivalently, in terms of the floor function $\lfloor \cdot \rfloor$, \begin{equation}\label{eq:S set def} \mathcal{S}_d(k_0) = \begin{cases} \mathcal{S}_d(\lfloor k_0 \rfloor ) & \text{if $d$ is even}, \\ \mathcal{S}_d(\lfloor k_0 + 1/2 \rfloor) & \text{if $d$ is odd}. \end{cases} \end{equation} \end{remark} \begin{definition}\label{Distance of the mean energy from the edge}(Distance of the mean energy from the edge of the spectrum) We define the parameter $\alpha_0\in(0,1]$ as a measure of how close $n_0\in(0,d-1)$ is to the edge of the energy spectrum, \iffalse \begin{equation} \alpha = \min \left\{ \frac{n_0+1}{d}, 1 - \frac{n_0}{d} \right\} = \min_{n \in \{0,1,...,d-1\}} 1 - \frac{|n-n_0|}{d}. \end{equation} The maximum value of $\alpha$ is $(d+1)/2d$ corresponding to $n_0 = (d-1)/2$. The minimum value is $\alpha = 1/d$, when $n_0=0$ or $n_0 = d-1$. c.f. similar measures of Eq. \eqref{eq:alpha_c def eq}, and Eq. \eqref{eq:alpha_0 def} in Def. \ref{def:stand alone def mathcal N}. \fi namely \begin{align}\label{eq:alpha_0 def} \alpha_0&=\left(\frac{2}{d-1}\right) \min\{n_0,(d-1)-n_0\}\\ &=1-\left|1-n_0\,\left(\frac{2}{d-1}\right)\right|\in(0,1]. \end{align} The maximum value $\alpha_0=1$ is obtained for $n_0=(d-1)/2$ when the mean energy is at the mid point of the energy spectrum, while $\alpha_0\rightarrow 0$ as $n_0$ approaches the edge values $0$ or $d-1$. c.f. similar measure, Eq. \eqref{eq:alpha_c def eq}. \end{definition} \begin{remark}[Comparison to time-states] As a first comparison, we repeat the analysis of the behaviour of the time-states (see Fig. \ref{peresbehaviour}), this time for the Gaussian state, i.e. we plot the expectation value and variance of the time operator, given the Gaussian initial state centered about $\ket{\theta_0}$, see Fig. \ref{gaussbehaviour}. \begin{figure}[h] \includegraphics[width=0.4\linewidth]{gaussbehaviour} \caption{The expectation value (blue) and variance (green) of the time operator Eq. \eqref{eq:t op main text def} given the initial state $\ket{\Psi(0)} \in \Lambda_{\sqrt{d},(d-1)/2}$, (symmetric, see Def. \ref{def:clock stat classes} and Remark \ref{rem:symme of clock}) for $d=8$, $T_0 = 1$. Compare to Fig. \ref{peresbehaviour}. The ideal case $\braket{\hat{t}} = t$ is in orange. \label{gaussbehaviour}} \end{figure} \end{remark} \begin{definition}\label{def:analytic ext. of clock states} \emph{(Analytic extension of clock states).} Corresponding to every $\ket{\Psi(k_0)} \in \Lambda_{\sigma,n_0}$, we define $\psi : {\mathbbm{R}} \rightarrow \mathbb{C}$ to be the analytic Gaussian function \begin{equation}\label{analyticposition} \psi(k_0;x) = A e^{-\frac{\pi}{\sigma^2}(x-k_0)^2} e^{i2\pi n_0(x-k_0)/d}. \end{equation} By definition, $\psi(k_0;k) = \braket{\theta_k| \Psi(k_0)}$ for $k \in \mathcal{S}_d(k_0)$, and thus $\psi$ is an analytic extension of the discrete coefficients $\braket{\theta_k|\Psi(k_0)}$. In the special case that the corresponding state $\ket{\Psi(k_0)} \in \Lambda_{\sigma,n_0}$ is normalised, $\psi$ will be denoted accordingly, namely \begin{equation}\label{eq:psi nor def} \psi_\textup{nor}(k_0;k):=\psi(k_0;k)\quad \text{iff } A \text{ satisfies Eq. \eqref{eq:A normalised}} \end{equation} \end{definition} \begin{remark}\label{interchangearguments} $\psi(k_0; k+y) = \psi(k_0-y; k)$. \end{remark} \begin{definition}\label{Continuous Fourier Transform no pot} \emph{(Continuous Fourier Transform as a function of dimension $d$ of clock state).} Let $\tilde{\psi} : {\mathbbm{R}} \rightarrow \mathbb{C}$ be defined as the continuous Fourier transform of $\psi$, \begin{align}\label{cftd} \tilde{\psi}(k_0;p) &= \frac{1}{\sqrt{d}} \int_{-\infty}^\infty \psi(k_0;x) \mathrm{e}^{-\mathrm{i} 2\pi px/d} dx = A \frac{\sigma}{\sqrt{d}} \mathrm{e}^{-\frac{\pi \sigma^2}{d^2} (p-n_0)^2} \mathrm{e}^{-i 2\pi n k_0}. \end{align} Similarly to above, we denote \begin{equation}\label{eq:psi f.t. nor def} \tilde\psi_\textup{nor}(k_0;p):=\tilde\psi(k_0;p)\quad \text{iff } A \text{ satisfies Eq. \eqref{eq:A normalised}}. \end{equation} \end{definition} \begin{lemma}[The clock state in the energy basis] This Lemma states that the continuous Fourier transform $\tilde{\psi}(k_0;p)$ is an exponentially good approximation (w.r.t. dimension $d$) of the clock state in the energy basis. Mathematically this is a statement of the closeness of the discrete Fourier transform (D.F.T.) to the continuous Fourier transform (C.F.T.) for Gaussian states. For simplicity, we state the result for the special case $\sigma=\sqrt{d}$, $n_0=(d-1)/2$. \begin{equation} \abs{ \braket{E_n | \Psi_\textup{nor}(k_0)} - \tilde{\psi}_\textup{nor}(k_0;n) } < \left( \frac{2^{\frac{9}{4}}}{1 - e^{-\pi}} \right) d^{-\frac{1}{4}} e^{-\frac{\pi d}{4}}, \quad n=0,1,2,\ldots, d-1. \end{equation} \end{lemma} \begin{proof} From the definition of the state \eqref{def:Gaussian clock states}, and the relation between the time states and energy states \eqref{finitetimestates_main} \begin{align} \braket{E_n | \Psi(k_0)} &= \frac{1}{\sqrt{d}} \;\;\; \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) e^{-i 2\pi nk/d}. \end{align} By definition, this is the D.F.T. of the $\psi(k_0;k)$, state in the time basis. To prove that this approximates the C.F.T. of the state $\psi(k_0,k)$, we first convert the finite sum above to the infinite sum $k\in\mathbb{Z}$, and bound the difference using Lemma \ref{G0}, \begin{equation} \left| \frac{1}{\sqrt{d}} \;\;\; \sum_{k \in \mathcal{S}(d,k_0)} \psi_\textup{nor}(k_0;k)\, e^{-i 2\pi nk/d} - \frac{1}{\sqrt{d}} \;\;\; \sum_{\mathclap{\substack{k \in \mathbb{Z}}}} \psi_\textup{nor}(k_0;k)\, e^{-i 2\pi nk/d} \right| < \left( \frac{2^{\frac{5}{4}}}{1 - e^{-\pi}} \right) d^{-\frac{3}{4}} e^{-\frac{\pi d}{4}}, \end{equation} where we have used the bound for $A$ for normalized states derived in Section \ref{Normalizing the clock state}. Applying the Poisson summation formula (Corollary \ref{poissonsummation}) on the infinite sum, \begin{equation} \frac{1}{\sqrt{d}} \;\;\; \sum_{k \in \mathbb{Z}} \psi_\textup{nor}(k_0;k) e^{-i 2\pi nk/d} = \sum_{m \in \mathbb{Z}} \tilde{\psi}_\textup{nor}(k_0;n + md). \end{equation} Using Lemma \ref{G0}, we can approximate the sum by the $m=0$ term, \begin{equation} \left| \tilde{\psi}_\textup{nor}(k_0;n) - \sum_{m \in \mathbb{Z}} \tilde{\psi}_\textup{nor}(k_0;n + md) \right| < \left( \frac{2^{\frac{5}{4}}}{1 - e^{-\pi}} \right) d^{-\frac{1}{4}} e^{-\frac{\pi d}{4}}. \end{equation} Adding the two error terms, we arrive at the Lemma statement. \end{proof} \begin{remark}[Symmetry of clock states]\label{rem:symme of clock} Now that we know the C.F.T. $\psi$ to be a good approximation of the clock state in the energy basis, we can note the two most important features. First, the width of the state in the energy basis is of the same order ($\sqrt{d}$) as that in the time basis \eqref{analyticposition}. The states are thus symmetric w.r.t. the uncertainty in either basis. In general when $\sigma\neq \sqrt{d}$, the widths of the states in the time and energy bases are of different order. Second, the state $\ket{\Psi_\textup{nor}(k_0)}$ is centered about the value $n_0=(d-1)/2$. Considering that the range of values for the energy number is from $0$ to $d-1$, the energy of the clock state is exactly at the middle of the spectrum. These observations motivate the following definitions. \end{remark} \begin{definition}(Gaussian clock state classes)\label{def:clock stat classes} We give the following names to states $\ket{\Psi(k_0)}\in\Lambda_{\sigma,n_0}$ depending on the relationship between their width $\sigma\in(0,d)$, and the clock dimension $d$: \begin{itemize} \item[1)] \textup{symmetric} if $\sigma^2=d$.\\\vspace{-0.4cm} \item[2)] \textup{time squeezed} if $\sigma^2<d$.\\\vspace{-0.4cm} \item[3)] \textup{energy squeezed} if $\sigma^2>d$. \end{itemize} Furthermore, we add the adverb, \textup{completely} when in addition $n_0=(d-1)/2$, i.e. $\ket{\Psi(k_0)}\in\Lambda_{\sigma,n_0}$ has mean energy centered at the middle of the spectrum of $\hat H_c$. \end{definition} \section{Proof of quasi-continuity of Gaussian clock states} \label{sec:Continuity of the Gaussian clock state} Recall the \emph{continuity} of the ideal clock \eqref{idealcontinuity}. The time-translated state of the clock was the same as the space-translated state, for arbitrary translation size. Our first major result is to derive an analogous statement for the finite clock. This is the subject of this section, with the Theorem stated at the end (Theorem \ref{gaussiancontinuity}). {\bf Sketch of proof:} The mean position $k_0$ of the clock state appears twice in the expression for the state, first as the mean of the Gaussian $\psi(k_0;k)$, and second as determining the set of integers $k$ over which the time states $\ket{\theta_k}$ are defined. Therefore, in order to prove that time translations are (approximately) equivalent to translations in position, we prove that an arbitrary time transation of the clock by a value $\Delta$ shifts both $k_0$ to $k_0+\Delta$ and the set $\mathcal{S}(k_0)$ to $\mathcal{S}(k_0+\Delta)$. By the properties of $\mathcal{S}(k_0)$, we see that it changes at definite values of $k_0$ (integer values if $d$ is even, and half-integer values if $d$ is odd). We proceed in a number of steps. First we prove that an infinitesimal time translation by $\delta$ is approximately equivalent to shifting only the mean $k_0$ by $\delta$. We then use the Lie-Product formula to extend the statement to finite time-translations of $\Delta$ that are small enough so that $\mathcal{S}(k_0+\Delta)$ is the same as $\mathcal{S}(k_0)$. We then bound the error involved in switching the set $\mathcal{S}(k_0)$ to $\mathcal{S}(k_0+1)$ (without any other change in the state). Together the two results then provide a manner of calculating the state for arbitrary time translations: first move only the mean $k_0$ for one unit, then switch $\mathcal{S}(k_0)$ by one integer, and repeat. We proceed by deriving all the necessary technical lemmas which are necessary for Theorem \ref{gaussiancontinuity}. We first prove the quasi-continuity of the Gaussian clock states under infinitesimal time-translations, followed by extending the proof to arbitrary times. \begin{lemma}\label{infinitesimaltimetranslation} The action of the clock Hamiltonian on a Gaussian clock state for infinitesimal time of order $\delta$ may be approximated by an infinitesimal translation of the same order on the analytic extension of the clock state. Precisely speaking, for $\ket{\Psi(k_0)} \in \Lambda_{\sigma,n_0}$, \begin{align}\label{infshift} e^{-i \delta \frac{T_0}{d} \hat{H}_c} \ket{\Psi(k_0)} = e^{-i \delta \frac{T_0}{d} \hat{H}_c} \sum_{k\in\mathcal{S}_d(k_0)} \psi(k_0; k) \ket{\theta_k} = \sum_{k\in\mathcal{S}_d(k_0)} \psi (k_0+\delta; k) \ket{\theta_k} + \ket{\epsilon}, \end{align} where the $l_2$ norm of the error $\ket{\epsilon}$ is bounded by \begin{align} \ltwo{\ket{\epsilon}} &\leq \delta \epsilon_{total} + \delta^2 C^\prime, \quad \text{where $C^\prime$ is $\delta$ independent, and } \\ \frac{\epsilon_{total}}{2\pi Ad} &< \begin{cases}\label{continuityerror}\displaystyle \left( 2 \sqrt{d} \left( \frac{1}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} + \frac{2}{1 - e^{-\pi}} + \frac{1}{2} + \frac{1}{2\pi d} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle \left( 2 \sigma \left( \frac{\alpha_0}{2} + \frac{1}{2\pi\sigma^2} + \frac{1}{1 - e^{-\pi\sigma^2\alpha_0}} \right) e^{-\frac{\pi\sigma^2}{4}\alpha_0^2} + \left( \frac{1}{1 - e^{-\frac{\pi d}{\sigma^2}}} + \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} + \frac{d}{2\sigma^2} + \frac{1}{2\pi d} \right) e^{-\frac{\pi d^2}{4\sigma^2}} \right) &\mbox{otherwise} \end{cases} \end{align} \end{lemma} \begin{remark} Before we begin the proof, note that the R.H.S. of \eqref{infshift} is not necessarily $\ket{\Psi (k_0+\delta)}$, since the summation is still over the set of integers $\mathcal{S}_d(k_0)$, and not $\mathcal{S}_d(k_0+\delta)$ which may be different, see \eqref{gaussianclock} and the remark \ref{clockcentered} that follows. \end{remark} \begin{proof} \begin{align} e^{-i \delta \frac{T_0}{d} \hat{H}_c} \ket{\Psi(k_0)} &= e^{-i \delta \frac{T_0}{d} \sum_{m=0}^{d-1} \frac{2\pi}{T_0} m \ket{E_m}\bra{E_m}} \sum_{k\in \mathcal{S}_d(k_0)} \psi(k_0;k) \ket{\theta_k}. \end{align} Switching to the basis of energy states so as to apply the Hamiltonian, and back to the basis of time states, \eqref{finitetimestates} - \eqref{finiteenergystates}, \begin{align} e^{-i \delta \frac{T_0}{d} \hat{H}_c} \ket{\Psi(k_0)} &= \sum_{k,l\in\mathcal{S}_d(k_0)} \psi(k_0,k) \left( \frac{1}{d} \sum_{n=0}^{d-1} e^{-i2\pi n(k+\delta-l)/d} \right) \ket{\theta_l} \end{align} We label the coefficient in the time basis of the exact state and the approximation as \begin{align} c_l(\delta) &= \braket{\theta_l | e^{-i \delta \frac{T_0}{d} \hat{H}_c} | \Psi(k_0)} = \sum_{k\in\mathcal{S}_d(k_0)} \psi(k_0,k) \left( \frac{1}{d} \sum_{n=0}^{d-1} e^{-i2\pi n(k+ \delta-l)/d} \right), \\ c^\prime_l(\delta) &= \bra{\theta_l} \sum_{k\in\mathcal{S}_d(k_0)} \psi (k_0+\delta; k) \ket{\theta_k}= \psi\left( k_0 + \delta; l \right). \label{idealcoeff} \end{align} By the properties of the analytic extension $\psi$, both of the above are analytic functions w.r.t. $\delta$, and we can express the difference between the coefficients via the Taylor series expansion about $\delta=0$, \begin{align}\label{firstTaylor} c_l(\delta) - c^\prime_l(\delta) &= c_l(0) - c^\prime_l(0) + \delta \left[ \frac{\partial c_l(\delta)}{\partial \delta} - \frac{\partial c^\prime_l(\delta)}{\partial \delta} \right]\bigg|_{\delta=0} + C \delta^2, \\ \text{where } C & = \frac{1}{2} \left( \text{max}_{|t|\leq \delta} \left| \frac{\partial^2 c_l(t)}{\partial t^2} - \frac{\partial^2 c^\prime_l(t)}{\partial t^2} \right| \right). \end{align} $C$ is upper bounded by a $\delta$ independent constant because the second derivatives w.r.t. $\delta$ of both $c_l(\delta)$ and $c^\prime_l(\delta)$ can be verified to be finite sums of bounded quantities, and therefore bounded. We now simplify \eqref{firstTaylor}. By direct substitution, $c_l(0) = c^\prime_l(0)$. For the first derivatives, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} \bigg|_{\delta=0} &= \left[ \frac{\partial}{\partial \delta} \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) \left( \frac{1}{d} \sum_{n=0}^{d-1} e^{-i2\pi n(k+\delta-l)/d} \right) \right]_{\delta=0} \\ &= \left( \frac{-i 2\pi}{d^2} \right) \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) \sum_{n=0}^{d-1} n e^{-i2\pi n(k-l)/d} \end{align} One can replace the finite sum over $k$ by an infinite sum, and bound the difference via Lemma \ref{G0}, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d^2} \right) \sum_{k\in\mathbb{Z}} \psi(k_0;k) e^{-i 2\pi nk/d} \sum_{n=0}^{d-1} n e^{i2\pi nl/d} + \epsilon_1, \\ \text{where} \;\;\;\;\; \left| \epsilon_1 \right| &< \begin{cases}\displaystyle 2\pi A \frac{e^{-\frac{\pi d}{4}}}{1 - e^{-\pi}} &\mbox{if } \sigma=\sqrt{d} \\[5pt]\displaystyle 2\pi A \frac{e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} &\mbox{otherwise} \end{cases}\label{eq:epsilon1 for continuity lema 1} \end{align} Applying the Poisson summation formula \ref{poissonsummation} upon the infinite sum, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d\sqrt{d}} \right) \sum_{n=0}^{d-1} \sum_{m\in\mathbb{Z}} \tilde{\psi}(k_0;n+md) \; n e^{i2\pi nl/d} + \epsilon_1 \end{align} Since $\sum_{n=0}^{d-1} \sum_{m\in\mathbb{Z}} f(n+md) = \sum_{s\in\mathbb{Z}} f(s)$, after some arithmetic manipulation, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d\sqrt{d}} \right) \left( \sum_{s\in\mathbb{Z}} \tilde{\psi}(k_0;s) \; s e^{i2\pi sl/d} - \sum_{n=0}^{d-1} \sum_{m=-\infty}^{\infty} \tilde{\psi}(k_0;n+md) \;md e^{i2\pi nl/d} \right) + \epsilon_1 \end{align} The second summation is a small contribution and may be bound in a similar way to $\epsilon_1$, via Lemmas \ref{G0}-\ref{G1}. \begin{align} \therefore \frac{\partial c_l(\delta)}{\partial \delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d\sqrt{d}} \right) \sum_{s\in\mathbb{Z}} \tilde{\psi}(k_0;s) \; s e^{i2\pi sl/d} + \epsilon_2 + \epsilon_1, \\ \text{where} \;\;\;\;\; & \label{eq:epsilon2 for continuity lema 1}\\ \abs{\epsilon_2} &< \begin{cases}\displaystyle 4\pi A \sqrt{d} \left( \frac{1}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi d}} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 4\pi A \sigma \left( \frac{\alpha_0}{2} + \frac{1}{2\pi\sigma^2} + \frac{1}{1 - e^{-\pi\sigma^2\alpha_0}} \right) e^{-\frac{\pi\sigma^2}{4}\alpha_0^2} &\mbox{otherwise} \end{cases} \end{align} Apply the Poisson summation on the remaining sum, Corollaries \ref{poissonsummation}, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d\sqrt{d}} \right) \sqrt{d} \left( \frac{-id}{2\pi} \right) \sum_{m\in\mathbb{Z}} \frac{\partial}{\partial x} \psi \left( k_0; x + l \right) \bigg|_{x=md} + \epsilon_2 + \epsilon_1 \end{align} Substituting $y=-x$, and using $\psi(k_0;l-y) = \psi(k_0+y,l)$, \ref{interchangearguments}, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} &= \sum_{m\in\mathbb{Z}} \frac{\partial}{\partial y} \psi (k_0+y; l) \bigg|_{y=md} +\epsilon_2 + \epsilon_1. \end{align} The $m=0$ term in the sum above is exactly the derivative $\partial c_l^\prime(\delta)/\partial \delta$ \eqref{idealcoeff}, we may bound the remainder of the sum using Lemmas \ref{G0}-\ref{G1}, \begin{align} \frac{\partial c_l(\delta)}{\partial \delta} &= \frac{\partial c_l^\prime(\delta)}{\partial \delta} + \epsilon_3 + \epsilon_2 + \epsilon_1 \\ \text{where} \;\;\;\;\; \left| \epsilon_3 \right| &< \begin{cases}\displaystyle 4\pi A \left( \frac{1}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 4\pi A \left( \frac{d}{2\sigma^2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} &\mbox{otherwise} \end{cases} \label{eq:epsilon3 for continuity lema 1} \end{align} The Taylor series expansion \eqref{firstTaylor} thus ends up as \begin{align} c_l(\delta) - c^\prime_l(\delta) &= \delta ( \epsilon_1 + \epsilon_2 + \epsilon_3 ) + C \delta^2, \end{align} where using Eqs. \eqref{eq:epsilon1 for continuity lema 1},\eqref{eq:epsilon2 for continuity lema 1},\eqref{eq:epsilon3 for continuity lema 1}, the sum of the errors is bounded by $\abs{\epsilon_1 + \epsilon_2 + \epsilon_3} \leq \epsilon_{4}$, \begin{align} \epsilon_{4}< \begin{cases}\displaystyle 2\pi A \left( 2 \sqrt{d} \left( \frac{1}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} + \frac{2}{1 - e^{-\pi}} + \frac{1}{2} + \frac{1}{2\pi d} \right) e^{-\frac{\pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 2\pi A \Bigg( 2 \sigma \left( \frac{\alpha_0}{2} + \frac{1}{2\pi\sigma^2} + \frac{1}{1 - e^{-\pi\sigma^2\alpha_0}} \right) e^{-\frac{\pi\sigma^2}{4}\alpha_0^2} \\ \quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad+ \left( \frac{1}{1 - e^{-\frac{\pi d}{\sigma^2}}} + \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} + \frac{d}{2\sigma^2} + \frac{1}{2\pi d} \right) e^{-\frac{\pi d^2}{4\sigma^2}} \Bigg) &\mbox{otherwise} \end{cases} \end{align} Note that $\epsilon_{total}$ is independent of the index $l$. Reconstructing the ideal and approximate states from the coefficients $c_l$ and $c_l^\prime$, we arrive at the error in the state, \begin{align} \ltwo{\ket{\epsilon}} \leq \sum_{l\in\mathcal{S}_d(k_0)} \abs{c_l(\delta) - c_l^\prime(\delta)} \leq \delta d \epsilon_{4} + d C \delta^2. \end{align} \end{proof} \begin{lemma}[Finite time translations within a single time step]\label{smalltimetranslation} Given an initial mean position $k_0$ and a translation of $\Delta$, s.t. $\mathcal{S}_d(k_0) = \mathcal{S}_d(k_0+\Delta)$, then \begin{equation} e^{-i \frac{T_0}{d} \Delta \hat{H}_c} \ket{\Psi(k_0)} = \ket{\Psi(k_0+\Delta)} + \ket{\epsilon}, \text{ where } \ltwo{\ket{\epsilon}} < \Delta \epsilon_{total}, \end{equation} where $\epsilon_{total}$ is defined in \eqref{continuityerror}. \end{lemma} \begin{proof} Split the translation $\Delta$ into $M$ equal steps. Then from Lemma \ref{infinitesimaltimetranslation}, for $n\in\{0,1,...,M-1\}$, \begin{equation} e^{-i \frac{T_0}{d} \frac{\Delta}{M} \hat{H}_c} \ket{\Psi(k_0 + n \Delta/M)} = \ket{\Psi(k_0 + (n+1) \Delta/M)} + \ket{\epsilon}, \text{ where } \ltwo{\ket{\epsilon}} < \frac{\Delta}{M} \epsilon_{total} + C^\prime \frac{\Delta^2}{M^2}. \end{equation} Applying Lemma \ref{unitaryerroraddition}, when we add all the steps, we obtain, \begin{align} \ltwo{ \ket{\Psi(k_0+\Delta)} - \left( e^{-i \frac{T_0}{d} \frac{\Delta}{M} \hat{H}_c} \right)^M \ket{\Psi(k_0)} } &< M \left( \frac{\Delta}{M} \epsilon_{total} + C^\prime \frac{\Delta^2}{M^2} \right), \\ \therefore \ltwo{ \ket{\Psi(k_0+\Delta)} - e^{-i \frac{T_0}{d} \Delta \hat{H}_c} \ket{\Psi(k_0)} } &< \Delta \epsilon_{total} + C^\prime \frac{\Delta^2}{M}. \end{align} We are free to choose any positive integer $M$, so we take the limit $M\rightarrow\infty$, which recovers the statement of the Lemma. \end{proof} At this point, we have already proven the continuity of the clock state for time translations that are finite, but small (i.e. small enough that the range $\mathcal{S}_d(k_0)$ remains the same). In order to generalize the statement to arbitrary translations we need to be able to shift the range itself, which is the goal of the following Lemma. \begin{lemma}[Shifting the range of the clock state]\label{rangeshift} If $d$ is even, and the mean of the clock state $k_0$ is an integer, or alternatively, if $d$ is $odd$ and $k_0$ is a half integer, then \begin{align} \ltwo{ \sum_{k \in \mathcal{S}_d(k_0-1)} \psi(k_0;k) \ket{\theta_k} - \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) \ket{\theta_k} } &< \epsilon_{step}, \\ \text{where} \quad \abs{\epsilon_{step}} &< \begin{cases}\displaystyle 2 A e^{-\frac{ \pi d}{4}} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle 2 A e^{-\frac{ \pi d^2}{4\sigma^2}} &\mbox{otherwise} \end{cases} \end{align} \end{lemma} \begin{proof} We prove the statement for even $d$, the proof for odd $d$ is analogous. By definition \eqref{gaussianclock}, $\mathcal{S}_d(k_0)$ is a set of $d$ consecutive integers. Thus the only difference between $\mathcal{S}_d(k_0)$ and $\mathcal{S}_d(k_0-1)$ is the leftmost integer of $\mathcal{S}_d(k_0-1)$ and the rightmost integer of $\mathcal{S}_d(k_0)$, which differ by precisely $d$. By direct calculation, these correspond to the integers $k_0-d/2$ and $k_0+d/2$. These are the only two terms that do not cancel out in the statement of the Lemma, \begin{equation} \ltwo{ \sum_{k \in \mathcal{S}_d(k_0-1)} \psi(k_0;k_0) \ket{\theta_k} - \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) \ket{\theta_k} } = \ltwo{ \psi(k_0;k_0-d/2) \ket{\theta_{k_0-d/2}} - \psi(k_0;k_0+d/2) \ket{\theta_{k_0+d/2}} }. \end{equation} But $\ket{\theta_{k_0-d/2}} = \ket{\theta_{k_0+d/2}}$, \begin{equation} \ltwo{ \sum_{k \in \mathcal{S}_d(k_0-1)} \psi(k_0;k_0) \ket{\theta_k} - \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) \ket{\theta_k} } = \abs{ \psi(k_0;k_0-d/2) - \psi(k_0;k_0+d/2) }. \end{equation} By direct substitution of $\psi(k_0;k_0-d/2)$ and $\psi(k_0;k_0+d/2)$ from \eqref{analyticposition}, we arrive at the Lemma statement. \end{proof} \begin{theorem}[Quasi-continuity of Gaussian clock states]\label{gaussiancontinuity} Let $k_0,t \in {\mathbbm{R}}$. Then the effect of the Hamiltonian $\hat{H}_c$ for the time $t$ on $\ket{{\Psi}_\textup{nor}(k_0)} \in\Lambda_{\sigma,n_0}$ is approximated by \begin{equation}\label{eq:main eq theorem Q continuity} \mathrm{e}^{-\mathrm{i} t \hat{H}_c} \ket{\Psi_\textup{nor}(k_0)} = \ket{\Psi_\textup{nor}(k_0+t \frac{d}{T_0})} + \ket{\epsilon}, \end{equation} where, in the large $d$ limit \begin{align} \varepsilon_c:=\ltwo{ \ket{\epsilon} } = \begin{cases}\displaystyle \frac{t}{T_0} \, \mathcal{O}\left( d^{9/4}\right)\mathrm{e}^{-\frac{\pi}{4}d\alpha_0^2} +\mathcal{O}\left(d^{-1/4}\right)\mathrm{e}^{-\frac{\pi}{4}d} &\mbox{if } \sigma=\sqrt{d} \\[10pt]\displaystyle \frac{t}{T_0} \left( \mathcal{O}\left( d^2\sigma^{1/2} \right)\mathrm{e}^{-\frac{\pi}{4}\sigma^2\alpha_0^2}+\mathcal{O}\left( 1+\frac{d^3}{\sigma^{3/4}} \right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}} \right) +\mathcal{O}\left(1+\frac{1}{\sigma^{1/2}}\right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}}+\mathcal{O}\left(\mathrm{e}^{-\frac{\pi}{4}\sigma^2}\right) &\mbox{otherwise} \end{cases} \end{align} More precisely, \begin{align} \varepsilon_c=\ltwo{ \ket{\epsilon} } &< |t| \frac{d}{T_0} \epsilon_{total} + \left( |t| \frac{d}{T_0} + 1 \right) \epsilon_{step} +\epsilon_\textup{nor}(t), \end{align} where, \begin{align} \epsilon_{total} =\! \begin{cases}\displaystyle 2\pi A d \left( 2 \sqrt{d} \left( \frac{\alpha_0}{2} + \frac{1}{2\pi d} + \frac{1}{1 - e^{-\pi\alpha_0}} \right) e^{-\frac{\pi d}{4}\alpha_0^2} + \left(\frac{2}{1 - e^{-\pi}} + \frac{1}{2} + \frac{1}{2\pi d} \right) e^{-\frac{\pi d}{4}}\right)\! &\mbox{if } \sigma=\sqrt{d} \\[10pt] 2\pi A d \left( 2 \sigma \left( \frac{\alpha_0}{2} + \frac{1}{2\pi\sigma^2} + \frac{1}{1 - e^{-\pi\sigma^2\alpha_0}} \right) e^{-\frac{\pi\sigma^2}{4}\alpha_0^2} + \left( \frac{1}{1 - e^{-\frac{\pi d}{\sigma^2}}} + \frac{1}{1 - e^{-\frac{\pi d^2}{\sigma^2}}} + \frac{d}{2\sigma^2} + \frac{1}{2\pi d} \right) e^{-\frac{\pi d^2}{4\sigma^2}} \right)\! &\mbox{otherwise}, \end{cases} \end{align} with $A=\mathcal{O}(\sigma^{-1/2})$ and is upper bounded by Eq. \eqref{eq:up low bounds for A normalize}, and \begin{align} \epsilon_\textup{nor}(t)&=\sqrt{\frac{\sum_{k\in \mathcal{S}_d(k_0)} \mathrm{e}^{-\frac{2\pi}{\sigma^2}(k-k_0-\frac{d}{T_0}t)^2}}{\sum_{l\in \mathcal{S}_d(k_0)} \mathrm{e}^{-\frac{2\pi}{\sigma^2}(k-k_0)^2}}}-1\leq \begin{cases}\displaystyle 8\sqrt{\frac{2}{d}}\frac{ \mathrm{e}^{-\frac{\pi d}{2}}}{1 - \mathrm{e}^{-\pi d}}\quad &\forall\, t\in{\mathbbm{R}} \;\;\mbox{ if } \sigma=\sqrt{d}\\[10pt]\displaystyle \frac{4\sqrt{2}}{\sigma}\left(\frac{ \mathrm{e}^{-\frac{\pi d^2}{2\sigma^2}}}{1 - \mathrm{e}^{-\frac{2\pi d}{\sigma^2}}}+\frac{ \mathrm{e}^{-\frac{\pi \sigma^2}{2}}}{1 - \mathrm{e}^{-\pi \sigma^2}}\right) \quad &\forall\, t\in{\mathbbm{R}} \;\;\mbox{ otherwise} \end{cases}\\ \epsilon_{step} &< \begin{cases} 2 A e^{-\frac{ \pi d}{4}} &\mbox{if } \sigma=\sqrt{d}\\ 2 A e^{-\frac{ \pi d^2}{4\sigma^2}} &\mbox{otherwise.} \end{cases} \end{align} \end{theorem} \emph{Intuition.} The discrete clock mimics the perfect clock, with an error that grows linearly with time, and scales better than any inverse polynomial w.r.t. the dimension of the clock. The optimal decay is when the state is completely symmetric, (see Def. \ref{def:clock stat classes}). This gives exponentially small error in $d$, the clock dimension, with a decay rate coefficient $\pi/4$. When the initial clock state is symmetric, but not completely symmetric, the error is still exponential decay but now with a modified coefficient which decreases as the clock's initial state's mean energy approaches either end of the spectrum of $\hat H_c$. \begin{proof} For, $t\geq 0$, directly apply the previous two Lemmas (\ref{smalltimetranslation}, \ref{rangeshift}) in alternation, first to move $k_0$ from one integer to the next, then to switch from ${\lfloor k_0\rfloor}$ to ${\lfloor k_0+1\rfloor}$ if $d$ is even, and ${\lfloor k_0+1/2\rfloor}$ to ${\lfloor k_0+1/2+1\rfloor}$ if $d$ is odd, and finally arriving at \begin{align} \mathrm{e}^{-it \hat{H}_c} \ket{{\Psi}(k_0)} &= \ket{{\Psi}(k_0+ \frac{d}{T_0} t)} + \ket{\epsilon},\quad\quad\ltwo{\ket{\epsilon}} \leq t \frac{d}{T_0} \epsilon_{total}+\left(t \frac{d}{T_0}+1\right)\epsilon_{step}. \end{align} To conclude Eq. \eqref{eq: main eq in control theorem} for $t>0$, one now has to normalize the states using bounds from Section \ref{Re-normalizing the clock state} and then use Lemma \ref{unitaryerroraddition} to upper bound the total error. For $t<0$, simply evaluate Eq. \eqref{eq:main eq theorem Q continuity} for a time $|t|$ followed by multiplying both sides of the equation by $\exp(\mathrm{i} t\hat H_c)$, mapping $k_0\rightarrow k_0-t\;d/T_0$ and noting the unitary invariance of the $l_2$ norm of $\ket{\epsilon}$. \end{proof} \section{Autonomous quasi-control of Gaussian clock states}\label{sec:Proof of quasi-continuous control of Gaussian clock states} \subsection{Theorem and proof of autonomous quasi-control} Recall the \textit{control} of the idealized clock \ref{idealregularity}. The time translated state of the clock under its natural Hamiltonian plus a potential for a time $t$, was the same as the space-translated state multiplied by a phase factor, with the phase given by the integral over the potential up to time $t$. Our second major result is to derive an analogous statement for the finite clock. This is the subject of this Section, with the result in the form of a Theorem stated at the end of the Section, Theorem (\ref{movig through finite time}). The proof will follow similar lines to the proof of the continuity of the Gaussian clock states of the previous Section (\ref{sec:Continuity of the Gaussian clock state}). The main extra complications will be two: \begin{itemize} \item[1)] Unlike in the previous Section (\ref{sec:Continuity of the Gaussian clock state}), the Hamiltonian will now include a potential term, i.e. $\hat H=\hat H_c+\hat V_d$ where $\hat H_c$ and $\hat V_d$ do not commute. To get around this difficulty, we will have to employ the Lie product formula to split the time evolution into consecutive infinitesimal time steps with separate contributions from $\hat H_c$ and $\hat V_d$. Bounding the infinitesimal time step contribution from $\hat H_c$ will be the subject of Lemma \ref{lemm:1}, while the contribution from $\hat V_d$ will be bounded in Lemma \ref{lem:2 infinitesimal pot}. Combining both contributions will be the subject of Lemma \ref{Moving the clock through finite time, within unit angle}. \item [2)] The Fourier Transform of the clock state (see Def. \ref{Continuous Fourier Transform no pot}), will be replaced by the Fourier Transform of the clock state plus a phase factor which depends on the potential. Unlike in the previous case, one will not be able to perform the Fourier Transform analytically, thus requiring to upper bound how quickly it decays with clock dimension $d$. This turns out not to be so simple, and is the subject of Lemma \ref{lem:the crucial fourierbound with potential} which is the main mathematically challenging difference between this Section and Section \ref{sec:Continuity of the Gaussian clock state}. \end{itemize} \begin{definition}\label{def:continous pot} \emph{(Continuous Interaction Potential)}. Let ${V}_0 : {\mathbbm{R}} \rightarrow {\mathbbm{R}}$ be an infinitely differentiable function of period $2\pi$, normalized so that \begin{equation}\label{eq:theta normalisation} \int_x^{x+2\pi} {V}_0(x^\prime) dx^\prime = \Omega, \;\;\;\;\; -\pi\leq\Omega<\pi,\quad x\in {\mathbbm{R}} \end{equation} and define ${V}_d : {\mathbbm{R}} \rightarrow {\mathbbm{R}} $ as \begin{equation}\label{potstretch} {V}_d(x) = \frac{2\pi}{d}\,{V}_0 \left( \frac{2\pi x}{d} \right). \end{equation} For convenience of notation, we also define the phase function \begin{equation}\label{eq: Theta def} \Theta(\Delta;x)=\int_{x-\Delta}^x dy V_d(y). \end{equation} \end{definition} \begin{definition}\label{def:interaction pot def} \emph{(Interaction Potential).} Let $\hat{V}_d$ be a self-adjoint operator on $\mathcal{H}_c$ defined by \begin{equation} \hat{V}_d = \frac{d}{T_0} \sum_k {V}_d(k) \ketbra{\theta_k}{\theta_k}. \end{equation} Thus ${V}_d(x)$ is a continuous extension of the discrete elements $\braket{\theta_k|\hat{V}_d|\theta_k}T_0/d$. Note that because ${V}_d$ has a period of $d$, the summation in the above expression may run over any sequence of $d$ consecutive integers without affecting the operator $\hat{V}_d$. \end{definition} \begin{remark} It is noteworthy that Theorem \ref{movig through finite time} holds under the generalization of the potential $V_0$ to the new definition $V_0:{\mathbbm{R}}\rightarrow {\mathbbm{C}}$ with $V_0$ still infinitely differentiable and with period $2\pi$, while replacing constraint Eq. \eqref{eq:theta normalisation} with the constraint $ \int_x^{x+2\pi} {V}_0(x^\prime) dx^\prime = \Omega+\mathrm{i}\Omega', \;\;\; -\pi\leq\Omega<\pi,$ \;\;\; $\Omega'\leq 0$, $\quad x\in {\mathbbm{R}}$. Such a generalization has not been considered in this manuscript for simplicity and since in this case $\hat V_d$ in Def. \ref{def:interaction pot def} is no longer self-adjoint, and thus neither is $\hat H_c+\hat V_d$. The more general setup could be useful for using the potential to perform weak measurements on the clock however. \end{remark} \begin{definition}\label{def:stand alone def mathcal N}\emph{(Decay rate parametres).} Let $b$ be any real number satisfying \begin{equation}\label{eq:b def eq} b\geq\; \sup_{k\in{\mathbbm{N}}^+}\left(2\max_{x\in[0,2\pi]} \left| V_0^{(k-1)}(x) \right|\,\right)^{1/k}, \end{equation} where $ V_0^{(p)}(x)$ is the $p^\textup{th}$ derivative with respect to $x$ of $ V_0(x)$ and $V_0^{(0)}:=V_0$. We can use $b$ to define $\mathcal{N}\in{\mathbbm{N}}^0$ as follows \begin{equation} \mathcal{N}= \begin{cases}\displaystyle \left\lfloor \frac{\pi\alpha_0^2}{2\left(\frac{\pi\kappa\alpha_0}{\ln(\pi\alpha_0 d)}b+1\right)^2} \,d \right\rfloor, &\mbox{if } \sigma=\sqrt{d}\\[20pt]\displaystyle \left\lfloor \frac{\pi\alpha_0^2}{2\left(\frac{\pi\kappa\alpha_0}{\ln(\pi\alpha_0\sigma^2)}b+\frac{d}{\sigma^2}\right)^2} \left(\frac{d}{\sigma}\right)^2 \right\rfloor, &\mbox{otherwise} \end{cases} \end{equation} where $\kappa=0.792$ and $\alpha_0\in(0,1)$ characterizes how close the mean energy of the initial state of the clock is to either extremum of the energy spectrum, defined in Def. \ref{Distance of the mean energy from the edge}. \iffalse namely \begin{align}\label{eq:alpha_0 def} \alpha_0&=\left(\frac{2}{d-1}\right) \min\{n_0,(d-1)-n_0\}\\ &=1-\left|1-n_0\,\left(\frac{2}{d-1}\right)\right|\in(0,1),\quad \text{for } n_0\in(0,d-1). \end{align} c.f. similar measures Def. \eqref{Distance of the mean energy from the edge} , and Eq. \eqref{eq:alpha_c def eq}. \fi \end{definition} We can use the above definitions to define the following parametre. \begin{definition}\label{def:decay rate params}\emph{(Exponential decay rate parameter)}. We define the rate parameter as \begin{equation} \bar\upsilon= \frac{\pi\alpha_0\kappa}{\ln\left(\pi\alpha_0\sigma^2 \right)}b \end{equation} where recall $\kappa=0.792$. \end{definition} We will often require that $\bar\upsilon\geq 0$. This is equivalent to requiring $\pi\alpha_0 \sigma^2>1$ if $b>0$, and is always satisfied in the special case $b=0$.\\ To proceed further, we will need a generalized definition of \ref{def:analytic ext. of clock states} which incorporates a potential. \begin{definition}\label{def:analyticmomentum control version} \emph{(Continuous Fourier Transform of the control analytic extension)} Let $\tilde{\psi}(k_0;\cdot) : {\mathbbm{R}} \rightarrow \mathbb{C}$ be defined as the continuous Fourier transform of $\psi$ multiplied by a phase function, \begin{align}\label{analyticmomentum control version} \tilde{\psi}(k_0,\Delta ;p) &= \frac{1}{\sqrt{d}} \int_{-\infty}^\infty \psi(k_0 ;x)\,\mathrm{e}^{-\mathrm{i} \int_{x-\Delta}^x V_d(x')dx'} \mathrm{e}^{-\mathrm{i} 2\pi px/d} dx , \end{align} where $\psi$ is defined in \ref{def:analytic ext. of clock states} and $ V_d$ in Def. \ref{def:continous pot}. \end{definition} \begin{remark} We observe that $\tilde \psi(k_0,0 ;x)=\tilde \psi(k_0;x)$ (see Def. \ref{def:analytic ext. of clock states}). \end{remark} \begin{lemma}\label{lem:the crucial fourierbound with potential}\emph{(Bounding a sampled version of the Fourier Transform of the clock state with potential).} Let $\bar\epsilon_2$ be defined by \begin{equation} \bar\epsilon_2=\frac{\mathrm{i} 2\pi}{\sqrt{d}}\sum_{n=0}^{d-1}\sum_{k=-\infty}^\infty k\, \mathrm{e}^{\mathrm{i} 2\pi n l/d}\, \tilde \psi(k_0,\Delta;n+k d). \end{equation} We have the bound \begin{equation} |\bar\epsilon_2|< \begin{cases} (2\pi)^{5/4}\sigma^{3/2} A \left(1+\frac{\pi^{2}}{8}\right)\sqrt{\frac{\mathrm{e}}{2}\frac{\alpha_0}{(\bar\upsilon\sigma^2/d+1)}}\exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(\frac{d}{\sigma^2}+\bar\upsilon\right)^2} \left(\frac{d}{\sigma}\right)^2 \right)\quad &\text{if } \mathcal{N}\geq 8 \text{ and } \bar\upsilon\geq 0\\ \frac{3^{7/4}}{\sqrt{2\pi}\mathrm{e}}\frac{A(8+\pi^{2})}{\alpha_0^3}\left(\frac{\kappa\sqrt{6\pi}}{\ln(3)}b+\frac{d}{\sigma}\right)^3 \left( \frac{\sigma}{d^3} \right)\quad &\text{otherwise} \label{eq:lemma sampled ft 2} \end{cases} \end{equation} where $b$, $\mathcal{N}$, and $\alpha_0$ are defined in Def. \ref{def:stand alone def mathcal N} while $\bar\upsilon$ in Def. \ref{def:decay rate params}. \end{lemma} \begin{proof} Outline: to bound $|\bar\epsilon_2|$, the main challenge is in bounding the Fourier transform $\tilde \psi(\cdot,\cdot;\cdot)$, and show that it is exponentially decaying in $d$ (for $d$ large enough such that the first if condition in Eq. \eqref{eq:lemma sampled ft 2} is satisfied). In order to achieve this, we will integrate by parts the Fourier transform $N=1,2,3,\ldots$ times followed by taking absolute values, thus generating a different bound for every $N$. We will then choose to bound the Fourier transform using a different bound (i.e. different $N$), depending of the value of $d$, i.e. $N=N(d)$. We will have to bound the derivatives produced by integrating by parts $N$ times. We will use results from combinatorics for this. The proof will make essential use of: the binomial theorem, the generalized Leibniz rule, Rodriguez formulas for Hermite polynomials, orthogonality conditions of Hermite polynomials, the Cauchy-Schwarz inequality, Fa\a di Bruno's formula, Bell polynomials, Bell numbers, analytic upper bounds for Bell numbers, Sterling's formula, and the Fundamental theorem of calculus. Also note that definitions \ref{def:stand alone def mathcal N} and \ref{def:decay rate params} have been defined more generally in the proof. One could use these slightly more general definitions to tighten the bound in Eq. \eqref{eq:lemma sampled ft 2} for small $d$ if desired. We have that \begin{equation}\label{eq:ep2 def} |\bar\epsilon_2|\leq \frac{2 \pi}{\sqrt{d}} \sum_{n=0}^{d-1}\left( \sum_{s=-\infty} ^{-1} |s\,\tilde \psi( k_0,\Delta; n+sd)| +\sum_{s=1}^\infty |s\,\tilde \psi( k_0,\Delta; n+sd)| \right), \end{equation} where from definitions \ref{def:analyticmomentum control version}, \ref{def:analytic ext. of clock states} we have \begin{equation}\label{int tilde phi} \tilde \psi( k_0,\Delta;p)=\frac{A}{\sqrt{d}} \int _{-\infty}^\infty dx\, \mathrm{e}^{-\frac{\pi}{\sigma^2}(x-k_0)^2}\mathrm{e}^{\mathrm{i} 2\pi n_0(x-k_0)/d} \mathrm{e}^{-\mathrm{i} \theta(x)} \mathrm{e}^{-\mathrm{i} 2 \pi p x/d}. \end{equation} where we denote \begin{equation}\label{eq:theta mischa def} \theta(x)= \frac{2 \pi}{d}\int_{x-\Delta}^x dy V_0(2\pi y/d), \end{equation} where $ V_0(\cdot)$ is a smooth, real, periodic function with period $2\pi$ defined in Eq. \eqref{eq:theta normalisation} Performing the change of variable $z=2 \pi(k_0-x)/d$, we find in Eq. \eqref{int tilde phi} \begin{equation}\label{eq:U in int 1st time} \tilde \psi( k_0,\Delta; p)=\frac{A}{\sqrt{d}}\frac{d}{2\pi} \mathrm{e}^{-\mathrm{i} p 2\pi k_0/d}\int _{-\infty}^\infty dz\,U(z) \mathrm{e}^{\mathrm{i}( p+\gamma) z}, \end{equation} where \begin{equation} U(z):= \mathrm{e}^{-\left(\frac{d }{2\sqrt{\pi}\sigma}\right)^2 z^2} \mathrm{e}^{-\mathrm{i} \tilde\theta(z)}\mathrm{e}^{-\mathrm{i} (n_0+\gamma)z}, \end{equation} with \begin{equation}\label{eq:theta tilde mischa def} \tilde\theta(z):=\theta\left(-\frac{d z}{2 \pi}+k_0 \right), \end{equation} and $\gamma\in \big(-(d-1)/2,(d-1)/2 \big).$ We now integrate by parts $N$ times the integral in Eq. \eqref{eq:U in int 1st time}, differentiating $U(\cdot)$ once in every iteration. This requires that $U(\cdot)$ is differentiable $N$ times and since we will require $N$ to be unbounded from above, hence the requirement that $ V_0(\cdot)$ is smooth. Taking this all into account, we have \begin{equation} \tilde \psi(k_0,\Delta; p)=-\frac{A}{\sqrt{d}} \frac{d}{2\pi}\mathrm{e}^{-\mathrm{i} p 2\pi k_0/d}\frac{1}{\big(-\mathrm{i}\left(p+\gamma\right)\big)^{N}}\int _{-\infty}^\infty dz\,U^{(N)}(z) \mathrm{e}^{\mathrm{i}\big( p+\gamma\big) z},\quad N\in{\mathbbm{N}}^+, \end{equation} where $U^{(N)}(z)$ denotes the $N^\text{th}$ derivative of $U(z)$ w.r.t. $z$. Thus taking absolute values, we achieve \begin{equation}\label{eq:div pars ineq} \tilde \psi( k_0,\Delta; p)\leq\frac{A}{\sqrt{d}}\frac{d}{2\pi} |p+\gamma|^{-N}\int _{-\infty}^\infty dz\,\left|U^{(N)}(z)\right|,\quad N\in{\mathbbm{N}}^+. \end{equation} We now substitute Eq. \eqref{eq:div pars ineq} into Eq. \eqref{eq:ep2 def} obtaining the upper bound \begin{align}\label{eq:ep2 1sr up bound} |\bar\epsilon_2|&\leq \frac{2\pi}{\sqrt{d}}\frac{A}{\sqrt{d}}\frac{d}{2\pi}\left(\int_{-\infty}^\infty dz \left| U^{(N)}(z)\right|\right) \sum_{n=0}^{d-1}\sum_{s=1}^\infty\left( \frac{|s|}{ \left|n+sd+\gamma\right|^{N}} + \frac{|s|}{ \left|n-sd+\gamma\right|^{N}} \right)\\ &\leq \left( A G d\int_{-\infty}^\infty dz \left| U^{(N)}(z)\right|\right) \left(\frac{d (1-|\beta|)}{2}\right)^{-N} , \quad N=3,4,5,\ldots \end{align} where $\beta\in(-1,1)$ and \begin{align}\label{def: G definition} G:= \frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=1}^\infty\frac{|s|}{\left(\frac{2}{d(1-|\beta|)}\right)^N} \left( \frac{1}{ \big|n+sd+\gamma\big|^{N}} + \frac{1}{ \big|n-sd+\gamma\big|^{N}} \right). \end{align} The rational to defining $G$ in this way, is that we will soon parametrize $\gamma$ in terms of $\beta$ such that $G$ will be upper bounded by a $d$ independent constant. We will now find this constant before proceeding to bound Eq. \eqref{eq:ep2 1sr up bound}. First we will parametrize $\gamma$ is such a way that the symmetry in the summations of $G$ becomes transparent. Let $\gamma=-(d-1)/2-y_0$, $y_0\in \big[0,-(d-1)/2\big)$. Substituting into Eq. \eqref{def: G definition}, we find \begin{align} G=& \frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=1}^\infty\frac{|s|}{\left(\frac{2}{d(1-|\beta|)}\right)^N} \left( \frac{1}{ \big|n+sd-(d-1)/2-y_0\big|^{N}} + \frac{1}{ \big|n-sd-(d-1)/2-y_0\big|^{N}} \right)\\ =& \frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=1}^\infty\frac{|s|}{\left(\frac{2}{d(1-|\beta|)}\right)^N} \left( \frac{1}{ \big|n+sd-(d-1)/2-y_0\big|^{N}} + \frac{1}{ \big|-n-sd+(d-1)/2-y_0\big|^{N}} \right)\\ =& \frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=1}^\infty\frac{|s|}{\left(\frac{2}{d(1-|\beta|)}\right)^N} \left( \frac{1}{ \big|n+sd-(d-1)/2-|y_0|\big|^{N}} + \frac{1}{ \big|n+sd-(d-1)/2+|y_0|\big|^{N}} \right),\label{eq:G intermediate} \end{align} and thus the sum only depends on the modulus of $y_0$. We will now make the change of variable $y_0=(d-1)\beta/2$ with $\beta\in(-1,1)$ leading to \begin{equation}\label{eq:gamma in terms of beta} \gamma= -\left(\frac{d-1}{2}\right) (1+\beta), \quad \beta\in(-1,1). \end{equation} Plugging this into Eq. \eqref{eq:G intermediate} leads to \begin{align} G=& \frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=1}^\infty \left( \frac{|s|}{\left|\frac{(2n+1+|\beta|)}{(1-|\beta|)d}-\frac{(1+|\beta|)-2s}{(1-|\beta|)}\right|^{N}} + \frac{|s|}{\left|\frac{(2n+1-|\beta|)}{(1-|\beta|)d}+\frac{(1-|\beta|)+2s}{(1-|\beta|)}\right|^{N}} \right)\\ =&\frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=0}^\infty \left( \frac{s+1}{\left(\frac{(2n+1+|\beta|)}{(1-|\beta|)d}+1+\frac{2s}{1-|\beta|}\right)^{N}} +\sum_{s=1}^\infty \frac{s}{\left(\frac{(2n+1-|\beta|)}{(1-|\beta|)d}+1+\frac{2s}{(1-|\beta|)}\right)^{N}} \right) \\ =&\frac{1}{d}\sum_{n=0}^{d-1}\sum_{s=0}^\infty \left( \frac{s+1}{\left(1+2s\right)^{N}} +\sum_{s=1}^\infty \frac{s}{\left(1+2s\right)^{N}} \right)=1+\sum_{s=0}^\infty \frac{1}{\left(2s+1\right)^{N-1}}\\ \leq & 1+\sum_{s=0}^\infty \frac{1}{(2s+1)^2}=1+\frac{\pi^{2}}{8}\approx 2.234\quad \text{for } N=3,4,5,\ldots\,. \end{align} Our next task will be to bound $\int_{-\infty}^\infty dz \left| U^{(N)}(z)\right|$. We will start by dividing $U(z)$ into a product of unitaries: $U(z)=U_1(z)U_2(z)U_3(z)$ where \begin{equation} U_1(z):= \mathrm{e}^{-d z^2/4\pi},\quad U_2(z):= \mathrm{e}^{-\mathrm{i} \tilde \theta(z)},\quad U_3(z):= \mathrm{e}^{-\mathrm{i}(n_0+\gamma)z}. \end{equation} We will take advantage of the distinct properties of $U_1,U_2,U_3$. We will start by using the \textit{general Leibniz rule} for $n$ times differentiable functions $u$ and $v$: \begin{equation} (uv)^{(n)}=\sum_{k=0}^n \binom{n}{k} u^{(k)}v^{(n-k)}, \end{equation} where $\binom{n}{k}$ is the \textit{binomial coefficient} we are using the standard superscript round bracket notation to indicated derivatives. We thus find \begin{align} \int_{-\infty}^\infty dz\left| U^{(N)}(z)\right| &=\int_{-\infty}^\infty dz\left| \sum_{k=0}^N \binom{N}{k} \left(U_1(z)U_2(z)\right)^{(k)} U_3^{(N-k)}(z)\right|\\ &=\int_{-\infty}^\infty dz\left| \sum_{k=0}^N \binom{N}{k}\sum_{q=0}^k \binom{k}{q}U_1^{(q)}(z)U_2^{(k-q)}(z)\left(-\mathrm{i}(n_0+\gamma)\right)^{N-k} U_3(z)\right|\\ &\leq \sum_{k=0}^N \binom{N}{k}\sum_{q=0}^k \binom{k}{q}|n_0+\gamma|^{N-k}\int_{-\infty}^\infty dz\left|U_1^{(q)}(z)U_2^{(k-q)}(z)\right|\\\label{eq:U N dev in terms of U1 U2} \end{align} We now need to relate $U_1^{(q)}$ to the Hermite polynomials in order to bound the integral. The Rodriguez formula for the $n^\text{th}$ Hermite polynomial $H_n(x)$ is \cite{Rodriguez} \begin{equation} H_n(x)=(-1)^n \mathrm{e}^{x^2}\frac{d^n}{dx^n}\left(\mathrm{e}^{-x^2}\right), \quad n\in {\mathbbm{N}}^0. \end{equation} By the change in variable $x=\frac{d}{2\sigma\sqrt{\pi}}z$ we can relate $U_1^{(q)}$ to the Hermite polynomials: \begin{equation}\label{eq:U in terms of Hermite} U^{(n)}_1(z)=\left(\frac{d}{2\sigma\sqrt{\pi}}\right)^{n}(-1)^n\, \mathrm{e}^{-\left(\frac{d}{2\sigma\sqrt{\pi}}\right)^2 z^2} H_n\left(\frac{d}{2\sigma\sqrt{\pi}}z\right),\quad n\in {\mathbbm{N}}^0. \end{equation} Due to the periodic nature of $ V_0(\cdot)$, we have that $|U_2^{(k)}(z)|$ is bounded in $z\in{\mathbbm{R}}$ for $k\in {\mathbbm{N}}^0$. We can thus use the Cauchy--Schwarz inequality in conjunction with Eqs. \eqref{eq:U N dev in terms of U1 U2}, \eqref{eq:U in terms of Hermite} to obtain \begin{align} &\int_{-\infty}^\infty dz\left| U^{(N)}(z)\right| \\ &\leq \sum_{k=0}^N \binom{N}{k}\sum_{q=0}^N \binom{N}{q}|n_0+\gamma|^{N-k}\left(\frac{d}{2\sigma\sqrt{\pi}}\right)^{q}\sqrt{ \int_{-\infty}^\infty dy\, \mathrm{e}^{-\left(\frac{d}{2\sigma\sqrt{\pi}}\right)^2 y^2}\left|U_2^{(k-q)}(z) \right|^2 \int_{-\infty}^\infty dz H_q^2\left(\frac{d}{2\sigma\sqrt{\pi}}\,z\right)\mathrm{e}^{-\left(\frac{d}{2\sigma\sqrt{\pi}}\,z\right)^2 z^2} }\\ &\leq \sum_{k=0}^N \binom{N}{k}\sum_{q=0}^N \binom{k}{q}|n_0+\gamma|^{N-k}\left(\frac{d}{2\sigma\sqrt{\pi}}\right)^{q-1/2} C_{k-q}\,\sqrt{ \int_{-\infty}^\infty dy\, \mathrm{e}^{-\left(\frac{d}{2\sigma\sqrt{\pi}}\right)^2 y^2} \int_{-\infty}^\infty dx\, H_q^2\left(x\right)\mathrm{e}^{-x^2} }\label{eq:U N dev in terms of H_n} \end{align} where we have defined \begin{equation}\label{eq:def Ck} C_k\geq \left|U_2^{(k)}(z)\right|,\quad \forall z\in {\mathbbm{R}},\;\;\; k\in {\mathbbm{N}}^0. \end{equation} Using the orthogonality conditions of the Hermite polynomials\cite{Rodriguez}: \begin{equation} \int_{-\infty}^\infty dy H_p(y) H_q(y)\mathrm{e}^{- y^2}=\delta_{p,q} \sqrt{\pi}\, 2^q (q!),\quad p,q\in {\mathbbm{N}}^0 \end{equation} Eq. \eqref{eq:U N dev in terms of H_n} reduces to \begin{align}\label{eq:U N dev in terms of C k} \int_{-\infty}^\infty dz\left| U^{(N)}(z)\right| \leq & \frac{2\sigma\pi}{d}\sum_{k=0}^N \binom{N}{k}|n_0+\gamma|^{N-k}\sum_q ^k\binom{k}{q} \left(\frac{d}{\sigma\sqrt{2\pi}}\right)^{q} C_{N-q} \sqrt{(q!)}. \end{align} Before we can continue, we will now bound $C_k$ in terms of derivatives of $\tilde \theta(\cdot)$. For this, let us first recall Fa\`a di Bruno's formula written in terms of the Incomplete Bell Polynomials $B_{n,k}(x_1,x_2,\ldots,x_{n-k+1})$ \cite{Faa_di_Bruno}: for $u$,$v$ $n$-times differentiable functions, we have \begin{equation} \frac{d^n}{dx^n}u\left(v(x)\right)=\sum_{k=0}^n u^{(k)}\left( v(x)\right) B_{n,k}\left(v^{(1)}(x),v^{(2)}(x),\ldots,v^{(n-k+1)}(x)\right). \end{equation} By choosing $U(x)=\mathrm{e}^x$, $v(x)=-\mathrm{i} \tilde \theta(x)$, it follows \begin{equation}\label{eq:Un bell poly} U_2^{(n)}(x)=U_2(x) \sum_{k=1}^n B_{n,k}\left(-\mathrm{i} \tilde\theta^{(1)}(x),\ldots,-\mathrm{i} \tilde\theta^{(n-k+1)}(x)\right)=: U_2(x) B_{n}\left(-\mathrm{i} \tilde\theta^{(1)}(x),\ldots,-\mathrm{i} \tilde\theta^{(n)}(x) \right), \end{equation} where $B_n$ are the Complete Bell Polynomials. Using the formula\cite{Faa_di_Bruno} \begin{equation} B_{n,k}(x_1,x_2,\ldots,x_{n-k+1})=\sum_{\{j_k\}_k} \frac{n!}{j_1!j_2!\ldots j_{n-k+1}!} \left(\frac{x_1}{1!}\right)^{j_1}\left(\frac{x_2}{2!}\right)^{j_2}\ldots \left(\frac{x_{n-k+1}}{(n-k+1)!}\right)^{j_{n-k+1}}, \end{equation} where $j_1+j_2+\ldots+j_{n-k+1}=k$ and $j_1+2j_2+\ldots+({n-k+1})j_{n-k+1}=n$, we see that if $x_i\leq a b^i$, for $a,b>0$ $i=1,2,3,\ldots ,{n-k+1}$, we have \begin{equation} \left|B_{n,k}(x_1,x_2,\ldots,x_{n-k+1})\right| \leq b^n B_{n,k}(a,a,\ldots,a). \end{equation} Let \begin{equation}\label{eq: dev tilde theta constraint} \left| \tilde \theta^{(n)}(x)\right|\leq a\, b^n, \end{equation} for some $a,b>0$ for $n\in {\mathbbm{N}}^+$, Eq. \eqref{eq:Un bell poly} gives us \begin{equation} \left|U_2^{(n)}(x)\right| \leq b^n B_n(a,a,\ldots,a). \end{equation} Setting $a=1$ and noting that $B_n(1,1,\ldots,1)=\textup{Be}_n$, where $\textup{Be}_n$ is the $n^\textup{th}$ Bell number\cite{Faa_di_Bruno}, we achieve \begin{equation}\label{eq: U2 in Bell num} \left|U_2^{(n)}(x)\right| \leq b^n\, \textup{Be}_n. \end{equation} Using Eqs. \eqref{eq:def Ck}, \eqref{eq:U N dev in terms of C k}, \eqref{eq: U2 in Bell num}, and introducing variables $\upsilon,b_2\geq 0$ via the definition \begin{equation}\label{eq:b upsilon and b_2 def} b=\upsilon b_2, \end{equation} we have \begin{align} \int_{-\infty}^\infty dz\left| U^{(N)}(z)\right| \leq & \frac{2\pi\sigma}{d}\sum_{k=0}^N \binom{N}{k}|n_0+\gamma|^{N-k}\sum_{q=0}^k \binom{k}{q}\upsilon^{k-q} \max_{p=0,\ldots,k}\left\{\left(\frac{d}{\sigma\sqrt{2\pi}}\right)^{p} \sqrt{(p!)}\,b_2^{k-p}\, \textup{Be}_{k-p} \right\}\\ \leq & \frac{2\pi\sigma}{d}\sum_{k=0}^N \binom{N}{k}|n_0+\gamma|^{N-k}\sum_{q=0}^k \binom{k}{q}\upsilon^{k-q} \max_{p=0,\ldots,N}\left\{\left(\frac{d}{\sigma\sqrt{2\pi}}\right)^{p} \sqrt{(p!)}\,b_2^{N-p}\,\textup{Be}_{N-p} \right\},\\ \leq & \frac{2\pi\sigma}{d}\left(|n_0+\gamma|+\upsilon+1\right)^N \max_{p=0,\ldots,N}\left\{\left(\frac{d}{\sigma\sqrt{2\pi}}\right)^{p} \sqrt{(p!)}\,b_2^{N-p}\, \textup{Be}_{N-p} \right\},\label{eq:U N 1st max prob} \end{align} where we have used twice the identity \begin{equation} \sum_{k=0}^n\binom{n}{k} g_1^k g_2^{n-k}=(g_1+g_2)^n\quad \forall g_1,g_2,\in {\mathbbm{R}},\, n\in {\mathbbm{N}}^0. \end{equation} We will now proceed to upper bound the maximisation problem in Eq. \eqref{eq:U N 1st max prob}. Using Sterling's bound for factorials and the a bound for the Bell numbers \cite{Berendr}, \begin{equation} n!\leq \mathrm{e}\, n^{n+1/2} \mathrm{e}^{-n},\quad \textup{Be}_n< \left(\frac{\kappa\, n}{\ln(n+1)}\right)^n, \text{with } \kappa=0.792,\quad n\in{\mathbbm{N}}^+ \end{equation} together with $\textup{Be}_0=1$, we can thus write \begin{align}\label{eq:max f def} \max_{p=0,\ldots,N}\left\{\left(\frac{d}{\sigma\sqrt{2\pi}}\right)^{p} \sqrt{(p!)}\,b_2^{N-p}\, \textup{Be}_{N-p} \right\}\\ \leq \sqrt{\mathrm{e}}\, \exp \left(\max_{q=0,\ldots,N}\left\{f(q)\right\}\right), \end{align} where we have defined $f: 0\cup [1,N]\rightarrow {\mathbbm{R}}$ for $N=3,4,5,\ldots$ \begin{equation}\label{def:f def} f(x):= \begin{cases} N\ln b_2+N\ln\left(\frac{\kappa N}{\ln(N+1)}\right), & \quad \text{if } x=0 \\ x\ln\left(\frac{d}{\sqrt{2\pi}\sigma}\right)+(x/2+1/4)\ln(x)-\frac{x}{2}+(N-x)\ln\left(\frac{\kappa (N-x)}{\ln(N-x+1)}\right)+(N-x)\ln(b_2), & \quad \text{if } x\in[1,N)\\ N\ln\left(\frac{d}{\sqrt{2\pi}\sigma}\right)+(N/2+1/4)\ln(N)-\frac{N}{2}, & \quad \text{if } x=N. \end{cases} \end{equation} Note that $f(x)$ is continuous on the interval $x\in [1,N]$. By explicit calculation, we have \begin{equation}\label{eq:2nd dev of f } f^{(2)}(x)=\frac{1}{N-x}+\frac{1}{2x}\left( 1-\frac{1}{2x}\right)+\frac{G(N-x)}{(N-x)(N-x+1)^2\ln^2(N-x+1)} \text{ for}\quad x\in[1,N),\,\, N=3,4,5,\ldots \end{equation} where \begin{equation} G(x):=(x+1)^2\ln^2(x+1)+x^2-(x+2)x\ln(x+1),\quad x\geq 0. \end{equation} Due to the following two properties satisfied by $G$, \begin{align} G(0)&=0,\\ \frac{d}{dx} G(x)&=\frac{x^2+2(1+x)^2\ln^2(1+x)}{1+x}>0,\quad\text{for } x>0, \end{align} we conclude $G(x)>0$ for $x>0$ and thus from Eq. \eqref{eq:2nd dev of f }, \begin{equation} f^{(2)}(x)>0 \text{ for } x \in[1,N), \quad \forall \,N= 3,4,5,\ldots \end{equation} hence $f(x)$ is convex on $x\in[1,N)$ and we can write Eq. \eqref{eq:max f def} as \begin{equation}\label{eq:max written with f q} \max_{p=0,\ldots,N}\left\{\left(\frac{d}{\sqrt{2\pi}\sigma}\right)^{p} b_2^{N-p}\, \textup{Be}_{N-p}\, \sqrt{(p!)}\right\} \leq \sqrt{\mathrm{e}}\, \exp \left(\max_{p=0,1,N}\left\{f(p)\right\}\right). \end{equation} We now want $\max_{p=0,1,N}\left\{f(p)\right\}=f(N)$. This is true if $1\geq f(0)/f(N)$ and $1\geq f(1)/f(N)$. By direct calculation using Eq. \eqref{def:f def}, we can solve these constraints for $b_2$. We find \begin{align} b_2&\leq \frac{d}{\sqrt{2\pi \mathrm{e}}\sigma}\frac{1}{\kappa}\frac{\ln(N+1)}{N^{1/2-1/4N}}\label{eq:f0/fN ineq}\\ b_2&\leq \frac{d}{\sqrt{2\pi \mathrm{e}}\sigma}\frac{1}{\kappa}\frac{\ln(N)}{(N+1)N^{-\left(N/2(N-1)+1/4(N-1)\right)}}. \label{eq:f1/fN ineq} \end{align} Therefore, we need $b_2$ to satisfy both Eqs. \eqref{eq:f0/fN ineq}, \eqref{eq:f1/fN ineq}, namely \begin{align}\label{eq:b2 condition} b_2&\leq \min\{b_L,b_R\},\\ b_L&:=\frac{d}{\sqrt{2\pi \mathrm{e}}\sigma}\frac{1}{\kappa}\frac{\ln(N+1)}{N^{1/2-1/4N}}\\ b_R&:= \frac{d}{\sqrt{2\pi \mathrm{e}}\sigma}\frac{1}{\kappa}\frac{\ln(N)}{(N+1)N^{-\left(N/2(N-1)+1/4(N-1)\right)}} \end{align} Thus if Eq. \eqref{eq:b2 condition} is satisfied, from Eqs. \eqref{eq:U N 1st max prob}, \eqref{eq:max f def}, and \eqref{eq:max written with f q}, we achieve \begin{align} \int_{-\infty}^\infty dz\left| U^{(N)}(z)\right| \leq & \frac{2\pi\sigma}{d} \left(|n_0+\gamma|+\upsilon+1\right)^N \sqrt{\mathrm{e}}\,\mathrm{e}^{f(N)}. \end{align} And hence plugging this into Eq. \eqref{eq:ep2 1sr up bound} \begin{align} |\bar\epsilon_2| &\leq Ad\,G \frac{2\pi\sigma\sqrt{\mathrm{e}}}{d}\left(\frac{(1-|\beta|)d}{2}\right)^{-N}\left( |n_0+\gamma|+\upsilon+1 \right)^N \mathrm{e}^{f(N)} \\ &\leq 2\pi\sigma A G \sqrt{\mathrm{e}} N^{1/4} \exp{\left(-N\ln\left(\frac{(1-|\beta|)d}{2}\right)+N\ln(|n_0+\gamma|+\upsilon+1)+N\ln\left( \frac{d}{\sqrt{2\pi}\sigma} \right)+N\ln\sqrt{N}-N/2\right)}\\ &\leq 2\pi\sigma A G \sqrt{\mathrm{e}} N^{1/4} \exp{\left(N\ln\left(\frac{|n_0+\gamma|+\upsilon+1}{(1-|\beta|)}\sqrt{\frac{2}{\pi\mathrm{e}}\frac{N}{\sigma^2}}\right)\right)}, \quad N=3,4,5,\ldots,\label{eq:ep in terms of N exp form} \end{align} Recall that our objective is to prove that $|\bar\epsilon_2|$ decays exponentially fast in $d$. For this, we will choose $N$ depending on the value of $d$. Although there is no explicit $d$ dependency in the exponential in Eq. \eqref{eq:ep in terms of N exp form}, recall that $\sigma$ is a function of $d$ and as such we will parametrize $N$ in terms of $\sigma$. For the exponential in Eq. \eqref{eq:ep in terms of N exp form} to be negative, we want \begin{equation} 0<\frac{|n_0+\gamma|+\upsilon+1}{(1-|\beta|)}\sqrt{\frac{2}{\pi\mathrm{e}}\frac{N}{\sigma^2}}<1 \end{equation} to hold. Solving for $N$ gives us \begin{equation}\label{eq:N up bound} N< \sigma^2\left(\frac{1-|\beta|}{|n_0+\gamma|+\upsilon+1} \right)^2 \frac{\pi \mathrm{e}}{2}. \end{equation} We thus set \begin{equation} N=N(\sigma)=\left\lfloor \sigma^2\left(\frac{1-|\beta|}{|n_0+\gamma|+\upsilon+1} \right)^2 \frac{\pi \mathrm{e}}{2\chi} \right\rfloor \end{equation} where $\chi$ is a free parameter we will choose such that Eq. \eqref{eq:N up bound} holds while optimizing the bound. Using the bounds $\lfloor x \rfloor \leq x$ for $x\in {\mathbbm{R}}$, and noting that Eq. \eqref{eq:ep in terms of N exp form} is monotonically increasing in $N=3,4,5,\ldots,$ we find plugging into Eq. \eqref{eq:ep in terms of N exp form} \begin{align} |\bar\epsilon_2| < \sigma A G (2\pi)^{5/4}\sqrt{\frac{\sigma\, (1-|\beta|)}{2(|n_0+\gamma|+\upsilon+1)}}\left(\frac{e}{\chi}\right)^{1/4} \,\exp\left(-\frac{\pi\mathrm{e}}{4\chi}\left(\frac{\sigma\, (1-|\beta|)}{|n_0+\gamma|+\upsilon+1} \right)^2 \ln(\chi)\right), \quad N(\sigma)=3,4,5,\ldots. \end{align} We now choose $\chi=\mathrm{e}$ to maximise $\ln (\chi)/\chi$ in the exponential and choose the parametrization \begin{equation}\label{eq:n0 with beta 0} n_0=\left(\frac{d-1}{2}\right)(1+\beta_0). \quad \beta_0\in[-1,1] \end{equation} Recalling Eq. \eqref{eq:gamma in terms of beta}, we thus achieve the final bound \begin{equation}\label{eq: exp decay ep2 bound} |\bar\epsilon_2| < \sigma A G (2\pi)^{5/4}\sqrt{\frac{\sigma\, (1-|\beta|)}{(d-1)|\beta_0-\beta|+2(\upsilon+1)}} \,\exp\left(-\frac{\pi}{4}\left(\frac{\sigma\, 2(1-|\beta|)}{(d-1)|\beta_0-\beta|+2(\upsilon+1)} \right)^2\right), \quad N(\sigma)=3,4,5,\ldots, \end{equation} with \begin{equation}\label{eq:N(d)} N=N(\sigma)=\left\lfloor \sigma^2\left(\frac{1-|\beta|}{(d-1)|\beta_0-\beta|/2+\upsilon+1} \right)^2 \frac{\pi}{2} \right\rfloor, \end{equation} where recall that $\beta\in(-1,1)$ is a free parametre which we can choose to optimise the bound. In the case that $N(\sigma)$ given by Eq. \eqref{eq:N(d)} does not satisfy $N(\sigma)=3,4,5,\ldots$, we will bound $|\bar\epsilon_2|$ by setting $N=3$ in Eq. \eqref{eq:ep in terms of N exp form}. Taking into account definitions \eqref{eq:gamma in terms of beta}, \eqref{eq:n0 with beta 0}, this gives \begin{equation}\label{eq: exp decay ep2 bound N(d)=0,1,2} |\bar\epsilon_2|< 2\pi\sqrt{\mathrm{e}}3^{1/4}\sigma A G \exp{\left(3\ln\left(\frac{|\beta_0-\beta|(d-1)/2+\upsilon+1}{1-|\beta|}\frac{1}{\sigma}\right)\right)}, \quad N(\sigma)=0,1,2. \end{equation} We will now work out an explicit bound for $b_2$ defined via Eq. \eqref{eq:b2 condition}. We will use $N(\sigma)$ (Eq. \eqref{eq:N(d)}) to achieve a definition of $b_2$ as a function of $\sigma$. We start by lower bounding $\min\{b_L,b_R\}$ \begin{align} \min\{b_L,b_R\}&=\frac{d}{\sqrt{2\pi}\kappa\sigma}\min\left\{ \frac{\ln(N+1)}{N^{1/2-1/4N}} , \frac{\ln(N)}{(N-1)N^{-(N/2(N-1)+1/4(N-1))}}\right\}\\ &\geq \frac{d}{\sqrt{2\pi}\kappa\sigma} \min\left\{ \inf_{x\geq 3}\frac{1}{x^{-1/4x}} , \inf_{x\geq 3}\frac{1}{x^{-1/4(x-1)}}\right\} \min\left\{ \frac{\ln(N+1)}{N^{1/2}} , \frac{\ln(N)}{(N-1)N^{-N/2(N-1)}}\right\}.\label{eq:intermediate 1 low bound min} \end{align} Note that the derivatives of $1/x^{-1/4x}$ and $1/x^{-1/4(x-1)}$ are both negative for $x\geq 3$ and thus $ \inf_{x\geq 3}\frac{1}{x^{-1/4x}}=\lim_{x\rightarrow \infty}\frac{1}{x^{-1/4x}}=1$ and $\inf_{x\geq 3}\frac{1}{x^{-1/4(x-1)}}=\lim_{x\rightarrow\infty }\frac{1}{x^{-1/4(x-1)}}=1$. Thus from Eq. \eqref{eq:intermediate 1 low bound min}, we find \begin{align} \min\{b_L,b_R\}&\geq \frac{d}{\sqrt{2\pi}\kappa\sigma} \min\left\{ \frac{\ln(N+1)}{N^{1/2}} , \frac{\ln(N)}{(N-1)N^{-(N-1)/2(N-1)}}\right\}\\ &\geq \frac{d}{\sqrt{2\pi}\kappa\sigma}\sqrt{N}\,\frac{\ln(N)}{N} \min\left\{ \frac{\ln(N+1)}{\ln(N)} , \frac{N}{(N-1)}\right\}\\ &\geq \frac{d}{\sqrt{2\pi}\kappa\,\sigma}\frac{\ln(N)}{\sqrt{N}}.\label{eq:lowebound min inter 2} \end{align} We now upper bound $N$. From Eqs. \eqref{eq:N(d)}, \eqref{eq:gamma in terms of beta}, \eqref{eq:n0 with beta 0}, it follows \begin{equation} N\leq \sigma^2\frac{\pi}{2}\left( \frac{1-|\beta|}{|n_0+\gamma|+\upsilon+1} \right)^2\\ \leq 2\pi \sigma^2 \left( \frac{1-|\beta|}{|\beta_0-\beta|(d-1)+2} \right)^2.\label{eq:up bound N} \end{equation} Now noting that \begin{equation} \frac{d}{dx}\frac{\ln(x)}{\sqrt{x}}=\frac{1}{x\sqrt{x}}\left(1-\frac{\ln(x)}{2}\right)<0 , \end{equation} for $x>\mathrm{e}^2\approx 7.39$, we can use Eq. \eqref{eq:up bound N} to lower bound Eq. \eqref{eq:lowebound min inter 2}. We find \begin{align} \min\{b_L,b_R\}\geq \mho,\quad \text{if } N(\sigma)\geq 8 \text{ and } \mho \geq 0 \end{align} where $N(\sigma)$ is given by Eq. \eqref{eq:N(d)} and we have defined \begin{equation}\label{def:mho} \mho:=\frac{d}{\sigma^2}\frac{1}{2\pi\,\kappa}\frac{|\beta_0-\beta|(d-1)+2}{1-|\beta|}\ln\left(2\pi\sigma^2\left( \frac{1-|\beta|}{|\beta_0-\beta|(d-1)+2} \right)\right). \end{equation} The constraint $\mho\geq 0$ is for consistency with the requirement $\upsilon\geq 0$. Recall Eq. \eqref{eq:b2 condition}, namely that $b_2$ must satisfy $\min\{b_L,b_R\} \geq b_2$, thus taking into account Eq. \eqref{eq:N(d)} and recalling that $\upsilon=b/b_2$ a consistent solution is \begin{equation}\label{eq:b_2 for N >=8} b_2=\mho\quad \text{if } \mathcal{N} \geq 8 \text{ and } \mho \geq 0 \end{equation} where we have defined \begin{equation}\label{def:mathcal N} \mathcal{N}:=\left\lfloor \sigma^2\left(\frac{1-|\beta|}{(d-1)|\beta_0-\beta|/2+b/\mho+1} \right)^2 \frac{\pi}{2} \right\rfloor \end{equation} When the if condition in Eq. \eqref{eq:b_2 for N >=8} is not satisfied, we can find a bound for $|\bar\epsilon_2|$ by setting $N=3$ in Eq. \eqref{eq:ep in terms of N exp form}. This gives us Eq. \eqref{eq:lemma sampled ft 2}. For $\upsilon'$, we bound $b_2$ by evaluating $\min\{b_L,b_R\}$ for $N=3$ using the bound Eq. \eqref{eq:lowebound min inter 2}. This gives us \begin{equation}\label{eq:b_2 for N<4,5,6} b_2=\frac{\ln(3)}{\sqrt{6\pi}\kappa}\frac{d}{\sigma},\quad \text{if } \mathcal{N}<8 \text{ and/or } \mho< 0 \end{equation} We will now workout what the constraint Eq. \eqref{eq: dev tilde theta constraint} with $a=1$, $b=\upsilon b_2$ imposes on potential function $V_0$. From definitions Eq. \eqref{eq:theta mischa def} and \eqref{eq:theta tilde mischa def}, we have \begin{equation} \tilde \theta (z)=\frac{2\pi}{d}\int_{-\frac{d z}{2\pi}+k_0-\Delta}^{-\frac{d z}{2\pi}+k_0} dy V_0(2\pi y/d)=\int_{-z+\tilde k_0-\tilde \Delta}^{- z+\tilde k_0} dy V_0(y)=\left( \mathcal{V}(- z+\tilde k_0)- \mathcal{V}(- z+\tilde k_0-\tilde\Delta)\right), \end{equation} where we have defined the re-scaled constants $\tilde k_0:=2\pi k_0/d$, $\tilde \Delta:=2\pi \Delta/d$ and have used the Fundamental Theorem of Calculus, to write the integral in terms of $\mathcal{V}$, where $\mathcal{V}^{(1)}(x)=V_0(x)$. We can now take the first $n$ derivatives of $\tilde\theta$: \begin{align} \tilde\theta^{(n)}(z)&= \left(-1\right)^n\frac{d^n}{dy^n}\left( \mathcal{V}(y)- \mathcal{V}(y-\tilde\Delta)\right)\\ &= \left(-1\right)^{n-1}\frac{d^{n-1}}{dy^{n-1}}\left( V_0(y-\tilde\Delta)-V_0(y)\right)=\left(-1\right)^{n-1}\frac{d^{n-1}}{dx^{n-1}} V_0(x)\Bigg{]}_y^{y-\tilde\Delta}\\ &=\left(-1\right)^{n-1} V_0^{(n-1)}(x)\Big{]}_y^{y-\tilde\Delta},\quad n\in {\mathbbm{N}}^+ \end{align} where $y:=-z+\tilde k_0$. Hence \begin{align} \left|\tilde\theta^{(n)}(z)\right|^{1/n}&= \left| V_0^{(n-1)}(x)\Big{]}_{y}^{y-\tilde\Delta} \right|^{1/n}\leq \left(\max_{x_1,x_2\in[0,2\pi]} \left| V_0^{(k-1)}(x)\Big{]}_{x_1}^{x_2} \right|\,\right)^{1/k}\\ &\leq \sup_{k\in {\mathbbm{N}}^+}\left(2\max_{x\in[0,2\pi]} \left| V_0^{(k-1)}(x)\right|\,\right)^{1/k} , \quad \forall\; n\in{\mathbbm{N}}^+ \text{ and } \forall\; z\in {\mathbbm{R}}. \end{align} Thus from Eqs. \eqref{eq: dev tilde theta constraint}, \eqref{eq:b upsilon and b_2 def} we conclude that $b\geq 0$ is any non negative number satisfying \begin{equation}\label{eq: b_1 explicit} b\geq\; \sup_{k\in {\mathbbm{N}}^+}\left(2\max_{x\in[0,2\pi]} \left| V_0^{(k-1)}(x)\right|\,\right)^{1/k}. \end{equation} and \begin{equation}\label{eq: upsilon explicit} \upsilon= \frac{b}{b_2}= \frac{\sigma^2}{d}\frac{2\pi\,\kappa(1-|\beta|)}{|\beta_0-\beta|(d-1)+2}\frac{b}{\ln\left(2\pi\sigma^2\left( \frac{1-|\beta|}{|\beta_0-\beta|(d-1)+2} \right)\right)}, \quad \text{if } \mathcal{N}\geq 8 \text{ and } \mho \geq 0 \end{equation} where we have used Eq. \eqref{eq:b_2 for N >=8}. For $\mathcal{N}<8$, we use Eq. \eqref{eq:b_2 for N<4,5,6} to achieve \begin{equation} \upsilon= \frac{b}{b_2}=b\,\frac{\sqrt{6\pi}\kappa}{\ln(3)}\frac{\sigma}{d},\quad \text{if } \mathcal{N}<8 \text{ and/or } \mho< 0. \end{equation} We are now ready to state the final bound. From Eqs. \eqref{eq: exp decay ep2 bound}, \eqref{eq:b_2 for N >=8} it follows \begin{equation} |\bar\epsilon_2|\leq (2\pi)^{5/4}\sigma A G\sqrt{\frac{\mathrm{e}\,(1-|\beta|)\,\sigma}{|\beta-\beta_0|(d-1)+2(\upsilon+1)}}\exp\left(x \right),\quad \text{if } \mathcal{N}\geq 8\text{ and } \mho \geq 0, \end{equation} with \begin{equation} x=-\pi\frac{\left(1-|\beta|\right)^2}{\left(\left(|\beta-\beta_0|(d-1)+2\right)\frac{d}{\sigma^2}+\frac{4\pi\kappa(1-|\beta|)b}{\left(|\beta_0-\beta|(d-1)+2\right)\ln\left (\frac{2\pi\sigma^2(1-|\beta|)}{|\beta_0-\beta|(d-1)+2}\right)}\right)^2} \left(\frac{d}{\sigma}\right)^2, \end{equation} where $\upsilon$ is given by Eq. \eqref{eq: upsilon explicit}, $\mathcal{N}$ by \eqref{def:mathcal N}, and $b$ by \eqref{eq: b_1 explicit}. Recall that $\beta\in(-1,1)$ is a free parameter which we may choose to optimize the bound. For $\mathcal{N}< 8$, the bound is achieved from Eqs. \eqref{eq: exp decay ep2 bound N(d)=0,1,2}, \eqref{eq:b_2 for N<4,5,6} \begin{equation} |\bar\epsilon_2|\leq \frac{3^{7/4}8}{\sqrt{2\pi}\mathrm{e}}A\,G \frac{\left( |\beta_0-\beta|(d-1)+2\left(\frac{\kappa\sqrt{6\pi}}{\ln(3)}\frac{\sigma}{d}\,b+1\right) \right)^3}{(1-|\beta|)^3}\frac{1}{\sigma^2},\quad \text{for } \mathcal{N}< 8\text{ and/or } \mho< 0. \end{equation} The optimal choice of $\beta$ might depend on $\sigma$ and $d$, however, whenever $\beta_0\neq \pm 1$, i.e. mean energy of the initial clock state is not at one of the extremal points $0,d$, asymptotically we see that the optimal choice is always $\beta_0=\beta$. Taking this into account and in order to simplify the bound, we will set $\beta_0=\beta$ to achieve \begin{equation} |\bar\epsilon_2|\leq (2\pi)^{5/4}\sigma^{3/2} A G\sqrt{\frac{\mathrm{e}}{2}\frac{\alpha_0}{(\upsilon+1)}}\exp\left(-\frac{\pi}{4}\left(\frac{\alpha_0}{\frac{d}{\sigma^2}+\frac{\pi\kappa\alpha_0}{\ln\left(\pi\alpha_0\sigma^2\right)}b}\right)^2 \left(\frac{d}{\sigma}\right)^2 \right),\quad \text{if } \mathcal{N}\geq 8 \text{ and } \mho\geq 0, \end{equation} and \begin{equation} |\bar\epsilon_2|\leq \frac{3^{7/4}8}{\sqrt{2\pi}\mathrm{e}}\frac{A\,G}{\alpha_0^3}\left(\frac{\kappa\sqrt{6\pi}}{\ln(3)}b+\frac{d}{\sigma}\right)^3 \left( \frac{\sigma}{d^3} \right),\quad \text{if } \mathcal{N}< 8 \text{ and/or } \mho< 0, \end{equation} where we have defined \begin{equation} \alpha_0:=1-|\beta_0|\in (0,1), \end{equation} which can be written in terms of $n_0$ as \begin{align} \alpha_0&=\min\{ 1+|\beta_0|,1-|\beta_0|\}=\min\{ 1+\beta_0,1-\beta_0\}\\ &=\left(\frac{2}{d-1}\right) \min\{n_0,(d-1)-n_0\}\\ &=1-\left|1-n_0\,\left(\frac{2}{d-1}\right)\right|,\quad \text{for } n_0\in(0,d-1) \end{align} where we have used Eq. \eqref{eq:n0 with beta 0}. Furthermore, $\mho$ and $\mathcal{N}$ can also be simplified when $\beta=\beta_0$. From Eqs. \eqref{def:mho} we find \begin{equation} \mho=\frac{d}{\sigma^2}\frac{\ln(\pi\alpha_0\sigma^2)}{\pi\alpha_0\kappa}, \end{equation} while from \eqref{def:mathcal N} it follows \begin{equation} \mathcal{N}=\left\lfloor \frac{\pi\alpha_0^2}{2\left(\frac{\pi\kappa\alpha_0}{\ln(\pi\alpha_0\sigma^2)}b+\frac{d}{\sigma^2}\right)^2} \left(\frac{d}{\sigma}\right)^2 \right\rfloor. \end{equation} \end{proof} Before we proceed, we now define a generalization of Def. \ref{def:Gaussian clock states} which includes a potential dependent phase. \begin{definition}\label{def:gauss clock steate with pot defintion}\emph{(Gaussian clock states with Potential)}. Let $\Lambda_{ V_0,\sigma,n_0}$ be the following space of states in the Hilbert space of the $d$ dimensional clock, \begin{equation}\label{eq:lambda set with pot def} \Lambda_{ V_0,\sigma,n_0}=\bigg\{ \ket{\bar{\Psi}(k_0,\Delta)}\in\mathcal{H}_\textup{c},\quad k_0,\Delta\in{\mathbbm{R}}\bigg\}, \end{equation} where \begin{align}\label{def:gauss clock steate with pot} \ket{\bar{\Psi}(k_0,\Delta)} = \sum_{\mathclap{\substack{k\in \mathcal{S}_d(k_0)}}}\mathrm{e}^{-\mathrm{i}\Theta(\Delta;k)}\psi(k_0;k)\ket{\theta_k}, \end{align} where $\Theta$, $\psi$ are defined in Eqs. \eqref{eq: Theta def}, \eqref{analyticposition}, respectively and depend on the parameters $\sigma \in (0,d)$, $n_0 \in (0,d-1)$ and function $ V_0$, which is defined in Def. \ref{def:continous pot}. $\mathcal{S}_d(k_0)$ is given by Eq. \eqref{eq: mathcal S def}.\\ In the special case that $\ket{\bar{\Psi}(k_0,\Delta)}$ is normalised, it will be denoted \begin{equation} \ket{\bar{\Psi}_\textup{nor}(k_0,\Delta)}=\ket{\bar{\Psi}(k_0,\Delta)}, \end{equation} and $A$ (see Eq. \eqref{analyticposition}) will satisfy Eq. \eqref{eq:A normalised}, s.t. $\braket{\bar{\Psi}_\textup{nor}(k_0,\Delta)|\bar{\Psi}_\textup{nor}(k_0,\Delta)}=1$. \end{definition} \begin{remark} In general $\Lambda_{\sigma,n_0}\subseteq \Lambda_{ V_0,\sigma,n_0}$ with equality if $\Delta=0$ or $ V_0(x)=0$ for all $x\in{\mathbbm{R}}$ (compare Eqs. \eqref{eq:lambda set def} and \eqref{eq:lambda set with pot def}). \end{remark} \begin{lemma}\label{lemm:1}\emph{(Infinitesimal evolution under the clock Hamiltonian).} The action of the unitary operator $e^{-i\frac{T_0}{d} \delta \hat{H}_c}$ on an element of $\Lambda_{ V_0,\sigma,n_0}$ may be approximated by a translation by $\delta\geq 0$ on the continuous extension of the clock state. Precisely speaking, \begin{align} e^{-i\frac{2\pi}{\omega d} \delta \hat{H}_c} \sum_{l\in\mathcal{S}_d(k_0)} \mathrm{e}^{-\mathrm{i} \Theta(\Delta;l)}\psi(k_0;l) \ket{\theta_l} &= \sum_{l\in\mathcal{S}_d(k_0)} \mathrm{e}^{-\mathrm{i} \Theta(\Delta;l-\delta)}\psi(k_0;l-\delta) \ket{\theta_l} + \ket{\epsilon}, \end{align} where the $l_2$ norm of the error $\ket{\epsilon}$ is bounded by \begin{align} \ltwo{\ket{\epsilon}} &\leq \delta \epsilon_T \sqrt{d}+C_1\delta^2\sqrt{d},\\ \epsilon_T &< \begin{cases} |\bar\epsilon_2|+2 A\left(\frac{2\pi}{1-\mathrm{e}^{-\pi}}+ \frac{\left(b+\frac{2\pi}{d}\right)}{1-\mathrm{e}^{-\pi d}} + \left( 2\pi+\pi d+ \frac{1}{d}\right)\right)\mathrm{e}^{-\frac{\pi}{4}d} &\mbox{if } \sigma=\sqrt{d}\\ |\bar\epsilon_2|+2 A\left(\frac{2\pi}{1-\mathrm{e}^{-\frac{\pi d}{\sigma^2}}}+ \frac{ \left(b+\frac{2\pi}{d}\right)}{1-\mathrm{e}^{-\pi \frac{d^2}{\sigma^2}}} + \left( 2\pi\frac{d}{\sigma^2}+\pi\frac{d^2}{\sigma^2}+ \frac{1}{d}\right)\right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}} &\mbox{otherwise} \end{cases}\\ &=\begin{cases} \left(\mathcal{O}(b)+\mathcal{O}\left( \frac{d^{3/2}}{\bar\upsilon+1}\right)^{1/2}+\mathcal{O}\left(d\right)\right) \exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(1+\bar\upsilon\right)^2} d \right) &\mbox{if } \sigma=\sqrt{d}\\ \left(\mathcal{O}(b)+\mathcal{O}\left( \frac{\sigma^3}{\bar\upsilon \sigma^2/d+1}\right)^{1/2}+\mathcal{O}\left(\frac{d^2}{\sigma^2}\right)\right) \exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(\frac{d}{\sigma^2}+\bar\upsilon\right)^2} \left(\frac{d}{\sigma}\right)^2 \right)+\mathcal{O}\left(\frac{d^2}{\sigma^2}\right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}} &\mbox{otherwise,} \end{cases} \end{align} with $|\bar\epsilon_2|$ given by Lemma \ref{lem:the crucial fourierbound with potential} and $C_1$ is $\delta$ independent. $\Theta$, $\psi$ are given by Eqs. \eqref{eq: Theta def}, \eqref{analyticposition} respectively. \end{lemma} \emph{Intuition}. This is simply the statement that for the class of Gaussian states that we have chosen, the effect of the clock Hamiltonian for an infinitesimal time is approximately the shift operator w.r.t. the angle space. The proof will follow along similar lines to that of Lemma \ref{infinitesimaltimetranslation} with the main difference being that we will now have to resort to Lemma \ref{lem:the crucial fourierbound with potential} in order to bound $\epsilon_2$ where as before bounding $\epsilon_2$ was straightforward and accomplished directly in the proof. \begin{proof} \begin{align} \mathrm{e}^{-\frac{\mathrm{i} T_0}{ d} \delta \hat{H}_c} \ket{\bar{\Psi}(k_0,\Delta)} &= e^{-\frac{\mathrm{i} 2\pi}{d} \delta \sum_{m=0}^{d-1} m\ket{E_m}\bra{E_m}} \sum_{k\in\mathcal{S}_d(k_0)} \!\!\! \mathrm{e}^{-\mathrm{i} \Theta(\Delta;k)}\psi(k_0;k) \ket{\theta_k}. \end{align} Switching the state to the basis of energy states, applying the Hamiltonian, and switching back (Eqs. \eqref{finitetimestates} and \eqref{finiteenergystates}), \begin{align} \mathrm{e}^{-\frac{\mathrm{i} T_0}{d} \delta \hat{H}_c} \ket{\bar{\Psi}(k_0,\Delta)} &= \sum_{k,l\in\mathcal{S}_d(k_0)} \!\!\! \mathrm{e}^{-\mathrm{i} \Theta(\Delta;k)}\psi(k_0;k) \left( \frac{1}{d} \sum_{n=0}^{d-1} e^{-i2\pi n(k+\delta-l)/d} \right) \ket{\theta_l} \end{align} We label the above state as $\ket{\bar{\Psi}^{exact}_\delta}$. On the other hand we label as $\ket{\bar{\Psi}^{approx}_\delta}$ the following expression, \begin{align} \ket{\bar{\Psi}^{approx}_\delta} &= \sum_{l\in\mathcal{S}_d(k_0)} \mathrm{e}^{-\mathrm{i} \Theta(\Delta;l-\delta)}\psi(k_0;l-\delta) \ket{\theta_l} \end{align} which is simply a translation by $+\delta$ of the continuous extension of the clock state. Both the coefficients $\braket{\theta_l|\bar{\Psi}^{exact}_\delta}$ and $\braket{\theta_l|\bar{\Psi}^{approx}_\delta}$ are twice differentiable with respect to $\delta$. For $\braket{\theta_l|\bar{\Psi}^{exact}_\delta}$ this is clear. In the case of $\braket{\theta_l|\bar{\Psi}^{approx}_\delta}$ we note that it is a function of the derivative of \begin{equation} \Theta(\Delta;l-\delta) \int_{l-\delta-\Delta}^{l-\delta} V_d(x')dx', \end{equation} with respect to $\delta$, and thus due to the fundamental theorem of calculus and the fact the $ V_d$ is a smooth function (and periodic), it follows that $\braket{\theta_l|\bar{\Psi}^{approx}_\delta}$ is differentiable with respect to $\delta$. By Taylor's remainder theorem, the difference can be expressed as \begin{equation}\label{Taylorshift} \braket{\theta_l|\bar{\Psi}^{approx}_\delta} - \braket{\theta_l|\bar{\Psi}^{exact}_\delta} = \braket{\theta_l|\bar{\Psi}^{approx}_0} - \braket{\theta_l|\bar{\Psi}^{exact}_0} + \delta \frac{d \left( \braket{\theta_l|\bar{\Psi}^{approx}_\delta} - \braket{\theta_l|\bar{\Psi}^{exact}_\delta} \right)}{d\delta} \bigg|_{\delta=0} + R(\delta), \end{equation} where \begin{equation}\label{Taylorshift C bound} \left| R(\delta) \right| \leq \frac{\delta^2}{2} \left( \text{max}_{|t|\leq|\delta|} \left| \frac{d^2 \left( \braket{\theta_l|\bar{\Psi}^{approx}_t} - \braket{\theta_l|\bar{\Psi}^{exact}_t} \right)}{dt^2} \right| \right) \leq C_1 \delta^2, \end{equation} and where $C_1$ is independent of $\delta$ because the second derivatives of both $\braket{\theta_l|\bar{\Psi}^{exact}_\delta}$ and $\braket{\theta_l|\bar{\Psi}^{approx}_\delta}$ w.r.t. $\delta$ are bounded for $\delta\in {\mathbbm{R}}$. The zeroth-order term vanishes, i.e. $\braket{\theta_l|\bar{\Psi}^{approx}_0} - \braket{\theta_l|\bar{\Psi}^{exact}_0}$ since $\ket{\bar{\Psi}^{exact}_0} = \ket{\bar{\Psi}^{approx}_0}$. For the first order term, \begin{align} \frac{d\braket{\theta_l|\bar{\Psi}^{approx}_\delta}}{d\delta}\bigg|_{\delta=0} &= \frac{d\;}{d\delta} \mathrm{e}^{-\mathrm{i} \Theta(\Delta;l-\delta)}\psi(k_0;l-\delta)\bigg|_{\delta=0} \\ \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} \bigg|_{\delta=0} &= \left[ \frac{d}{d\delta} \sum_{k\in\mathcal{S}_d(k_0)}\mathrm{e}^{-\mathrm{i} \Theta(\Delta;k)}\psi(k_0;k) \left( \frac{1}{d} \sum_{n=0}^{d-1} e^{-i2\pi n(k+\delta-l)/d} \right) \right]_{\delta=0} \\ &= \left( \frac{-i 2\pi}{d} \right) \sum_{k\in\mathcal{S}_d(k_0)}\mathrm{e}^{-\mathrm{i} \Theta(\Delta;k)}\psi(k_0;k) \left( \frac{1}{d} \sum_{n=0}^{d-1} n e^{-i2\pi n(k-l)/d} \right) \end{align} One can replace the finite sum over $k$ as an infinite sum, and bound the difference using Lemma \ref{G0} \begin{align} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d} \right) \sum_{k=-\infty}^\infty \mathrm{e}^{-\mathrm{i} \Theta(\Delta;k)}\psi(k_0;k) \left( \frac{1}{d} \sum_{n=0}^{d-1} n e^{-i2\pi n(k-l)/d} \right) + \epsilon_1,\label{eq:div exact before poisson} \end{align} where, \begin{align} \left| \epsilon_1 \right| &\leq \frac{2\pi}{d} \left( \sum_{k\in {\mathbbm{Z}} - \mathcal{S}_d(k_0)} \left|\psi(k_0;k)\right| \right) \frac{1}{d} \left( \sum_{n=0}^{d-1} n \right) \\ &< 4\pi A\left(\frac{\mathrm{e}^{-\frac{\pi d^2}{4\sigma^2}}}{1-\mathrm{e}^{-\frac{\pi d}{\sigma^2}}}\right) & \text{by Lemma \ref{G0}}. \end{align} Applying the Poisson summation formula (Corollary \ref{poissonsummation}) to the sum in Eq. \eqref{eq:div exact before poisson} we achieve \begin{equation} \sum_{k=-\infty}^\infty \mathrm{e}^{-\mathrm{i} \Theta(\Delta;k)}\psi(k_0;k) e^{-i2\pi nk/d}=\sum_{s=-\infty}^\infty \tilde{\psi}(k_0,\Delta;n+sd), \end{equation} where $\tilde{\psi}$ is given by Def. \ref{def:analyticmomentum control version}. Thus we have \begin{align} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d} \right) \frac{1}{\sqrt{d}} \sum_{n=0}^{d-1} \sum_{s=-\infty}^{\infty} \tilde{\psi}(k_0,\Delta;n+sd) \; n e^{i2\pi nl/d} + \epsilon_1. \end{align} Since $\sum_{n=0}^{d-1} \sum_{s=-\infty}^{\infty} f(n+sd) = \sum_{n=-\infty}^\infty f(n)$, one can manipulate the expression accordingly, \begin{align} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d} \right) \frac{1}{\sqrt{d}} \left( \sum_{n=-\infty}^\infty \tilde{\psi}(k_0,\Delta;n) \; n e^{i2\pi nl/d} - \sum_{n=0}^{d-1} \sum_{s=-\infty}^{\infty} \tilde{\psi}(k_0,\Delta;n+sd) \;sd e^{i2\pi nl/d} \right) + \epsilon_1 \end{align} The second summation is a small contribution and has been bound in Lemma \ref{lem:the crucial fourierbound with potential}, \begin{align} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} \bigg|_{\delta=0} &= \left( \frac{-i 2\pi}{d} \right) \frac{1}{\sqrt{d}} \sum_{n=-\infty}^\infty \tilde{\psi}(k_0,\Delta;n) \; n e^{i2\pi nl/d} + \epsilon_2 + \epsilon_1, \end{align} where \begin{align} \epsilon_2 &=\frac{\mathrm{i} 2\pi}{\sqrt{d}}\sum_{n=0}^{d-1}\sum_{k=-\infty}^\infty k\, \mathrm{e}^{\mathrm{i} 2\pi n l/d}\, \tilde \psi(k_0,\Delta;n+k d). \end{align} On the remaining sum, apply Eq. \eqref{eq: inv Fourier = dev} to write $\tilde \psi(k_0,\Delta;n) n$ in terms of the derivative of the Fourier transform $\tilde \psi(k_0,\Delta;n)$, followed by applying the Poisson summation formula, \begin{align} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} \bigg|_{\delta=0} &= \sum_{m=-\infty}^\infty \frac{d\;}{d\delta}\left(\mathrm{e}^{-\mathrm{i}\Theta(\Delta;l-\delta+md)}\psi(k_0;l-\delta+md)\right) \bigg|_{\delta=0} + \epsilon_1 + \epsilon_2. \end{align} Replacing the sum by the $m=0$ term, and bounding the difference \begin{align} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} &= \frac{d\;}{d\delta} \left(\mathrm{e}^{-\mathrm{i}\Theta(\Delta;l-\delta)}\psi(k_0;l-\delta)\right)\bigg|_{\delta=0} + \epsilon_1 + \epsilon_2 + \epsilon_3, \label{deltacoeff} \end{align} where \begin{align} \left| \epsilon_3 \right| &\leq \sum_{m\in{\mathbbm{Z}}-\{0\}} \abs{\frac{d\;}{d\delta} \left(\mathrm{e}^{-\mathrm{i}\Theta(\Delta;l-\delta+md)}\psi(k_0;l-\delta+md)\right)}_{\delta=0} \\ &\leq \sum_{m\in{\mathbbm{Z}}-\{0\}} \left( \abs{ V_d(l+md)- V_d(l+md-\Delta) }\abs{\psi(k_0;l+md)}+\abs{\frac{d\,}{d\delta} \psi(k_0;l-\delta+md) }_{\delta=0} \right)\\ &\leq \sum_{m\in{\mathbbm{Z}}-\{0\}} \left( b\abs{ \psi(k_0;l+md)}+\abs{ \frac{d\,}{d\delta}\psi(k_0;l-\delta+md)}_{\delta=0} \right)\\ &< \begin{cases} 2 A\left( \frac{\left(b+\frac{2\pi}{d}\right)}{1-\mathrm{e}^{-\pi d}} + \left( 2\pi+\pi d+ \frac{1}{d}\right)\right)\mathrm{e}^{-\frac{\pi}{4}d} &\mbox{if } \sigma=\sqrt{d}\\ 2 A\left( \frac{ \left(b+\frac{2\pi}{d}\right)}{1-\mathrm{e}^{-\pi \frac{d^2}{\sigma^2}}} + \left( 2\pi\frac{d}{\sigma^2}+\pi\frac{d^2}{\sigma^2}+ \frac{1}{d}\right)\right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}} &\mbox{otherwise,} \end{cases} \end{align} for all $l\in\mathcal{S}_d(k_0)$ and where $b$ is defined in \eqref{eq:b def eq}. To achieve the last two lines, we have used Lemmas \ref{G0}, \ref{G1}. Thus the first order term in Eq. (\ref{Taylorshift}) is composed of the terms $\epsilon_1,\epsilon_2,\epsilon_3$, and are bounded by \begin{align}\label{eq:total epsilon} \frac{d \braket{\theta_l|\bar{\Psi}^{exact}_\delta}}{d\delta} - \frac{d \braket{\theta_l|\bar{\Psi}^{approx}_\delta}}{d\delta} &= \epsilon_1 + \epsilon_2 + \epsilon_3 =: \epsilon_T. \end{align} Note that $\epsilon_T$ is independent of the index $l$. If we now define the error in the state as $\ket{\epsilon} = \ket{\bar{\Psi}^{approx}_\delta} - \ket{\bar{\Psi}^{exact}_\delta}$, then from the properties of the norm, and Eqs. \eqref{Taylorshift}, \eqref{Taylorshift C bound}, \eqref{eq:total epsilon}, \begin{align} \ltwo{\ket{\epsilon}}^2 &\leq \sum_{l\in\mathcal{S}_d(k_0)} \left( \bra{\bar{\Psi}_\delta^\text{approx}}- \bra{\bar{\Psi}_\delta^\text{exact}}\right)\ketbra{\theta_l}{\theta_l}\left( \ket{\bar{\Psi}_\delta^\text{approx}}- \ket{\bar{\Psi}_\delta^\text{exact}}\right)\\ &\leq \sum_{l\in\mathcal{S}_d(k_0)} \left| \braket{\theta_l|\bar{\Psi}_\delta^\text{approx}}- \braket{\theta_l|\bar{\Psi}_\delta^\text{exact}}\right|^2\\ &\leq \left| \delta \epsilon_T+C_1\delta^2\right|^2 d. \end{align} \end{proof} \begin{lemma}\label{lem:2 infinitesimal pot} \emph{(Infinitesimal evolution under the interaction potential).} The action of the unitary operator $e^{-i\frac{T_0}{d} \delta \hat{V}_d}$ on any state $\ket{\Phi} \in \mathcal{H}_c$ and $\delta\geq 0$, may be approximated by the following transformation, \begin{align}\label{eq:lemma eq for pot infini} \mathrm{e}^{-\mathrm{i}\frac{T_0}{ d} \delta \hat{V}_d} \ket{\Phi} &= \sum_{l\in\mathcal{S}_d(k_0)} \braket{\theta_l|\Phi} \mathrm{e}^{-\mathrm{i} \Theta(\delta;l)} \ket{\theta_l} + \ket{\epsilon}, \end{align} where the $l_2$ norm of the error $\ket{\epsilon}$ is bounded by \begin{align} \ltwo{\ket{\epsilon}} \leq C_2 \delta^2, \end{align} with $C_2$ being $\delta$ independent. $\hat V_d$ is defined in Def. \ref{def:interaction pot def}. \end{lemma} \begin{proof} Since the operator $\hat{V}_d$ is diagonal in the $\{\ket{\theta_k}\}$ basis, \begin{align}\label{eq:intermidiate theta basis pot infini} \mathrm{e}^{-\frac{\mathrm{i} T_0}{d} \delta \hat{V}_d} \ket{\Phi} = \mathrm{e}^{-\mathrm{i} \delta \sum_k {V}_d(k) \ket{\theta_k}\bra{\theta_k}} \sum_{l\in\mathcal{S}_d(k_0)} \ket{\theta_l}\braket{\theta_l|\Phi} = \sum_{l\in\mathcal{S}_d(k_0)} \ket{\theta_l}\braket{\theta_l|\Phi} e^{-\mathrm{i}\delta {V}_d(l)}. \end{align} We wish to replace $e^{- \mathrm{i} \delta {V}_d(l)}$ by $e^{-\mathrm{i} \int_{l-\delta}^l V_d(x^\prime) dx^\prime}$. The difference may be bound using Taylor's theorem (to second order), \begin{align}\label{eq: taylors pot infini} \mathrm{e}^{- \mathrm{i} \delta {V}_d(l)} - \mathrm{e}^{-\mathrm{i} \int_{l-\delta}^l V_d(x^\prime) dx^\prime} = A + B\delta + R_l(\delta), \end{align} where \begin{align}\label{Taylorshift C bound pot lemma} \left| R_l(\delta) \right|& \leq \frac{\delta^2}{2} \left( \text{max}_{|t|\leq|\delta|} \left| \frac{d^2 \left(\mathrm{e}^{- \mathrm{i} t {V}_d(l)} - \mathrm{e}^{-\mathrm{i} \int_{l-t}^l V_d(x^\prime) dx^\prime} \right)}{dt^2} \right| \right)\\ &\leq \frac{\delta^2}{2} \left( \text{max}_{t\in{\mathbbm{R}}} \left| \frac{d^2 \left(\mathrm{e}^{- \mathrm{i} t {V}_d(l)} - \mathrm{e}^{-\mathrm{i} \int_{l-t}^l V_d(x^\prime) dx^\prime} \right)}{dt^2} \right| \right)\\ &\leq C_2 \delta^2, \quad \forall\, l\in\mathcal{S}_d(k_0) \end{align} and $C_2$ is $\delta$ and $l$ independent. Such a $C_2$ exists, since $ V_d$ is periodic and smooth. Furthermore, by direct calculations $A=B=0$. Thus using the properties of the $l_2$ norm and Eqs. \eqref{eq:lemma eq for pot infini}, \eqref{eq:intermidiate theta basis pot infini}, \eqref{eq: taylors pot infini}, we find \begin{align} \ltwo{\ket{\epsilon}}^2 &\leq \left( \sum_{l\in\mathcal{S}_d(k_0)} |\! \braket{\theta_l|{\Phi}}\!|^* R_l(\delta)^* \bra{\theta_l}\right)\left( \sum_{k\in\mathcal{S}_d(k_0)} |\!\braket{\theta_k|\Phi}\!| R_k(\delta) \ket{\theta_k}\right)\\ &\leq \sum_{l\in\mathcal{S}_d(k_0)}|\!\braket{\theta_l|\Phi}\!| |R_l(\delta)|^2\\ &\leq C_2^2 \delta^4. \end{align} \end{proof} \begin{lemma}\label{Moving the clock through finite time, within unit angle} \emph{(Moving the clock through finite time, within unit angle).} Let $k_0 \in {\mathbbm{R}}$, and let $a \geq 0$ be s.t. $\mathcal{S}_d(k_0)=\mathcal{S}_d(k_0+a)$, where $\mathcal{S}_d$ is defined by Eq. \eqref{eq: mathcal S def}. Then the effect of the joint Hamiltonian $\hat{H}_c + \hat{V}_d$ for the time $\frac{T_0}{d}a$ on $\ket{\bar{\Psi}(k_0;\Delta)} \in \Lambda_{ V_0,\sigma,n_0}$ is approximated by \begin{align} \mathrm{e}^{-\mathrm{i} \frac{T_0}{d}a (\hat{H}_c + \hat{V}_d)} \ket{\bar{\Psi}(k_0,\Delta)} &= \ket{\bar{\Psi}(k_0+a,\Delta+a)} + \ket{\epsilon}, \end{align} where the $l_2$ norm of the error $\ket{\epsilon}$ is bounded by \begin{align} \ltwo{\ket{\epsilon}} \leq a\, \epsilon_T \sqrt{d}, \end{align} and $\epsilon_T$ is given by Lemma \ref{lemm:1}. \end{lemma} \emph{Intuition.} Here is the core of the proof. By applying the Lie product formula, we show that the effect of the combined clock and interaction Hamiltonians is to simply shift the continuous extension of the clock state, while simultaneously adding a phase function that is the integral of the potential that the clock passes through. The discrete clock is thus seen to mimic the idealised momentum clock $\hat H=\hat p$. Here it is assumed the mean of the state does not pass through an integer (the full result follows later). \begin{proof} Consider the sequential application of $\mathrm{e}^{-\mathrm{i} \frac{T_0}{d} \delta \hat{H}_c}$ followed by $\mathrm{e}^{-\mathrm{i} \frac{T_0}{d} \delta \hat{V}_d}$ on $\ket{{\bar{\Psi}}(k_0,\Delta)}$. From the previous two Lemmas \ref{lemm:1}, \ref{lem:2 infinitesimal pot}, \begin{align} \mathrm{e}^{-\mathrm{i} \frac{T_0}{d} \delta \hat{V}_d} \mathrm{e}^{-\mathrm{i} \frac{T_0}{d} \delta \hat{H}_c} \ket{\bar{\Psi}(k_0,\Delta)} = \ket{\bar{\Psi}(k_0+\delta,\Delta+\delta)} + \ket{\epsilon_\delta}, \end{align} where \begin{align} \ltwo{\ket{\epsilon}} \leq \delta \epsilon_T \sqrt{d}+\delta^2\left(C_1\sqrt{d} + C_2 \right). \end{align} where we combined the errors using the lemma on compiling unitary errors, Lemma \ref{unitaryerroraddition}. Consider the above transformation repeated $m$ times on the state. Combining the errors using Lemma \ref{unitaryerroraddition} as before, \begin{align} \left( \mathrm{e}^{-\mathrm{i} \frac{T_0}{d} \delta \hat{V}_d} \mathrm{e}^{-\mathrm{i} \frac{T_0}{d} \delta \hat{H}_c} \right)^m \ket{\bar{\Psi}(k_0,\Delta)} &=\ket{ \bar{\Psi} \left(k_0+ m\delta, \Delta+m\delta \right)} + \ket{\epsilon^{(m)}},\end{align} where \begin{align} \ltwo{\ket{\epsilon^{(m)}}} \leq m\delta \epsilon_T \sqrt{d}+m\delta^2\left(C_1\sqrt{d} + C_2 \right). \end{align} This holds as long as the center of the state $k_0+m\delta$ has not crossed an integer value, i.e. if $d$ is even, $\lfloor k_0\rfloor= \lfloor k_0+m\delta\rfloor$, or if $d$ is odd $\lfloor k_0+1/2\rfloor= \lfloor k_0+1/2+m\delta\rfloor$; else $\mathcal{S}_d(k_0)\neq\mathcal{S}_d(k_0+m\delta)$ and the previous Lemmas \ref{lemm:1}, \ref{lem:2 infinitesimal pot} in this section will not hold. To arrive at the lemma, set $\delta = a/m$, so that \begin{align}\label{eq:to power m with pot} \left( e^{-i \frac{T_0}{d} \frac{a}{m} \hat{V}_d} e^{-i \frac{T_0}{d} \frac{a}{m} \hat{H}_c} \right)^m \ket{\bar{\Psi}(k_0,\Delta)} &= \ket{\bar{\Psi} \left(k_0+ a, \Delta+a \right)} + \ket{\epsilon^{(m)}}, \end{align} where \begin{align} \ltwo{\ket{\epsilon^{(m)}}} \leq a \epsilon_T \sqrt{d}+m\left(\frac{a}{m}\right)^2\left(C_1\sqrt{d} + C_2 \right). \end{align} The Lie-Product formlula \cite{Lie}, states that for all $n \times n$ complex matrices $A$ and $B$, \begin{equation}\label{eq:lie} \mathrm{e}^{A+B} = \lim_{N \rightarrow \infty} \left( \mathrm{e}^{\frac{A}{N}} \mathrm{e}^{\frac{B}{N}} \right)^N. \end{equation} We now use Eq. \eqref{eq:lie} to finalise the lemma. Consider the limit $m\rightarrow\infty$. On the l.h.s. of Eq. \eqref{eq:to power m with pot}, by the Lie product formula, we have that \begin{equation} \lim_{m\rightarrow\infty} \left( e^{-i \frac{T_0}{d} \frac{a}{m} \hat{V}_d} e^{-i \frac{T_0}{d} \frac{a}{m} \hat{H}_c} \right)^m \ket{\bar{\Psi}(k_0,\Delta)}= e^{-i \frac{T_0}{d}a (\hat H_c+\hat V_d)}\ket{\bar{\Psi}(k_0,\Delta)}. \end{equation} On the r.h.s. of Eq. \eqref{eq:to power m with pot}, we have \begin{equation} \lim_{m\rightarrow\infty} \left( \ket{\bar{\Psi} \left(k_0+ a, \Delta+a \right)} + \ket{\epsilon^{(m)}} \right)= \ket{\bar{\Psi} \left(k_0+ a, \Delta+a \right)} + \ket{\epsilon^{(\infty)}}, \end{equation} where $\ket{\epsilon^{(\infty)}}:=\lim_{m\rightarrow\infty} \ket{\epsilon^{(m)}}$, and \begin{equation} \ltwo{\ket{\epsilon^{(\infty)}}} \leq a \epsilon_T \sqrt{d}. \end{equation} \end{proof} At this point, we have already proven the continuity of the clock state for time translations that are finite, but small (i.e. small enough that the range $\mathcal{S}_d(k_0)$ remains the same). In order to generalize the statement to arbitrary translations we need to be able to shift the range itself, which is the goal of the following Lemma. \begin{lemma}[Shifting the range of the clock state with potential]\label{lem:Moving the clock through one angle step} If $d$ is even, and the mean of the clock state $k_0$ is an integer, or alternatively, if $d$ is $odd$ and $k_0$ is a half integer, then \begin{align} \sum_{k \in \mathcal{S}_d(k_0)} \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k)}\psi(k_0;k) \ket{\theta_k} &=\sum_{k \in \mathcal{S}_d(k_0-1)} \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k)}\psi(k_0;k) \ket{\theta_k} +\ket{\bar\epsilon_{step}}, \\ \text{where} \quad \|\ket{\bar\epsilon_{step}}\|_2:=\bar\epsilon_{step} &< \begin{cases} 2 A e^{-\frac{ \pi d}{4}} &\mbox{if } \sigma=\sqrt{d}\\ 2 A e^{-\frac{ \pi d^2}{4\sigma^2}} &\mbox{otherwise.} \end{cases} \end{align} \end{lemma} \begin{proof} We prove the statement for even $d$, the proof for odd $d$ is analogous. By definition \eqref{gaussianclock}, $\mathcal{S}_d(k_0)$ is a set of $d$ consecutive integers. Thus the only difference between $\mathcal{S}_d(k_0)$ and $\mathcal{S}_d(k_0-1)$ is the leftmost integer of $\mathcal{S}_d(k_0-1)$ and the rightmost integer of $\mathcal{S}_d(k_0)$, which differ by precisely $d$. By direct calculation, these correspond to the integers $k_0-d/2$ and $k_0+d/2$. These are the only two terms that do not cancel out in the statement of the Lemma, \begin{align} \|\ket{\bar\epsilon_{step}}\|_2 &\leq \ltwo{ \sum_{k \in \mathcal{S}_d(k_0-1)} \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k)}\psi(k_0;k_0) \ket{\theta_k} - \sum_{k \in \mathcal{S}_d(k_0)} \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k)}\psi(k_0;k) \ket{\theta_k} } \\ &= \ltwo{ \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k_0-d/2)}\psi(k_0;k_0-d/2) \ket{\theta_{k_0-d/2}} - \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k_0+d/2)}\psi(k_0;k_0+d/2) \ket{\theta_{k_0+d/2}}} \end{align} But $\ket{\theta_{k_0-d/2}} = \ket{\theta_{k_0-d/2+d}}= \ket{\theta_{k_0+d/2}}$, giving \begin{align} \|\ket{\bar\epsilon_{step}}\|_2 &\leq \abs{ \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k_0-d/2)}\psi(k_0;k_0-d/2) - \mathrm{e}^{-\mathrm{i}\Theta(\Delta;k_0+d/2)}\psi(k_0;k_0+d/2) }\\ &\leq \abs{ \psi(k_0;k_0-d/2)}+\abs{\psi(k_0;k_0+d/2) }\\ &=2 A \mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}}. \end{align} By direct substitution of $\psi(k_0;k_0-d/2)$ and $\psi(k_0;k_0+d/2)$ from \eqref{analyticposition}, we arrive at the Lemma statement. The lemma is analogous to that of Lemma, \ref{rangeshift}, but now with a trivial extension due to the inclusion of the potential in the dynamics. \end{proof} \begin{theorem}[Moving the clock through finite time with a potential]\label{movig through finite time} Let $k_0,\Delta,t \in {\mathbbm{R}}$. Then the effect of the Hamiltonian $\hat{H}_c + \hat{V}_d$ for the time $t$ on $\ket{\bar{\Psi}_\textup{nor}(k_0,\Delta)} \in\Lambda_{ V_0,\sigma,n_0}$ is approximated by \begin{align}\label{eq: main eq in control theorem} \mathrm{e}^{-\mathrm{i} t (\hat{H}_c + \hat{V}_d)} \ket{\bar{\Psi}_\textup{nor}(k_0,\Delta)} &= \ket{\bar{\Psi}_\textup{nor}(k_0+ \frac{d}{T_0} t, \Delta+ \frac{d}{T_0} t)} + \ket{\epsilon},\quad\quad\ltwo{\ket{\epsilon}} \leq \varepsilon_v(t,d), \end{align} where, in terms of order in $b$, $|t|$, $d$ and $\sigma$, \begin{equation} \varepsilon_v(t,d)= \begin{cases} |t| \frac{d}{T_0}\left(\mathcal{O}\left( \frac{d^{3/2}}{\bar\upsilon+1}\right)^{1/2}+\mathcal{O}\left(d+b\right)\right) \exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(1+\bar\upsilon\right)^2} d \right)+\mathcal{O}\left( \mathrm{e}^{-\frac{\pi}{2}d} \right) &\mbox{if } \sigma=\sqrt{d}\\[10pt] |t| \frac{d}{T_0}\!\left(\mathcal{O}\left( \frac{\sigma^3}{\bar\upsilon \sigma^2/d+1}\right)^{1/2}\!\!+\mathcal{O}\left(\frac{d^2}{\sigma^2}+b\right)\right) \exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(\frac{d}{\sigma^2}+\bar\upsilon\right)^2} \left(\frac{d}{\sigma}\right)^2 \right)+\mathcal{O}\left(|t|\frac{d^2}{\sigma^2}+1\right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}}+\mathcal{O}\left( \mathrm{e}^{-\frac{\pi}{2}\sigma^2} \right) &\mbox{otherwise.} \end{cases} \end{equation} and $\ket{\bar{\Psi}_\textup{nor}(\cdot,\cdot)}$ is defined in Def. \eqref{def:gauss clock steate with pot defintion}, $\alpha_0$ is defined in Def. \ref{def:stand alone def mathcal N} while $\bar\upsilon$ is defined in Def. \ref{def:decay rate params}.\\ More precisely, we have \begin{equation}\label{eq:ltwo epsilon nor control theorem} \varepsilon_v(t,d)=|t| \frac{d}{T_0} \epsilon_T+\left(|t| \frac{d}{T_0}+1\right)\bar\epsilon_{step}+\epsilon_\textup{nor}(t), \end{equation} where \begin{align} \epsilon_T &< \begin{cases} |\bar\epsilon_2|+2 A\left(\frac{2\pi}{1-\mathrm{e}^{-\pi}}+ \frac{\left(b+\frac{2\pi}{d}\right)}{1-\mathrm{e}^{-\pi d}} + \left( 2\pi+\pi d+ \frac{1}{d}\right)\right)\mathrm{e}^{-\frac{\pi}{4}d} &\mbox{if } \sigma=\sqrt{d}\\[10pt] |\bar\epsilon_2|+2 A\left(\frac{2\pi}{1-\mathrm{e}^{-\frac{\pi d}{\sigma^2}}}+ \frac{ \left(b+\frac{2\pi}{d}\right)}{1-\mathrm{e}^{-\pi \frac{d^2}{\sigma^2}}} + \left( 2\pi\frac{d}{\sigma^2}+\pi\frac{d^2}{\sigma^2}+ \frac{1}{d}\right)\right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}} &\mbox{otherwise,} \end{cases} \end{align} with $A=\mathcal{O}(\sigma^{-1/2})$ and is upper bounded by Eq. \eqref{eq:up low bounds for A normalize} and, if $\sigma=\sqrt{d}$ \begin{equation} |\bar\epsilon_2|< \begin{cases} (2\pi)^{5/4}d^{3/4} A \left(1+\frac{\pi^{2}}{8}\right)\sqrt{\frac{\mathrm{e}}{2}\frac{\alpha_0}{(\bar\upsilon+1)}}\exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(1+\bar\upsilon\right)^2} d \right)\quad &\text{if } \mathcal{N}\geq 8\text{ and } \bar\upsilon\geq 0\\ \frac{3^{7/4}}{\sqrt{2\pi}\mathrm{e}}\frac{A(8+\pi^{2})}{\alpha_0^3}\left(\frac{\kappa\sqrt{6\pi}}{\ln(3)}b+\sqrt{d}\right)^3 d^{-5/2}\quad &\text{otherwise,} \end{cases} \end{equation} while in general \begin{equation} |\bar\epsilon_2|< \begin{cases} (2\pi)^{5/4}\sigma^{3/2} A \left(1+\frac{\pi^{2}}{8}\right)\sqrt{\frac{\mathrm{e}}{2}\frac{\alpha_0}{(\bar\upsilon\sigma^2/d+1)}}\exp\left(-\frac{\pi}{4}\frac{\alpha_0^2}{\left(\frac{d}{\sigma^2}+\bar\upsilon\right)^2} \left(\frac{d}{\sigma}\right)^2 \right)\quad &\text{if } \mathcal{N}\geq 8\text{ and } \bar\upsilon\geq 0\\ \frac{3^{7/4}}{\sqrt{2\pi}\mathrm{e}}\frac{A(8+\pi^{2})}{\alpha_0^3}\left(\frac{\kappa\sqrt{6\pi}}{\ln(3)}b+\frac{d}{\sigma}\right)^3 \left( \frac{\sigma}{d^3} \right)\quad &\text{otherwise} \end{cases} \end{equation} where $b$, $\mathcal{N}$, and $\alpha_0$ are defined in Def. \ref{def:stand alone def mathcal N} while $\bar\upsilon$ is defined in Def. \ref{def:decay rate params}. \begin{align} \epsilon_\textup{nor}(t)&=\sqrt{\frac{\sum_{k\in \mathcal{S}_d(k_0)} \mathrm{e}^{-\frac{2\pi}{\sigma^2}(k-k_0-\frac{d}{T_0}t)^2}}{\sum_{l\in \mathcal{S}_d(k_0)} \mathrm{e}^{-\frac{2\pi}{\sigma^2}(k-k_0)^2}}}-1\leq \begin{cases}\displaystyle 8\sqrt{\frac{2}{d}}\frac{ \mathrm{e}^{-\frac{\pi d}{2}}}{1 - \mathrm{e}^{-\pi d}}\quad &\forall\, t\in{\mathbbm{R}} \;\;\mbox{ if } \sigma=\sqrt{d}\\[10pt]\displaystyle \frac{4\sqrt{2}}{\sigma}\left(\frac{ \mathrm{e}^{-\frac{\pi d^2}{2\sigma^2}}}{1 - \mathrm{e}^{-\frac{2\pi d}{\sigma^2}}}+\frac{ \mathrm{e}^{-\frac{\pi \sigma^2}{2}}}{1 - \mathrm{e}^{-\pi \sigma^2}}\right) \quad &\forall\, t\in{\mathbbm{R}} \;\;\mbox{ otherwise} \end{cases}\\ \bar\epsilon_{step} &< \begin{cases} 2 A e^{-\frac{ \pi d}{4}} &\mbox{if } \sigma=\sqrt{d}\\ 2 A e^{-\frac{ \pi d^2}{4\sigma^2}} &\mbox{otherwise.} \end{cases} \end{align} \end{theorem} \emph{Intuition.} The discrete clock mimics the idealised clock, with an error that grows linearly with time, and scales better than any inverse polynomial w.r.t. the dimension of the clock. The optimal decay is when the state is symmetric, i.e. when $\sigma=\sqrt{d}$. This gives exponentially small error in $d$, the clock dimension. The rate of decay i.e. the coefficient in the exponential, is determined by to competing factors, $\bar\upsilon$ and $\alpha_0$. The latter being a measure of how clock the initial clocks states mean energy is to either the maximum or minimum value, while the former is a measure of how steep the potential function $\hat V_0$ is. The larges decay parameter, $\pi/4$ is only achieved asymptotically when both the mean energy of the initial clock state tends to the middle of the spectrum in the large $d$ limit and $\bar\upsilon$ tends to zero in the said limit. The latter limit, is true for arbitrarily steep potentials. \begin{remark} Observe by Eq. \eqref{eq:ltwo epsilon nor control theorem} that $\lim_{t\rightarrow 0}\varepsilon_v(0,d)\neq 0$, yet from Eq. \eqref{eq: main eq in control theorem} clearly $\|\ket{\epsilon}\|_2=0$ in this limit. One can trivially modify the proof to find upper bounds for $\varepsilon_v(t,d)$ which are of order $t$ in the $t\rightarrow 0$ limit if required. \end{remark} \begin{proof} For, $t\geq 0$, directly apply the previous two Lemmas (\ref{lem:Moving the clock through one angle step}, \ref{Moving the clock through finite time, within unit angle}) in alternation, first to move $k_0$ from one integer to the next, then to switch from ${\lfloor k_0\rfloor}$ to ${\lfloor k_0+1\rfloor}$ if $d$ is even, and ${\lfloor k_0+1/2\rfloor}$ to ${\lfloor k_0+1/2+1\rfloor}$ if $d$ is odd, and finally arriving at \begin{align} \mathrm{e}^{-it (\hat{H}_c + \hat{V}_d)} \ket{\bar{\Psi}(k_0,\Delta)} &= \ket{\bar{\Psi}(k_0+ \frac{d}{T_0} t, \Delta+ \frac{d}{T_0} t)} + \ket{\epsilon},\quad\quad\ltwo{\ket{\epsilon}} \leq t \frac{d}{T_0} \epsilon_T+\left(t \frac{d}{T_0}+1\right)\bar\epsilon_{step}. \end{align} To conclude Eq. \eqref{eq: main eq in control theorem} for $t>0$, one now has to normalize the states using bounds from Section \ref{Re-normalizing the clock state} and then use Lemma \ref{unitaryerroraddition} to upper bound the total error. For $t<0$, simply evaluate Eq. \eqref{eq: main eq in control theorem} for a time $|t|$ followed by multiplying both sides of the equation by $\exp(\mathrm{i} t(\hat H_c+\hat V_d))$, mapping $k_0\rightarrow k_0-t\;d/T_0$, $\Delta\rightarrow \Delta-t\;d/T_0$ and noting the unitary invariance of the $l_2$ norm of $\ket{\epsilon}$. \end{proof} \begin{corollary} \label{movig through finite time coro} \emph{(Evolution in the time basis with potential).} Let $k_0,t \in {\mathbbm{R}}$ and define a \textup{discreate wave function} in analogy with continuous wave functions, \begin{equation} \bar{\psi}(x,t):=\bra{\theta_x}\mathrm{e}^{-\mathrm{i} t (\hat{H}_c + \hat{V}_d)} \ket{\bar{\Psi}_\textup{nor}(k_0)}, \end{equation} for $x\in\mathcal{S}_d(k_0+t \,d/T_0) , $t\in{\mathbbm{R}}$. Then the time evolution of $\bar{\psi}$ is approximated by \begin{align}\label{eq:bar psi evolution} \bar{\psi}(x,t) &= \mathrm{e}^{-\mathrm{i} \int_{x-t\,d/T_0}^x dy V_d(y)}\,\bar{\psi}(x-\frac{d}{T_0} t,0) + \bar\varepsilon_v(t,d),\quad \quad |\bar\varepsilon_v(t,d)|\leq \varepsilon_v(t,d), \end{align} where $\bar\varepsilon_v\in{\mathbbm{C}}$ and $\varepsilon_v(t,d)$ is given by Eq. \eqref{eq:ltwo epsilon nor control theorem}. \end{corollary} \emph{(Intuition.)} This is the analogous statement to that of Eq. \eqref{eq:psi idealised control}, i.e. that up to an error $\varepsilon_v(t,d)$, the time evolution overlap of the Gaussian state in the time basis, by the clock Hamiltonian plus a potential term, is simply the time translated overlap multiplied by an exponentiated phase which integrates over the potential. \begin{proof} This is a direct consequence of Theorem \ref{movig through finite time}. Note that by definition, \begin{equation} \bar{\psi}(x,0)=\psi_\textup{nor}(k_0;x). \end{equation} Thus by taking the inner product of Eq. \eqref{eq: main eq in control theorem} with $\ket{\theta_k}$ followed by noting $\psi(k_0;k+y)=\psi(k_0-y;k)$ and $|\!\braket{\theta_k|\epsilon}\!|\leq\varepsilon_v(t,d)$, we achieve Eq. \eqref{eq:bar psi evolution}. \end{proof} \begin{remark} Note that when $t$ corresponds to one period, namely $t=T_0=2\pi/\omega$, it follows from the definitions that the state $\ket{\bar\Psi(k_0+\frac{d}{T_0} t,\Delta+\frac{d}{T_0} t)}$ in Theorem \ref{movig through finite time} has returned to its initial state, up to a global phase, namely \begin{equation} \mathrm{e}^{-\mathrm{i} T_0 (\hat{H}_c + \hat{V}_d)} \ket{\bar{\Psi}(k_0,\Delta)} = \mathrm{e}^{-\mathrm{i}\Omega}\ket{ \bar{\Psi} \left( k_0,\Delta \right)} + \ket{\epsilon}, \end{equation} where the $l$-two norm of $\ket{\epsilon}$ is as before. Similarly, for the discrete wafe function of Eq. \eqref{eq:bar psi evolution}, \begin{align} \bar{\psi}(x,T_0) &= \mathrm{e}^{-\mathrm{i} \Omega}\,\bar{\psi}(x,0) + \bar\varepsilon_v(T_0,d). \end{align} \end{remark} \subsection{Examples of Potential functions}\label{sec:Examples of Potential functions} \subsubsection{The cosine potential}\label{sec:The cosine potential} In this section, we calculate abound for $b$ explicitly for the cosine potential. Let \begin{equation}\label{def: cosine pot} V_0(x)=A_c \cos^{2n}\left( \frac{x}{2} \right), \end{equation} where $n\in{\mathbbm{N}}^+$ is a parameter which determines how steep the potential is and $A_c$ is a to be determined normalisation constant. In order to proceed, we use the identity \begin{equation} \cos^{2n}(x/2)=\frac{1}{2^{2n}} (\mathrm{e}^{ix/2}+\mathrm{e}^{-ix/2})^{2n}= \frac{1}{2^{2n}} \sum_{k=0}^{2n} \binom{2n}{k} \mathrm{e}^{ix k/2}\mathrm{e}^{-ix(2n-k)/2}=\frac{1}{2^{2n}} \sum_{k=0}^{2n} \binom{2n}{k} \mathrm{e}^{ix (k-n)}. \end{equation} Therefore, from Eq. \eqref{eq:theta normalisation}, it follows \begin{equation} \Omega=A_c \int_0^{2\pi}dy \cos^{2n}\left( \frac{y}{2} \right)=\frac{A_c}{2^{2n}} \sum_{k=0}^{2n} \binom{2n}{k}\int_0^{2\pi} \mathrm{e}^{ix (k-n)}=\frac{A_c}{2^{2n}} \sum_{k=0}^{2n} \binom{2n}{k}2\pi \delta_{k,n}= \frac{A_c}{2^{2n}} \binom{2n}{n} 2\pi, \end{equation} giving \begin{equation}\label{def: A c eq def for cos pot} A_c=\Omega \frac{2^{2 n}}{2\pi \binom{2n}{n}}=\Omega\frac{2 ^{2 n}}{2\pi}\frac{(n!)^2}{(2n)!}. \end{equation} Thus \begin{align} \left| V_0^{(q)}(x)\right|&=\left| \frac{\Omega}{2\pi}\frac{(n!)^2}{(2n)!}\sum_{k=0}^{2n}\binom{2n}{k}\left(\mathrm{i}(k-n)\right)^q\mathrm{e}^{i x(k-n)} \right|\leq \frac{|\Omega|}{2\pi}\frac{(n!)^2}{(2n)!}\sum_{k=0}^{2n}\binom{2n}{k}|k-n|^q\leq \frac{|\Omega|}{2\pi}\frac{(n!)^2}{(2n)!}\sum_{k=0}^{2n}\binom{2n}{k}n^q\label{eq:V devs up bound, inter line} \\ &=\frac{|\Omega|}{2\pi}\frac{(n!)^2}{(2n)!} 4^n n^q\leq \frac{\mathrm{e}^2}{(2\pi)^{3/2}}\frac{|\Omega|}{\sqrt{2}} n^{q+3/2},\quad q\in {\mathbbm{N}}^0, \end{align} Where we have used Stirling's approximation to bound the factorials. Note that one can improve upon this bound by explicity calculating in closed form the penultimate equality in line \eqref{eq:V devs up bound, inter line}. Thus we can upper bound the r.h.s. of Eq. \eqref{eq: b_1 explicit}, to achieve \begin{equation} \sup_{k\in {\mathbbm{N}}^+} \left(2\max_{x\in[0,2\pi]} \left| V_0^{(k-1)}(x)\right|\,\right)^{1/k}\leq \sup_{k\in {\mathbbm{N}}^+} \left(\frac{2\,\mathrm{e}^2}{(2\pi)^{3/2}}\frac{|\Omega|}{\sqrt{2}} n^{k+1/2} \right)^{1/k}. \end{equation} Thus using the formula \begin{equation} \frac{d}{dk} c^{1/k}=\frac{c^{1/k}}{k^2}\ln(1/c) \end{equation} and taking into account Eq. \eqref{eq: b_1 explicit}, we can conclude a value for $b$, namely \begin{equation}\label{eq:b final for cos example} b= \begin{cases} n \left(\frac{2\,\mathrm{e}^2}{(2\pi)^{3/2}}\frac{|\Omega|}{\sqrt{2}} n^{1/2} \right)^{1/k} \bigg{|}_{k=1} =\frac{2\,\mathrm{e}^2}{(2\pi)^{3/2}}\frac{|\Omega|}{\sqrt{2}}\,n\sqrt{n} &\mbox{if } |\Omega| \sqrt{n} \geq \frac{(2\pi)^{3/2}}{\sqrt{2}\mathrm{e}^2}\approx 1.507\\ n \lim_{k\rightarrow\infty} \left(\frac{2\,\mathrm{e}^2}{(2\pi)^{3/2}}\frac{|\Omega|}{\sqrt{2}} n^{1/2} \right)^{1/k}=n &\mbox{if } |\Omega| \sqrt{n} \leq \frac{(2\pi)^{3/2}}{\sqrt{2}\mathrm{e}^2}. \end{cases} \end{equation} \section{Clocks as Quantum control}\label{sec:Clocks as Quantum control} This section will be concerned with the implementation of unitaries on a finite dimensional quantum system with Hilbert space $\mathcal{H}_s$ via a control system. It will be in this section that the importance of Theorem \ref{movig through finite time} for controlling a quantum system will become apparent. To start with, we will consider implementing energy preserving unitaries. The first Section will consider the case that the implementation is via a time dependent Hamiltonian (Section\ref{sec:Implementing Energy preserving unitaries via a time dependent Hamiltonian}) followed by showing that the idealised clock can, via a time independent Hamiltonian, perfectly implement controlled energy preserving unitaries (Section \ref{sec:Automation via the idealised clock}). Finally, these previous two sections will serve as an introduction and provide context for the first important section about clocks as quantum control, namely how well (as a function of the clock dimension) can energy preserving unitaries be implemented via finite dimensional Gaussian clocks (Section \ref{sec:Implementing Energy preserving unitaries with the finite}). To finalize this discussion, we will then discuss how to perform non energy preserving unitaries using the finite dimensional clock in Section \ref{sec:non energy preserving unitaries}. This will require a source of energy and potentially coherence too. The purpose of the clock is a time peace, and to extract energy and or coherence from it will cause large degradation. As such we will do this by appending a battery to the setup. To conclude, in Section \ref{Clock Fidelity}, we will bound how the implementation of the unitary on the system via the finite dimension clock, disturbs the dynamics of the clock. \subsection{Implementing Energy preserving unitaries via a time dependent Hamiltonian}\label{sec:Implementing Energy preserving unitaries via a time dependent Hamiltonian} Consider a system of dimension $d_s$, that begins in a state $\rho_s\in\mathcal{S}(\mathcal{H}_s)$, and upon which one wishes to perform the unitary $U_s$ over a time interval $[t_i,t_f]$. Energy-preserving means that $[U_s,\hat H_s]=0$, where $H_s$ is the time-independent natural Hamiltonian of the system with point spectrum. The first step in automation is to convert the unitary into a time-dependent interaction, via \begin{equation}\label{unitaryasinteraction} U_s = e^{-\mathrm{i} \hat H^{int}_s}, \end{equation} where $\hat H^{int}_s$ also commutes with $\hat H_s$. The unitary can thus be implemented by the addition of $\hat H^{int}_s$ as an time-dependent interaction \begin{equation}\label{eq:system Ham t depen} \hat H = \hat H_s + \hat H^{int}_s \cdot g(t). \end{equation} If $g\in L({\mathbbm{R}}: {\mathbbm{R}}_{\geq 0})$ is a normalized pulse, i.e. \begin{align}\label{eq:g(t) pulse normalisation} \int_{t_i}^{t_f}dt g(t) &= 1, \end{align} with support interval $[t_i,t_f]$, then it is easily verified that the unitary $U$ will be implemented between $t_i$ and $t_f$, i.e. given the initial system state $\rho_s\in\mathcal{S}(\mathcal{H}_s)$, the state at the time $t$ is \begin{equation}\label{eq:rho t scho sol commute t depen} \rho_s(t) = \mathrm{e}^{-\mathrm{i} t\hat H_s} \mathrm{e}^{-\mathrm{i}\hat H^{int}_s\int_{t_i}^t dx\, g(x)} \rho_s\; \mathrm{e}^{+\mathrm{i} \hat H^{int}_s \int_{t_i}^t dx\, g(x)} \mathrm{e}^{+\mathrm{i} t\hat H_s}\\ = \mathrm{e}^{-\mathrm{i} t\hat H_s-\mathrm{i}\hat H^{int}_s\int_{t_i}^t dx\, g(x)} \rho_s\; \mathrm{e}^{+\mathrm{i} t \hat H_s+\mathrm{i} \hat H^{int}_s \int_{t_i}^t dx\, g(x)}. \end{equation} To gain a deeper understanding, we first note that since $\hat H_s$ and $\hat H_s^{int}$ commute, there must exist a mutually orthonormal basis, denoted by $\{\ket{\phi_j}\}_{j=1}^{d_s}$ such that \begin{align} \hat H_s &= \sum_{j=1}^{d_s} E_j \ketbra{\phi_j}{\phi_j}_s, \\ \hat H^{int}_s &= \sum_{j=1}^{d_s} \Omega_j \ketbra{\phi_j}{\phi_j}_s, \end{align} where, due to Eq. \eqref{unitaryasinteraction}, without loss of generality, we can confine $\Omega_j\in[-\pi,\pi)$, for $j=1,2,3,\ldots,d_s$. By writing the evolution of the free Hamiltonian in the $\{\ket{\phi_j}\}_{j=1}^{d_s}$ basis \begin{equation}\label{eq:rho s in mutual alphog basis} \rho(t)=\mathrm{e}^{-\mathrm{i} t \hat H_s}\rho_s \,\mathrm{e}^{\mathrm{i} t \hat H_s}=\sum_{m,n=1}^{d_s}\rho_{m,n}(t) \ketbra{\phi_m}{\phi_n}, \end{equation} Eq. \eqref{eq:rho t scho sol commute t depen} can be written in the form \begin{equation} \label{eq:rho as a function of t} \rho_s(t)=\sum_{m,n=1}^{d_s}\rho_{m,n}(t)\,\mathrm{e}^{-\mathrm{i} (\Omega_m-\Omega_n) \int_{t_i}^t dt\, g(t)} \ketbra{\phi_m}{\phi_n}. \end{equation} \subsection{Implementing Energy preserving unitaries with the idealised clock}\label{sec:Automation via the idealised clock} To remove the explicit time-dependence of Eq. \eqref{eq:system Ham t depen}, we insert a clock, and replace the background time parameter $t$ by the `time' degree of the clock. For the idealised clock, this corresponds to a Hamiltonian on $\mathcal{H}=\mathcal{H}_s\otimes\mathcal{H}_c$, given by \begin{equation}\label{eq:idealised s c ham} \hat H_{total}^{id} = \hat H_s \otimes {\mathbbm{1}}_c + {\mathbbm{1}}_s \otimes \hat{p}_c + \hat H^{int}_s \otimes g(\hat{x}_c), \end{equation} where $\hat x_c$ and $\hat p_c$ are the canonically conjugate position and momentum operators of a free particle in one dimension detailed in Section \ref{idealizedclock}. To verify that this Hamiltonian can indeed implement the unitary, we let the initial state of the system and clock be in a product form, \begin{equation} \rho_{sc}^{id}=\rho_s\otimes\ketbra{\Psi}{\Psi}_c. \end{equation} Writing the state $\rho_{sc}^{id}$ at a later time in the $\{\ket{\phi_j}\}_{j=1}^{d_s}$, we find \begin{equation}\label{eq:rho ideal s c} \rho_{sc}^{id}(t)=\mathrm{e}^{-\mathrm{i} t \hat H_{total}^{id}}\rho_{sc}^{id}\, \mathrm{e}^{\mathrm{i} t \hat H_{total}^{id}}= \sum_{m,n=1}^{d_s} \rho_{m,n}(t)\otimes\ketbra{\Phi_m(t)}{\Phi_n(t)}_c, \end{equation} with \begin{equation} \ket{\Phi_n(t)}_c:=\mathrm{e}^{-\mathrm{i} t( \hat{p}_c + \Omega_n g(\hat{x}_c))}\ket{\Phi}_c. \end{equation} \begin{lemma}\label{lem:doing unitary id} Let $\psi(x,0)\in D_0$ be the normalised wave-function with support on the interval $x\in[x_{\psi l},x_{\psi r}]$ associated with the state $\ket{\Phi}_c$. Let $g(x)$ have support on $x\in[x_{g l},x_{g r}]$ and denote $\rho_s^{id}(t)=\textup{tr}_c[\rho_{sc}^{id}(t)]$ as the partial trace over $\mathcal{H}_c$, with $\rho_{sc}^{id}(t)$ given by Eq. \eqref{eq:rho ideal s c}. Furthermore, assume that initally the wave packet is on the left of the potential function, namely $x_{\psi r}\leq x_{g l}$. Then the unitary is implemented perfectly in the time interval $x_{g l}-x_{\psi r} \leq t\leq x_{g r}-x_{\psi l}$, more precisely \begin{equation} \rho_s^{id}(t)= \begin{cases} \sum_{m,n=1}^{d_s}\rho_{m,n}(t) \ketbra{\phi_m}{\phi_n} &\mbox{if }\; t\leq x_{g l}-x_{\psi r}\\ \sum_{m,n=1}^{d_s}\rho_{m,n}(t) \ketbra{\phi_m}{\phi_n} \,\mathrm{e}^{-\mathrm{i} (\Omega_m-\Omega_n)} &\mbox{if } \, t\geq x_{g r}-x_{\psi l} \end{cases} \end{equation} \end{lemma} \emph{Intuition.} If one sets $t_i=x_{g l}-x_{\psi r}$, $t_f=x_{g r}-x_{\psi l}$, then the clock can control perfectly the system and mimics the time dependent Hamiltonian of Eq. \eqref{eq:system Ham t depen}. The Lemma is a consequence of the property noted in Eq. \eqref{eq:psi idealised control}, and hence the name given to dynamics of this form. \begin{proof} Taking the partial trace over the idealized clock in Eq. \eqref{eq:rho ideal s c}, we achieve \begin{equation}\label{eq:rho s id 1} \rho_{s}(t)= \sum_{m,n=1}^{d_s} \rho_{m,n}(t)\braket{\Phi_n(t)|\Psi_m(t)}_c, \end{equation} where using Eq. \eqref{eq:psi idealised control}, we find \begin{align} \braket{\Phi_n(t)|\Psi_m(t)}_c &=\int_{-\infty}^\infty dx\braket{\Phi_n(t)|x}_{\!\!c}\!\! \braket{x|\Psi_m(t)} =\int_{x_{\psi l}}^{x_{\psi r}} dx |\psi(x,0)|^2 \,\mathrm{e}^{-\mathrm{i} (\Omega_m-\Omega_n)\int_x^{x+t}g(x')dx'}.\label{eq:inner product n m} \end{align} Note that \begin{equation} \int_x^{x+t} g(x')dx'= \begin{cases}\label{eq:cases conditioons in idealisec control lemma} 0 &\mbox{if } x+t\leq x_{gl}\\ 1 &\mbox{if } x\leq x_{vl} \text{ and } x+t\geq x_{vr} \end{cases} \end{equation} Thus in order for $\exp({-\mathrm{i} (\Omega_m-\Omega_n)\int_x^{x+t}g(x')dx'})$ to be a constant factor in the integral of Eq. \eqref{eq:inner product n m}, we need the conditions on $x$ in Eq. \eqref{eq:cases conditioons in idealisec control lemma} to hold for all $x\in[x_{\psi l},x_{\psi r}]$. Noting the normalization of the wave function, we conclude the Lemma. \end{proof} One might also wonder whether or not the state of the clock is degraded in any way due to the action of performing the unitary. The wave-function only changes by a global phase, and thus the answer to this question is no. This means that one can implement any number of energy preserving unitaries $\{U_n\}_n$ on the system in any sequence of time intervals. This of course, is under the physically questionable assumption of the validity of the idealized clock. \subsection{Implementing Energy preserving unitaries with the finite dimensional Gaussian clock}\label{sec:Implementing Energy preserving unitaries with the finite} In this section we will see how well the finite dimensional Gaussian clock can mimic the behavior of the idealized clock studied in Section \ref{sec:Implementing Energy preserving unitaries via a time dependent Hamiltonian}. We will find that Theorem \ref{movig through finite time} implies that the clock can perform on the system any sequence of unitaries with an additive error which grows linearly in time and decays faster than any polynomial in clock dimension (the exact error will depend on a number of things, for example which Gaussian initial clock state).\\ In analogy with the idealized clock-system Hamiltonian, Eq. \eqref{eq:idealised s c ham}, we define \begin{equation}\label{eq:finite s c ham} \hat H_{sc} = \hat H_s \otimes {\mathbbm{1}}_c + {\mathbbm{1}}_s \otimes \hat{H}_c + \hat H^{int}_s \otimes \hat{V}_d, \end{equation} with $\Omega=1$, where $\hat H_c$ is defined in Eq. \eqref{finiteHamiltonian} and $\Omega$, $\hat{V}_d$ are defined in Eq. \eqref{eq: Theta def}. Furthermore, we denote the system and clock evolution under $\hat H_{sc}$ as \begin{equation}\label{eq:rho finite s c} \rho_{sc}'(t)=\mathrm{e}^{-\mathrm{i} t \hat H_{sc}}\rho_s\otimes\ketbra{\Psi_\textup{nor}(k_0)}{\Psi_\textup{nor}(k_0)}_c\, \mathrm{e}^{\mathrm{i} t \hat H_{sc}}. \end{equation} Recall that the clock has period $T_0$ and aims to implement the potential $V_0$ while the time dependent Hamiltonian aims to implement the potential $g$ over an interval $t\in[t_i,t_f]$. In order to make a fair comparison we will set \begin{equation}\label{eq:g in terms of V_0} g(x)=\frac{2\pi}{T_0} V_0(x\; 2\pi/T_0),\quad [t_i,t_f]=[0,T_0], \end{equation} which taking into account Eq. \eqref{eq:theta normalisation}, implies that Eq. \eqref{eq:g(t) pulse normalisation} is satisfied. Note that while Eq. \eqref{eq:g in terms of V_0} is a natural definition of $g$, it differs from the definition used in Section \ref{sec:Consequences of Quasi-Autonomous control}. We will remedy this in Corollary \ref{lem:new g}. \begin{lemma}[Implicit form]\label{lem:trace dist bound t-Ham Vs d dim clock} Let $\rho_s'(t):=\textup{tr}_c[\rho_{sc}'(t)]$, then for all $\rho_s(0)\in\mathcal{S}(\mathcal{H}_s)$, and $t\in[0,T_0]$, the trace distance between $\rho_s'(t)$ and the idealised case of controlling the system with a time dependent Hamiltonian, $\rho_s(t)$ (see Eq. \eqref{eq:rho t scho sol commute t depen}; $g$ given by Eq. \eqref{eq:g in terms of V_0}), is bounded by \begin{equation}\label{eq:trace dist rho rho finite and time} \|\rho_s(t)-\rho_s'(t)\|_1\leq \sqrt{d_s \textup{tr}[\rho_s^2(0)]}\left(2\varepsilon_v(t,d)+\varepsilon_v^2(t,d)+\epsilon_V(t,d) \right), \end{equation} where $\varepsilon_v(t,d)$ is evaluated for $\Omega=\pi$ and defined by Eq. \eqref{eq:ltwo epsilon nor control theorem} while\footnote{In the following summation, and for now on, we use the convention $\sum_{k=a}^b f(k)= f(a)+f(a+1)+\ldots+f(b)$, where $b-a\in{\mathbbm{N}}^0$} \begin{equation}\label{eq:def epsilon k} \epsilon_V(t,d):=\max_{\bar\kappa\in[0,1]} \sum_{k=k_0-d/2+\bar\kappa}^{k_0+d/2+\bar\kappa}\left( \epsilon_k^2+\epsilon_k\right)|\psi_\textup{nor}(k_0;k)|^2,\quad\quad \epsilon_k=2\pi\left| \int_{0}^{t\, 2\pi/T_0} dx\big( V_0(x)-V_0(x+2\pi k/d)\big)\right|. \end{equation} \end{lemma} \emph{Intuition}. The point of this Lemma is to make it manifest that the error invoked by using the finite dimensional Gaussian clock to control the quantum system on $\mathcal{H}_s$ (specifically, in this case, to perform an energy conserving unitary during a specific moment in time) is small. The two error terms on the r.h.s. of Eq. \eqref{eq:trace dist rho rho finite and time} involving $\varepsilon_v(t,d)$ decay fast with $d$ (see Eq. \eqref{eq:ltwo epsilon nor control theorem}) and only grow linearly in time. Note that $\epsilon_V(j\, T_0,d)=0$ for all $d\in {\mathbbm{N}}^+$, $j\in{\mathbbm{N}}^0$. More generally, the error term $\epsilon_V$, will be small for a clock initially centered at $k_0$, at time $t$ when $V_0(2\pi(t/T_0+k_0/d))\approx V_0(2\pi t/T_0)$, i.e. is relatively flat. Roughly speaking, this will be true for times $t$ before and after the unitary has been implemented as the example in Fig. \ref{fig:epsilon V exmaple plot} and the following Lemma will make explicit. Lemma \ref{lem:trace dist bound t-Ham Vs d dim clock} still holds if we do not assume the special choice of $g$ in Eq. \eqref{eq:g in terms of V_0}. For general $g$, we simply replace $\epsilon_k$ in Eq. \eqref{eq:def epsilon k} with Eq. \eqref{eq:epsilon k lemma}. As regards to the system size dependency, it varies between $\sqrt{d_s}$ growth for pure initial system states, to being $d_s$ independent in the case of maximally mixed initial system states. The latter is what one should expect, since clearly an initially maximally mixed state is invariant under the unitary dynamics and as such the error associated with implementing a unitary should not depend on its dimension. \begin{proof} Proceeding similarly to the derivation of Eq. \eqref{eq:rho s id 1} in the proof of Lemma \ref{lem:doing unitary id}, we achieve \begin{equation} \rho_{s}(t)= \sum_{m,n=1}^{d_s} \rho_{m,n}(t)\braket{\bar{\Phi}_n(t)|\bar{\Phi}_m(t)}, \end{equation} with \begin{equation}\label{eq:Psi bar def eq} \ket{\bar{\Phi}_l(t)}:=\mathrm{e}^{-\mathrm{i} t( \hat{H}_c + \Omega_l \hat{V}_d)}\ket{\Psi_\textup{nor}(k_0)}, \quad l=n,m. \end{equation} Thus noting that the Frobenius norm, $\|\cdot\|_F$, is an upper bound to the trace distance $\|\cdot\|_1/\sqrt{d_s}$, for a $d_s$ dimensional system, followed by using Eq. \eqref{eq:rho as a function of t}, \begin{align} \|\rho_s(t)-\rho_s'(t)\|_1& \leq \sqrt{d_s} \abs{\abs{\rho(t)-\rho'(t)}}_F=\sqrt{d_s}\;\sqrt{\sum_{n,m}|\rho_{nm}(t)|^2\left| e^{-\mathrm{i}(\Omega_m-\Omega_n)\int_{t_i}^{t_f}g(x)dx}-\braket{\bar{\Phi}_n(t)|\bar{\Phi}_m(t)}\right|^2}\\ & \leq \sqrt{d_s}\;\sqrt{\sum_{n,m}|\rho_{nm}(t)|^2 \max_{q,r}\left\{\left| e^{-\mathrm{i}(\Omega_m-\Omega_n)\int_{t_i}^{t_f}g(x)dx)}-\braket{\bar{\Phi}_r(t)|\bar{\Phi}_q(t)} \right|^2\right\}} \\ & \leq \sqrt{d_s\textup{tr}\left(\rho^2_s(0)\right)} \max_{q,r}\left\{\left| e^{-\mathrm{i}(\Omega_m-\Omega_n)\int_{t_i}^{t_f}g(x)dx}-\braket{\bar{\Phi}_r(t)|\bar{\Phi}_q(t)} \right|\right\}.\label{eq: rho - rho 1} \end{align} Applying Theorem \ref{movig through finite time}, \begin{equation} \braket{\bar{\Phi}_n(t)|\bar{\Phi}_m(t)}=\sum_{k\in\mathcal{S}(k_0+dt/T_0)} \mathrm{e}^{-\mathrm{i} (\Omega_n-\Omega_m)\int_{k-td/T_0}^k dy V_d(y)} \left| \psi_\textup{nor}(k_0;k-td/T_0)\right|^2+\tilde\epsilon_{n,m}, \end{equation} where \begin{align}\label{eq:tilde epsilon up bound} \tilde\epsilon_{n,m}&=\braket{\bar\Psi_\textup{nor}(k_0+td/T_0)|\epsilon}+\braket{\epsilon|\bar\Psi_\textup{nor}(k_0+td/T_0)}+ \braket{\epsilon|\epsilon},\\ |\tilde\epsilon_{n,m}|&\leq 2\varepsilon_v(t,d)+\varepsilon_v^2(t,d),\label{eq:up bund n m independ ep tilde} \end{align} with $\varepsilon_v(t,d)$ is given by Eq. \eqref{eq:ltwo epsilon nor control theorem}. Note that the r.h.s of Eq. \eqref{eq:tilde epsilon up bound}, does depend on indices $n$ and $m$. The r.h.s. of Eq. \eqref{eq:up bund n m independ ep tilde} on the other hand, is independent of $n$ and $m$.\\ Thus applying Eq. \eqref{eq:up bund n m independ ep tilde} to Eq. \eqref{eq: rho - rho 1}, we achieve \begin{align} \frac{\|\rho_s(t)-\rho_s'(t)\|_1}{ \sqrt{d_s\textup{tr}[\rho^2_s(0)]}} \leq & \max_{q,r}\left\{\left| 1-\sum_{k\in\mathcal{S}(k_0+dt/T_0)} \mathrm{e}^{\mathrm{i} (\Omega_q-\Omega_r)\left(\int_{t_i}^{t_f}g(x)dx-\int_{k-td/T_0}^k dy V_d(y)\right)} \left| \psi_\textup{nor}(k_0;k-td/T_0)\right|^2 \right|\right\}\\ &+2\varepsilon_v(t,d)+\varepsilon_v^2(t,d)\\ \leq & \max_{q\geq r}\left\{\left| 1-\sum_{k\in\mathcal{S}(k_0+dt/T_0)} \mathrm{e}^{\pm\mathrm{i} |\Omega_q-\Omega_r|\left(\int_{t_i}^{t_f}g(x)dx-\int_{k-td/T_0}^k dy V_d(y)\right)} \left| \psi_\textup{nor}(k_0;k-td/T_0)\right|^2 \right|\right\}\\ &+2\ \varepsilon_v(t,d)+\varepsilon_v^2(t,d)\\ =&\max_{q,r}\left\{ \rule{0cm}{1cm}\right.\sqrt{\left(1-\sum_{k\in\mathcal{S}(k_0+dt/T_0)} \cos(\pm g_k^{q,r})\left| \psi_\textup{nor}(k_0;k-td/T_0)\right|^2 \right)^2+}\\ &\,\,\quad\quad\,\,\,\,\,\,\,\,\overline{\left(\sum_{k\in\mathcal{S}(k_0+dt/T_0)}\sin(\pm g_k^{q,r})\left| \psi_\textup{nor}(k_0;k-td/T_0)\right|^2\right)^2}\left. \rule{0cm}{1cm}\right\}+2\varepsilon_v(t,d)+\varepsilon_v^2(t,d), \end{align} where \begin{equation}\label{eq:gk def in lemm} g_k^{q,r}:= |\Omega_q-\Omega_r|\left(\int_{t_i}^{t_f}g(x)dx- \int_{k-td/T_0}^k V_d(x)dx\right)\leq 2\pi\left|\int_{t_i}^{t_f}g(x)dx- \int_{k-td/T_0}^k V_d(x)dx\right|=:g_k. \end{equation} Thus using the identities $(|a|+|b|)^2\geq |a|^2+|b|^2$, $\sin(a)\leq |a|$, $-\cos(a)\leq a^2-1$, for $a,b\in{\mathbbm{R}}$, to achieve \begin{align} \frac{\|\rho_s(t)-\rho_s'(t)\|_1}{ \sqrt{d_s\textup{tr}[\rho^2_s(0)]}} \leq &\,2\varepsilon_v(t,d)+\varepsilon_v^2(t,d)+\sum_{k\in\mathcal{S}(k_0+dt/T_0)} (g_k^2+g_k)|\psi(k_0;k-td/T_0)|^2. \end{align} Performing the change of variable $y=k-td/T_0$, we find \begin{align} \frac{\|\rho_s(t)-\rho_s'(t)\|_1}{ \sqrt{d_s\textup{tr}[\rho^2_s(0)]}} \leq &\,2\varepsilon_v(t,d)+\varepsilon_v^2(t,d)+\sum_{y=\min\{\mathcal{S}(k_0+dt/T_0)\}-td/T_0}^{\max\{\mathcal{S}(k_0+dt/T_0)\}-td/T_0} (\epsilon_y^2+\epsilon_y)|\psi(k_0;y)|^2\\ \leq &\, 2\varepsilon_v(t,d)+\varepsilon_v^2(t,d)+\max_{\bar\kappa\in[0,1]}\sum_{y=k_0-d/2+\bar\kappa}^{k_0+d/2+\bar\kappa} (\epsilon_y^2+\epsilon_y)|\psi(k_0;y)|^2, \end{align} where \begin{equation}\label{eq:epsilon k lemma} \epsilon_k=2\pi\left| \int_{t_i}^t g(x)dx-\int_{k}^{k+td/T_0}V_d(x)dx\right|. \end{equation} and to achieve the last line, we have used the inequalities \begin{align} k_0-d/2+\bar\kappa &\leq \min\{\mathcal{S}(k_0+dt/T_0)\}-td/T_0,\\ k_0+d/2+\bar\kappa &\geq \max\{\mathcal{S}(k_0+dt/T_0)\}-td/T_0,\\ k_0+d/2+\bar\kappa &-(k_0-d/2+\bar\kappa)=d. \end{align} Finally, taking into account Eq. \eqref{eq:g in terms of V_0} followed by a change of variable, Eq. \eqref{eq:epsilon k lemma} can be written as \begin{equation} \epsilon_k=2\pi\left| \int_{0}^{t 2\pi/T_0} V_0(x)dx-\int_{k}^{k+td/T_0}V_d(x)dx\right|=2\pi\left| \int_{0}^{t\, 2\pi/T_0} dx\big( V_0(x)-V_0(x+2\pi k/d)\big)\right|. \end{equation} \end{proof} \begin{figure}[!htb] \minipage{0.32\textwidth} \includegraphics[width=\linewidth]{{epsilon_VPlot.png} \endminipage\hfill \minipage{0.32\textwidth} \includegraphics[width=\linewidth]{{epsilon_VplotMiddle.png}} \endminipage\hfill \minipage{0.32\textwidth}% \includegraphics[width=\linewidth]{{epsilon_VplotRight.png}} \endminipage \caption{{\bf Left}: Plot of $\epsilon_V$ as a function of time over one clock period $T_0$ with cosine potential $V_0(x)=A \cos^{2 n}((x-x_0)/2)$ from Section \ref{sec:The cosine potential} with parameters $k_0=0$, $\sigma=\sqrt{d}$, $T_0=20$, $d=20$, $n=60$, $x_0=\pi$. The unitary is to be implemented at a time $t\approx x_0 \,T_0/(2\pi)=T_0/2$ and the clock Gaussian superposition has its peak initially on the L.H.S. of the potential peak. Before the unitary is applied, to a time $t\approx 5$ the error is practically zero, during a period $t\approx 5$ to $t\approx 15$ the unitary is being applied and the error is large, but after $t\approx 15$, the unitary has been applied and the error is practically zero again. As one increases $n$ and $d$, the errors become arbitrarily small and the interval $t\in[5,15]$ in which there is a large error in the above plot becomes arbitrarily small. {\bf Middle and Right}: Same as Left but changing $x_0$ to $\pi/2$ and $3 \pi /2$ respectiverly, to illustrate how the same qualitative behaviour holds when the unitary is implemented at earlier and later times. \label{fig:epsilon V exmaple plot} \end{figure} \begin{corollary}[Explicit form]\label{corr:explicit bound on system trace distance} Let $V_0(x)\geq 0$ and have a unique global maximum in the interval $x\in[0,2\pi]$ at $x=x_0$ and $g$ be given by Eq. \eqref{eq:g in terms of V_0}. Let $\tilde\epsilon_V$ be such that \begin{equation}\label{eq:tilde ep V def} 1-\tilde\epsilon_V=\int_{x_{vl}}^{x_{vr}} dx\, V_0(x+x_0) \end{equation} for some $-\pi\leq x_{vl}< 0< x_{vr}\leq \pi$, where, for simplicity, we assume $x_{vl}=- x_{vr}$. Let the initial clock Gaussian state be centered at zero; $k_0=0$, and introduce the parameter $0< \gamma_\psi\leq 1$. Furthermore, let $x_0$ be such that $x_{vr}+\pi \gamma_\psi \leq x_0\leq 2\pi-x_{vr}-\pi \gamma_\psi$. Then, for all times $t$ satisfying \begin{equation}\label{eq:t before uni applied} \begin{split} 0&\leq t \,2 \pi/T_0\leq x_0-x_{vr}-\pi \gamma_\psi \end{split} \end{equation} or \begin{equation}\label{eq:t after uni applied} \begin{split} x_0+x_{vr}+\pi \gamma_\psi &\leq t \,2 \pi/T_0\leq 2\pi+x_0 - x_{vr}-\pi\gamma_\psi, \end{split} \end{equation} $\epsilon_V$ defined in Lemma \ref{lem:trace dist bound t-Ham Vs d dim clock}, is bounded by \begin{equation}\label{eq:epsilon V upper bound explicit version} \epsilon_V\leq 4\pi \left( (1+2\pi)A^2\frac{\mathrm{e}^{-2\pi\tilde\kappa\frac{d^2}{\sigma^2}}}{1-\mathrm{e}^{-4\pi\sqrt{\tilde\kappa}\,d/\sigma^2}}+2\tilde\epsilon_V (1+8\pi\tilde\epsilon_V)\right),\quad \tilde\kappa= \begin{cases} 0 &\mbox{if } d\gamma_\psi/2\leq 1\\ (\gamma_\psi/2-1/d)^2 &\mbox{otherwise} \end{cases} \end{equation} where $A$ is given by Eq. \eqref{eq:A normalised} and is a decreasing function of $\sigma$ (see Section \ref{Normalizing the clock state} for bounds). \end{corollary} \emph{Intuition}. If the potential is very peaked around $x_0$ and close to zero away from $x_0$, then one can choose $x_{vl}\approx x_{vr}\approx 0$ with $0 \leq \tilde \epsilon_V\ll 1$. Thus by choosing $\gamma_\psi\approx 0$, the intervals permitted for $x_0$ and $t$ by Eqs. \eqref{eq:t before uni applied} and Eqs. \eqref{eq:t after uni applied}, are almost the full range $[0,2\pi]$ and $[0,T_0]$ respectively, and only for times $t\approx x_0 T_0/(2\pi)$ will the error be large. Eqs. \eqref{eq:t before uni applied} correspond to the times \textit{before} the unitary is applied, while Eqs. \eqref{eq:t after uni applied} correspond to the times \textit{after} the unitary is applied. \begin{remark} The parameters $\gamma_\psi, x_{vr}$ can be chosen to depend on $t$ and/or $d$ in order to improve bounds if desired. \end{remark} \begin{proof} The proof consists in writing the sum in Eq. \eqref{eq:def epsilon k} in terms of three contributions; the left Gaussian tail of $|\psi_\textup{nor}(0;k)|$, the right Gaussian tail of $|\psi_\textup{nor}(0;k)|$, and the contribution from the center of the Gaussian $|\psi_\textup{nor}(0;k)|$. These terms are then all individually bounded.\\ Noting that $0\leq\epsilon_k\leq 2\pi $ and that $|\psi_\textup{nor}(0;k)|$ is normalized, we have from Eq. \eqref{eq:def epsilon k} \begin{align}\label{eq:ep V up bound coll proof} \epsilon_V(t,d)\leq \max_{\bar\kappa\in[0,1]}\Bigg(& \sum_{k=-d/2+\bar\kappa}^{-d\gamma_\psi/2+\bar\kappa}\left( \epsilon_k^2+\epsilon_k\right)|\psi_\textup{nor}(k_0;k)|^2+\sum_{k=d\gamma_\psi /2+\bar\kappa}^{d/2+\bar\kappa}\left( \epsilon_k^2+\epsilon_k\right)|\psi_\textup{nor}(k_0;k)|^2\\ &+\sum_{k=-d\gamma_\psi /2+1+\bar\kappa}^{d\gamma_\psi/2-1+\bar\kappa}\left( \epsilon_k^2+\epsilon_k\right)|\psi_\textup{nor}(k_0;k)|^2 \Bigg)\\ \leq 2\pi(1+2&\pi)\max_{\bar\kappa\in[0,1]} \bigg(\sum_{k=-d/2+\bar\kappa}^{-d\gamma_\psi/2+\bar\kappa}|\psi_\textup{nor}(k_0;k)|^2+\sum_{k=d\gamma_\psi /2+\bar\kappa}^{d/2+\bar\kappa}|\psi_\textup{nor}(k_0;k)|^2\bigg)\label{eq:first line epsilon bound proof}\\ + \max_{\bar\kappa\in[0,1]}&\max_{k\in\mathcal{I}_{\gamma_\psi}}\left( \epsilon_k^2+\epsilon_k\right),\label{eq:last line epsilon bound proof} \end{align} where $\mathcal{I}_{\gamma_\psi}:=\{-d\gamma_\psi /2+1+\bar\kappa,\,-d\gamma_\psi /2+2+\bar\kappa,\ldots,d\gamma_\psi/2-1+\bar\kappa \}$. \\ First we will bound the terms in line \ref{eq:first line epsilon bound proof}. By a change of variable, followed by using Eq. \eqref{eq:psi nor def}, we find \begin{align} & 2\pi(1+2\pi)\max_{\bar\kappa\in[0,1]}\bigg(\sum_{k=-d/2+\bar\kappa}^{-d\gamma_\psi/2+\bar\kappa}|\psi_\textup{nor}(k_0;k)|^2+\sum_{k=d\gamma_\psi /2+\bar\kappa}^{d/2+\bar\kappa}|\psi_\textup{nor}(k_0;k)|^2\bigg)\\ &= 2\pi(1+2\pi)A^2\max_{\bar\kappa\in[0,1]}\bigg(\sum_{x=0}^{d(1+\gamma_\psi)/2}\mathrm{e}^{-\frac{2\pi}{\sigma^2}(x-d\gamma_\psi/2+\bar\kappa)^2}+\sum_{x=0}^{d(1-\gamma_\psi)/2}\mathrm{e}^{-\frac{2\pi}{\sigma^2}(x+d\gamma_\psi/2+\bar\kappa)^2}\bigg)\\ &\leq 2\pi(1+2\pi)A^2\max_{\bar\kappa\in[0,1]}\bigg(\frac{\mathrm{e}^{-\frac{2\pi}{\sigma^2}(d\gamma_\psi/2-\bar\kappa)^2}}{1-\mathrm{e}^{-4\pi|d\gamma_\psi/2-\bar\kappa|/\sigma^2}}+\frac{\mathrm{e}^{-\frac{2\pi}{\sigma^2}(d\gamma_\psi/2+\bar\kappa)^2}}{1-\mathrm{e}^{-4\pi|d\gamma_\psi/2+\bar\kappa|/\sigma^2}}\bigg)\label{eq:used gaus tail bound here}\\ &\leq 4\pi(1+2\pi)A^2\frac{\mathrm{e}^{-2\pi\tilde\kappa\frac{d^2}{\sigma^2}}}{1-\mathrm{e}^{-4\pi\sqrt{\tilde\kappa}\,d/\sigma^2}}, \end{align} with $\tilde \kappa:=\min_{\bar\kappa\in[0,1]} \{ (\gamma_\psi/2-\bar\kappa/d)^2, (\gamma_\psi/2+\bar\kappa/d)^2\}=\min_{\bar\kappa\in[0,1]}(\gamma_\psi/2-\bar\kappa/d)^2$ and where in line \ref{eq:used gaus tail bound here}, the bound from Lemma \ref{G0} has been used.\\ Now we will bound the term in line \ref{eq:last line epsilon bound proof}. By the change of variables $y=x-x_0$, $y'=x+2\pi k/d-x_0$, Eq. \eqref{eq:def epsilon k} can be written in the form \begin{equation}\label{eq:epsilon k intermediate} \epsilon_k=2\pi\left| \int_{-x_0}^{t\, 2\pi/T_0-x_0} dx\, V_0(x-x_0)-\int_{\frac{2\pi}{d}k-x_0}^{t\, 2\pi/T_0+\frac{2\pi}{d}k-x_0} dx \,V_0(x-x_0)\right|. \end{equation} Taking into account that \begin{equation} 1=\int_{-\pi}^\pi dx V_0(x+x_0), \end{equation} together with Eq. \eqref{eq:tilde ep V def} and the periodicity of the potential, it follows that \begin{equation}\label{eq:tilde ep equal} \tilde \epsilon_V=\int_{x_{vr}-2\pi}^{x_{vl}} dx V_0(x+x_0). \end{equation} Thus from Eq. \eqref{eq:epsilon k intermediate}, we have that $\epsilon_k\leq 4\pi \tilde \epsilon_V$ if \begin{align}\label{eq:intermediate befoe uni proof} \begin{split} x_{vr}-2\pi\leq -x_0&\leq t 2\pi/T_0-x_0\leq x_{vl}\\ x_{vr}-2\pi\leq -x_0\leq 2\pi \frac{k}{d}-x_0 &\leq t 2\pi/T_0+2\pi\frac{k}{d}-x_0\leq x_{vl} \quad \forall\, k\in \mathcal{I}_{\gamma_\psi}. \end{split} \end{align} Taking into account $t\geq 0$ and that $\min \mathcal{I}_{\gamma_\psi}= \bar\kappa-d\gamma_\psi/2+1$, $\max \mathcal{I}_{\gamma_\psi}= \bar\kappa+d\gamma_\psi/2-1$, Eqs. \eqref{eq:intermediate befoe uni proof} imply Eqs. \eqref{eq:t before uni applied}. To derive Eqs. \eqref{eq:t after uni applied}, we first note that \begin{align}\label{eq:ep k after uni int} \epsilon_k =&2\pi\Bigg| \int_{-x_0}^{x_{vl}} dx\, V_0(x+x_0)+\int_{x_{vl}}^{x_{vr}} dx\, V_0(x+x_0)+\int_{x_{vr}}^{t\, 2\pi/T_0-x_0} dx\, V_0(x+x_0)\\ &-\int_{\frac{2\pi}{d}k-x_0}^{x_{vl}} dx\, V_0(x+x_0)-\int_{x_{vl}}^{x_{vr}} dx\, V_0(x+x_0)-\int_{x_{vr}}^{t\, 2\pi/T_0+\frac{2\pi}{d}k-x_0} dx\, V_0(x+x_0)\Bigg|\\ \leq &2\pi\Bigg| \int_{-x_0}^{x_{vl}} dx\, V_0(x+x_0)\Bigg|+2\pi\Bigg|\int_{x_{vr}}^{t\, 2\pi/T_0-x_0} dx\, V_0(x+x_0)\Bigg|\\ &+ 2\pi\Bigg|\int_{\frac{2\pi}{d}k-x_0}^{x_{vl}} dx\, V_0(x+x_0)\Bigg|+2\pi\Bigg|\int_{x_{vr}}^{t\, 2\pi/T_0+\frac{2\pi}{d}k-x_0} dx\, V_0(x+x_0)\Bigg|. \end{align} In order for $\epsilon_k$, $k\in \mathcal{I}_{\gamma_\psi}$ to be bounded by $8\pi \tilde \epsilon_V$, the following conditions derived from Eq. \eqref{eq:ep k after uni int} are sufficient: \begin{equation}\label{eq: constrinats 3 coll} \begin{split} x_{vr}-2\pi &\leq -x_0\leq x_{vl}\\ x_{vr}-2\pi &\leq \frac{2\pi}{d}k-x_0\leq x_{vl}\quad \forall \,k\in \mathcal{I}_{\gamma_\psi} \end{split} \end{equation} and \begin{equation}\label{eq: constrinats 4 coll} \begin{split} x_{vr} &\leq t2\pi/T_0-x_0\leq x_{vl}+2\pi\\ x_{vr} &\leq t2\pi/T_0+\frac{2\pi}{d}k-x_0\leq x_{vl}+2\pi\quad \forall \,k\in \mathcal{I}_{\gamma_\psi}. \end{split} \end{equation} Eqs. \eqref{eq: constrinats 3 coll}, \eqref{eq: constrinats 4 coll} imply Eq. \eqref{eq:t after uni applied}. Thus whenever Eqs. \eqref{eq:t before uni applied} or Eqs. \eqref{eq:t after uni applied} are satisfied, the last term in Eq. \eqref{eq:ep V up bound coll proof} is bounded by \begin{equation} \max_{\bar\kappa\in[0,1]}\max_{k\in\mathcal{I}_{\gamma_\psi}}\left( \epsilon_k^2+\epsilon_k\right)\leq 8\pi \tilde \epsilon_V( 8\pi \tilde \epsilon_V+1), \end{equation} \end{proof} Finally, via the following Corollary, we will now show how up to a small correction, the bounds derived in Lemma \ref{lem:trace dist bound t-Ham Vs d dim clock} are the same for $g$ as defined in Section \ref{sec:Consequences of Quasi-Autonomous control}. The asymptotic bounds in the large $d$ limit for the fidelity of the unitary implementation of the system reported in Section \ref{sec:Consequences of Quasi-Autonomous control}, will then be derived in the next Section, \ref{sec: control exmaples}. \begin{corollary}[Section \ref{sec:Consequences of Quasi-Autonomous control} form]\label{lem:new g} Let $V_0$ and $\tilde \epsilon_V$ satisfy the same conditions as in Corollary \ref{corr:explicit bound on system trace distance}, and let $g$ be the normalised pulse from Section \ref{sec:Implementing Energy preserving unitaries via a time dependent Hamiltonian} with support interval $[t_i,t_f]=[t_1,t_2]$ with $t_1=T_0(x_0-x_{vr}-\pi\gamma_\psi)/(2\pi)$, $t_2=T_0(x_0+x_{vr}+\pi\gamma_\psi)/(2\pi)$, then for all $t\in[0,t_1]\cup[t_2,T_0]$ and $0<t_1<t_2<T_0$, \begin{equation} \label{eq: state dist in appendix in main text form} \| \rho_s(t)-\rho'_s(t)\|_1 \leq \sqrt{d_s \textup{tr}[\rho_s^2(0)]} \left( 2\varepsilon_v(t,d)+\varepsilon_v^2(t,d)+\varepsilon_s(t,d) \right), \end{equation} where \begin{equation} \varepsilon_s(t,d) :=\epsilon_V(d)+2\pi T_0 \tilde \epsilon_V(2\pi T_0\tilde{\epsilon}_V+1), \end{equation} with $\epsilon_V(d)$ bounded by Eq. \eqref{eq:epsilon V upper bound explicit version}, and $\tilde{\epsilon}_V$ satisfies Eq. \eqref{eq:tilde ep V def}. $\rho_s(t)$ is given by Eq. \eqref{eq:rho t scho sol commute t depen} and $\rho'_s(t)=\textup{tr}[\rho'_{sc}(t)]$ by Eq. \eqref{eq:rho finite s c}. \end{corollary} \begin{proof} Let $g_1$ denoted $g$ as defined in this Corollary and $g_2$ denote $g$ as defined via Eq. \eqref{eq:g in terms of V_0}. With these functions, define $\epsilon_g(t)=\int_0^t dx(g_2(x)-g_1(x))$, $\rho_{S,q}(t)=U_q(t)\rho_s(0) U_q^\dag(t),$ $U_q(t)=\mathrm{e}^{-\mathrm{i} \hat H_s^{int}\int_0^t g_q(t)}$, $q=1,2$. First note \begin{equation} \int_{t_1}^{t_2}dx g_2(x)=\frac{2\pi}{T_0}\int_{t_1}^{t_2}dx V_0(x 2\pi/T_0) =\int_{-x_{vr}-\pi\gamma_\psi}^{x_{vr}+\pi\gamma_\psi}dx V_0(x+x_0)\geq \int_{-x_{vr}}^{x_{vr}}dx V_0(x+x_0)=1-\tilde\epsilon_V. \end{equation} Thus \begin{equation} 1=\int_0^{T_0} dx\, g_2(x) \geq \int_0^{t_1} dx\, g_2(x)+1-\tilde{\epsilon}_V +\int_{t_2}^{T_0} dx\, g_2(x)\implies \tilde{\epsilon}_V\geq \begin{cases} \int_0^{t_1} dx\, g_2(x)\\ \vspace{-0.3cm}\\ \int_{t_2}^{T_0} dx \,g_2(x) \end{cases} \end{equation} \label{eq:tilde epsilon V} Using Eq. \eqref{eq:tilde epsilon V}, $\|\cdot\|_1/\sqrt{d_s}\leq \|\cdot\|_F$, where $\|\cdot\|_F$ is the Frobenius norm for a $d_s$ dimensional system, and the bounds $\sin(a)\leq |a|$, $-\cos(a)\leq a^2-1$, $a\in{\mathbbm{R}}$, we find \begin{align}\label{eq:trigabgle term boud} \| \rho_{s,1}(t)-\rho_{s,2}(t)\|_1 &= \| \rho_{s}(0)-U_1^\dag U_2(t)\rho_{s}(0) U_2^\dag U_1(t)\|_1\leq \sqrt{d_s} \| \rho_{s}(0)-U_1^\dag U_2(t)\rho_{s}(0) U_2^\dag U_1(t)\|_F\\ &= \sqrt{d_s} \sqrt{\sum_{m,n} \big{|}\rho_{m,n}(t)-\mathrm{e}^{-\mathrm{i} (\Omega_m-\Omega_n)\epsilon_g(t)} \rho_{m,n}(t)\big{|}^2}\leq \sqrt{d_s\textup{tr}[\rho_s^2(0)]}\max_{m,n}|1-\mathrm{e}^{-\mathrm{i} (\Omega_m-\Omega_n)\epsilon_g(t)}|\\ & \leq 2\pi(2\pi|\epsilon_g(t)|+1)|\epsilon_g(t)|. \end{align} Using Eq. \eqref{eq:tilde epsilon V}, we thus conclude $|\epsilon_g(t)|\leq \tilde{\epsilon}_V T_0$, for $t\in[0,t_1]\cup[t_2,T_0]$. Finally, note that $\| \rho_s(t)-\rho_{s,1}(t)\|_1 \leq \| \rho_s(t)-\rho_{s,2}(t)\|_1 +\| \rho_{s,1}(t)-\rho_{s,2}(t)\|_1$ via the triangle inequality. $\| \rho_s(t)-\rho_{s,2}(t)\|_1$ was bounded in Eq. \eqref{eq:trace dist rho rho finite and time}, with $\epsilon_V(t,d)$ bounded in Corollary \ref{corr:explicit bound on system trace distance} for $t\in[0,t_1]\cup[t_2,T_0]$. Using these two bounds and Eq. \eqref{eq:trigabgle term boud}, Eq, \eqref{eq: state dist in appendix in main text form} follows. To finalize the proof, note that $x_0=\frac{t_2-t_1}{2}\frac{2\pi}{T_0}$, $x_{vr}+\pi\gamma=\frac{t_2-t_1}{2}\frac{2\pi}{T_0}$. As such, the condition $x_{vr}+\pi \gamma_\psi \leq x_0\leq 2\pi-x_{vr}-\pi \gamma_\psi$ from Lemma \ref{corr:explicit bound on system trace distance} is satisfied for all $0<t_1<t_2<T_0$ and Eq. \eqref{eq:epsilon V upper bound explicit version} is valid. \end{proof} \subsubsection{Examples}\label{sec: control exmaples} We can now bound $\tilde{\epsilon}_V$ for the cosine potential defined in Section \ref{sec:The cosine potential} with the aim of understanding the significance of Corollaries \ref{corr:explicit bound on system trace distance}, \ref{lem:new g}. We find the bound \begin{equation}\label{eq: up bound tilde epsilon V} \tilde \epsilon_V\leq \frac{(\pi-x_{vr})\mathrm{e}^2}{4\pi\sqrt{\pi}}\sqrt{n}\cos^{2n}(x_{vr}/2), \quad\text{if } \cos(x_{vr})\leq 1-\frac{1}{n}. \end{equation} By choosing $n=n(d)$ in different ways, one can achieve different decay rates. This choice will also effect the decay rates of $\varepsilon_v(t,d)$ defined in Eq. \ref{eq:ltwo epsilon nor control theorem}, the significance of which will be clarified in Section \ref{Clock Fidelity}. We give three examples of $n=n(d)$: \begin{itemize} \item[1)]\emph{Power law decay.} Let \begin{equation} n=\gamma_1\frac{\ln d}{-2\ln \cos(x_{vr}/2)},\quad \gamma_1>0. \end{equation} Then \begin{equation} \tilde \epsilon_V\leq \frac{(\pi-x_{vr})\mathrm{e}^2}{4\pi\sqrt{\pi}} \sqrt{\gamma_1\frac{\ln d}{-2\ln \cos(x_{vr}/2)}}\,d^{-\gamma_1}, \end{equation} if \begin{equation}\label{eq:gamma 1 exmaple condition} \gamma_1\geq \frac{-2\ln \cos(x_{vr}/2)}{\left(1-\cos(x_{vr})\right)}\frac{1}{\ln d}. \end{equation} Note that $\lim_{d\rightarrow \infty} \tilde\epsilon_V d^m=0$ for all $\gamma_1>m>0$. Using the definition of the clock rate parameter $\bar\upsilon$ (Def. \ref{def:decay rate params}), we find \begin{equation} \bar\upsilon =\frac{\ln d}{\ln (\pi \alpha_0 \sigma^2)} \frac{2 \mathrm{e}^2}{(2\pi)^{3/2} \sqrt{2}}\frac{\pi \kappa \alpha_0}{(-2\ln\cos (x_{vr}/2))^{3/2}}\gamma_1^{3/2} \sqrt{\ln d}. \end{equation} Thus from Eq. \eqref{eq:ltwo epsilon nor control theorem}, for $\sigma=\sqrt{d}$ \begin{equation}\label{eq:epsilon decay rate exmaple 1} \varepsilon_v(t,d)=\mathcal{O}\left(t\, poly(d)\,\exp\left(-\frac{\pi}{4}\frac{\alpha_0^2 \chi_1^2 \gamma_1^3}{\left( 1+ \chi_1 \gamma_1^{3/2}/\sqrt{\ln d} \right)^2} \frac{d}{\ln d}\right) \right), \end{equation} with \begin{equation} \chi_1:= \frac{(2\pi)^{3/2}\sqrt{2}}{2\,\mathrm{e}^2} \left(1+ \frac{\ln(\pi\alpha_0)}{\ln d}\right) \frac{\left(-2\ln(\cos x_{vr}/2)\right)^{3/2}}{\pi \kappa \alpha_0} \end{equation} Furthermore, using Eq. \eqref{eq:epsilon V upper bound explicit version}, for Eqs. \ref{eq:trace dist rho rho finite and time}, \eqref{eq: state dist in appendix in main text form} one has that \begin{equation} \frac{\|\rho_s(t)-\rho'_s(t)\|_1}{\sqrt{d_s\textup{tr}[\rho_s^2(0)]}}=\mathcal{O}\left(\varepsilon_v(t,d)\right)+\mathcal{O}\left(\mathrm{e}^{-2\pi\tilde k (d/\sigma)^2}\right)+\mathcal{O}\left(\tilde\epsilon_V\right). \end{equation} Thus since $\varepsilon_v(t,d)$ decays faster than any power of $d$, for $\sigma=\sqrt{d}$ and any $\gamma_\psi\in(0,1]$, we have that \begin{equation} \lim_{d\rightarrow \infty} \|\rho_s(t)-\rho'_s(t)\|_1 \,d^m=0, \end{equation} for all constants $\gamma_1>m>0$. \item[2)] \emph{System error faster than power-law decay.} By parametrizing $\gamma_1$ in example 1) to increase sufficiently slowly with $d$ rather than being constant, we can achieve that the trace distance error between $\rho_s$ and $\rho_s'$ (Eqs. \eqref{eq:trace dist rho rho finite and time} and \eqref{eq: state dist in appendix in main text form}) decays with $d$ faster than any power law in $d^m$ for the optimal clock decay $\sigma=\sqrt{d}$. The penalty of this improvement, will be a worse decay rate for $\varepsilon_v(t,d)$ than that of Eq. \eqref{eq:epsilon decay rate exmaple 1} . Let \begin{equation}\label{eq:gamma 1 equlity eg} \gamma_1=\frac{\pi}{4}\alpha_0^2 \chi_2^2\frac{d^{1/4}}{\sqrt{\ln d}},\quad \chi_2:=\chi_1\left(\frac{\pi}{4}\alpha_0^2\chi_1^2\right)^{-3/8}. \end{equation} such that Eq. \eqref{eq:gamma 1 exmaple condition} is satisfied, then for $\sigma=\sqrt{d}$, \begin{align}\label{eq:epsilon decay rate exmaple 1,2} \varepsilon_v(t,d)&=\mathcal{O}\left(t\, poly(d)\,\,\mathrm{e}^{- \frac{\pi}{4}\alpha_0^2 \chi_2^2 \,d^{1/4}\sqrt{\ln d}}\,\right)\\ \tilde \epsilon_V&\leq \frac{(\pi-x_{vr})\mathrm{e}^2 \alpha_0\chi_2}{8\pi} \frac{(d\ln d)^{1/4}}{\sqrt{-2\ln \cos(x_{vr}/2)}}\,\mathrm{e}^{- \frac{\pi}{4}\alpha_0^2 \chi_2^2 \,d^{1/4}\sqrt{\ln d}}, \end{align} and thus up to polynomial factors, both $\varepsilon_v(t,d)$ and $\tilde\epsilon_V$ have exponential decay in $d^{1/4}\sqrt{\ln d}$ and \begin{equation} \frac{\|\rho_s(t)-\rho'_s(t)\|_1}{\sqrt{d_s\textup{tr}[\rho_s^2(0)]}}=\mathcal{O}\left(\left( t \,poly(d) +(d\ln d)^{1/4}\right)\mathrm{e}^{- \frac{\pi}{4}\alpha_0^2 \chi_2^2 \,d^{1/4}\sqrt{\ln d}}\,\right), \end{equation} from which we conclude \begin{equation}\label{eq: faster than power decay trace dist} \lim_{d\rightarrow \infty} \|\rho_s(t)-\rho_s'(t)\|_1\, d^m=\lim_{d\rightarrow \infty} \varepsilon_v(t,d)\, d^m=0, \end{equation} for all constant $m>0$ and for all $t$ such that Eq. \eqref{eq:t before uni applied} and Eq. \eqref{eq:t after uni applied} hold. Furthermore, the constraint of Eq. \eqref{eq:gamma 1 exmaple condition} can be re-written in terms of the new parametrizations to achieve, using Eqs. \eqref{eq:gamma 1 exmaple condition} and \eqref{eq:gamma 1 equlity eg} the constraint \begin{equation} \frac{4}{\pi\alpha_0^2\chi_2^2}\frac{-2\ln\cos(x_{vr}/2)}{1-\cos (x_{vr})}\leq d^{1/4}\sqrt{\ln d}. \end{equation} This constraint is satisfied for all constant $x_{vr}\in(0,2\pi)$ for sufficiently large $d$ (recall that by definition, $\chi_2$ depends on $x_{vr}$). Similarly, $\gamma_\psi$ can be any constant in $(0,1)$. \item[3)]\emph{Smallest clock error.} In this case, we will bound how quickly the error $\tilde \epsilon_V$ can decay with $d$ while maintaining the smallest disturbance imposed on the clock due to performing the unitary (see section \ref{Clock Fidelity} for details), that is to say, while maintaining a constant rate parameter $\bar\upsilon$, for the optimal clock decay $\sigma=\sqrt{d}$.\\ Let \begin{equation} n=\frac{\gamma_3}{-2\ln\cos(x_{vr})} (\ln d)^{2/3},\quad \gamma_3>0. \end{equation} Then $\tilde\epsilon_V$ decays with increasing $d$ at a rate \begin{equation} \tilde\epsilon_V\leq \frac{(\pi-x_{vr})\mathrm{e}^2}{4\pi\sqrt{\pi}} \sqrt{\frac{\gamma_3}{-2\ln\cos(x_{vr})}} (\ln d)^{2/3} d^{-\gamma_3(\ln d)^{-1/3}}, \quad\text{if } \gamma_3\geq \frac{-2\ln \cos(x_{vr})}{\left(1-\cos(x_{vr})\right)}\frac{1}{(\ln d)^{2/3}}, \end{equation} and the clock rate parameter (Def. \ref{def:decay rate params}) is \begin{equation} \bar\upsilon =\frac{\ln d}{\ln (\pi \alpha_0 \sigma^2)} \frac{2 \mathrm{e}^2}{(2\pi)^{3/2} \sqrt{2}}\frac{\pi \kappa \alpha_0}{(-2\ln\cos (x_{vr}))^{3/2}}\gamma_3^{3/2}, \end{equation} which is bounded in the $d\rightarrow \infty$ limit. Thus for this parametrisation, $\varepsilon_v(t,d)$ has exponential decay in $d$ and consequently the clock's disturbance decays exponentially when $\sigma=\sqrt{d}$, (recall Theorem \ref{movig through finite time}). \end{itemize} We now present a quick proof of Eq. \eqref{eq: up bound tilde epsilon V} \begin{proof} From Eqs. \eqref{eq:tilde ep equal} and \eqref{def: cosine pot} we find \begin{align}\label{eq: tilde ep V int written in convex form} \tilde \epsilon_V&=\int_{x_{vr}-2\pi}^{x_{vl}}dx A_c \cos^{2 n}\left(\frac{x}{2}\right)=A_c \left( \int_{x_{vr}-2\pi+2\pi}^{-\pi+2\pi} dx \cos^{2n}(x/2-\pi)-\int_{\pi}^{-x_{vl}}dx \cos^{2n}(-x/2) \right)\\ &= 2 A_c\int_{x_{vr}}^\pi dx \cos^{2 n}(x/2)=2 A_c (\pi-x_{vr})\int_0^1 dy \cos^{2 n}\left( (y\, x_{vr}+(1-y) \pi)/2 \right). \end{align} We now note the convexity of the potential, namely on the interval $x\in(x_{vr},\pi)$, \begin{equation} \frac{d^2\;}{dx^2} \cos^{2n}(x/2)=\frac{n}{2}\cos^{2(n-1)}(x/2) \left(n(1-\cos(x))-1\right)\geq 0 \implies \cos(x)\leq 1-\frac{1}{n}, \end{equation} thus noting that $\max_{x\in[x_{vr},\pi]}\cos(x)=\cos(x_{vr})$, $\cos^{2n}(x/2)$ is convex on $x\in[x_{vr},\pi]$ if $\cos(x_{vr})\leq 1-1/n$ and $ \cos^{2 n}\left( (y\, x_{vr}+(1-y) \pi)/2 \right)\leq y \cos^{2 n}(x_{vr}/2)+(1-y) \cos^{2 n}(\pi/2) =y \cos^{2 n}(x_{vr}/2)$, $y\in[0,1]$. Hence using Eq. \eqref{eq: tilde ep V int written in convex form}, we conclude \begin{equation} \tilde \epsilon_V\leq A_c(\pi-x_{vr})\cos^{2n}(x_{vr}/2), \quad\text{if } \cos(x_{vr})\leq 1-\frac{1}{n}. \end{equation} Using Eq. \eqref{def: A c eq def for cos pot} with $\Omega=1$ and Sterling's formula, we find \begin{equation} A_c= \frac{2^{2 n}}{2\pi \binom{2n}{n}}=\frac{2 ^{2 n}}{2\pi}\frac{(n!)^2}{(2n)!}\leq \frac{\mathrm{e}^2}{4\pi\sqrt{\pi}}\sqrt{n}, \end{equation} thus giving us \begin{equation} \tilde \epsilon_V\leq \frac{(\pi-x_{vr})\mathrm{e}^2}{4\pi\sqrt{\pi}}\sqrt{n}\cos^{2n}(x_{vr}/2), \quad\text{if } \cos(x_{vr})\leq 1-\frac{1}{n}. \end{equation} \end{proof} \subsection{How to perform any timed unitary} \label{sec:non energy preserving unitaries} So far we have concerned ourselves with performing energy preserving unitaries only. Often, one may want to perform any unitary on the system $\mathcal{H}_s$. The clock provides a source of timing, but not energy. To perform non-energy preserving unitaries, one will need an energy source. To achieve this, one can introduce a battery system with Hilbert space $\mathcal{H}_b$. Then, the clock can perform energy preserving unitaries on a state $\rho_{sb}$ on $\mathcal{H}_s\otimes\mathcal{H}_b$ according the procedure described in Section \ref{sec:Implementing Energy preserving unitaries with the finite}. The initial state of the battery is chosen such that it has enough energy to implement the unitary and will be such that an energy preserving unitary over $U_{sb}$ on a state $\rho_s\otimes\rho_b$ will induce any desired unitary $U'_{s}$, on $\rho_s$, namely $\textup{tr}_b[U_{sb} \rho_s\otimes\rho_b U_{sb}^\dag]= U'_{s} \rho_s U^{\prime \dag}_{s}$, where $[U_{sb},\hat H_s\otimes{\mathbbm{1}}_b+{\mathbbm{1}}_s\otimes \hat H_s]=0$, $[U'_{s},\hat H_s ]\neq 0$. The procedure which allows one to find a state $\rho_b\in\mathcal{S}(\mathcal{H}_b)$, a Hamiltonian $\hat H_b$ on $\mathcal{H}_b$ and an energy preserving unitary $U_{sb}$ on $\mathcal{H}_{sb}$, such that any desired unitary $U'_s$ on $\mathcal{H}_s$ can be performed, is the topic of \cite{Aberg}. Therefore, using the finite dimension clock together with a battery as discussed in this Section, one can achieve any timed non-energy preserving unitary on states on $\mathcal{H}_s$ using the results of \cite{Aberg}. This will be developed with explicit error bounds in an upcoming paper. \subsection{Clock Fidelity}\label{Clock Fidelity} If the potential $V_0(x)$ is zero for all $x\in{\mathbbm{R}}$, then the system on $\mathcal{H}_s$ and the clock on $\mathcal{H}_c$ will evolve in time as independent systems. After a time $T_0$, as we have seen, the clock state will return to its initial state. However, whenever $V_0\neq 0$, the system-clock interaction will disturb the dynamics of the clock such that after a time $T_0$, the clock state will not have returned to its initial state. Lemma \ref{Lem:clock fidelity} in this Section bounds quantitatively how large this disturbance is, and how quickly it decays with clock dimension \begin{lemma}[Clock disturbance]\label{Lem:clock fidelity} Let $\rho_c'(t):=\textup{tr}[\rho_{sc}'(t)]$, $\rho_{sc}'$ defined in Eq. \eqref{eq:rho finite s c}, be the state of the clock at time $t$. We have the bound, \begin{equation}\label{eq:trace distance bound} \frac{1}{2}\| \rho_c'(0)-\rho_c'(T_0)\|_1\leq \varepsilon_v(T_0,d), \end{equation} where $\varepsilon_v(\cdot,\cdot)$ is defined in Eq. \eqref{eq:ltwo epsilon nor control theorem}. \end{lemma} \emph{Intuition} The bound on the trace distance given by Eq. \eqref{eq:trace distance bound}, decays quickly with $d$ while increasing linearly in $T_0$ as is evident from the definition of $\varepsilon_v(\cdot,\cdot)$. It bounds the disturbance which is \textit{only} originating from the implementation of the unitary on the system on $\mathcal{H}_s$, since $\| \rho_c'(0)-\rho_c'(T_0)\|_1=0$ if $V_0 (x)=0$ for all $x\in {\mathbbm{R}}$. \begin{proof} The result follows from Theorem \ref{eq: main eq in control theorem} and simple identities.\\ By taking the partial trace over the system Hilbert space $\mathcal{H}_s$, from Eq. \eqref{eq:rho finite s c} we achieve\\ From Eq. \eqref{eq:rho finite s c}, it follows \begin{equation}\label{} \rho_{sc}^{\prime}(t)= \sum_{m,n=1}^{d_s} \rho_{m,n}(t)\ketbra{\phi_m}{\phi_n}\otimes\ketbra{\bar\Phi_m(t)}{\bar\Phi_n(t)}_c, \end{equation} where we have used definitions Eqs. \eqref{eq:rho s in mutual alphog basis}, \eqref{eq:Psi bar def eq}. Taking the partial trace over the system Hilbert space $\mathcal{H}_s$, we achieve \begin{equation} \rho_{c}^{\prime}(t)= \sum_{n=1}^{d_s} \rho_{n,n}(0)\ketbra{\bar\Phi_n(t)}{\bar\Phi_n(t)}_c. \end{equation} Thus the \textit{quantum Fidelity} $F$ is \begin{align}\label{eq: clock inital vs final fidelity in proof 1} F( \rho_c'(0),\rho_c'(t))&=\textup{tr}\left[\sqrt{\sqrt{\rho_{c}^{\prime}(0)}\rho_{c}^{\prime}(t)\sqrt{\rho_{c}^{\prime}(t)}}\right]\\ &=\textup{tr}\left[\sqrt{\ketbra{\Psi_\textup{nor}(k_0)}{\Psi_\textup{nor}(k_0)}\left(\sum_{n=1}^{d_s} \rho_{n,n}(0)\hat\Gamma_n(t) \ketbra{\Psi_\textup{nor}(k_0)}{\Psi_\textup{nor}(k_0)}\hat\Gamma_n^\dag(t)\right)\ketbra{\Psi_\textup{nor}(k_0)}{\Psi_\textup{nor}(k_0)}}\right]\\ &=\textup{tr}\left[\sqrt{\ketbra{\Psi_\textup{nor}(k_0)}{\Psi_\textup{nor}(k_0)}\left(\sum_{n=1}^{d_s} \rho_{n,n}(0)\left|\bra{\Psi_\textup{nor}(k_0)}\hat\Gamma_n(t)\ket{\Psi_\textup{nor}(k_0)}\right|^2\right)}\right]\\ &=\sqrt{\sum_{n=1}^{d_s} \rho_{n,n}(0)\left|\bra{\Psi_\textup{nor}(k_0)}\hat\Gamma_n(t)\ket{\Psi_\textup{nor}(k_0)}\right|^2}, \end{align} where we have defined \begin{equation} \hat\Gamma_n(t):= \mathrm{e}^{-\mathrm{i} t( \hat{H}_c + \Omega_n \hat{V}_d)}. \end{equation} Applying Theorem \ref{eq: main eq in control theorem}, we find \begin{align}\label{eq:modulus sq of exp Gamma t} \left|\bra{\Psi_\textup{nor}(k_0)}\hat\Gamma_n(T_0)\ket{\Psi_\textup{nor}(k_0)}\right|^2 &=\left(\mathrm{e}^{-\mathrm{i}\Omega_n}+\braket{\Psi_\textup{nor}(k_0)|\epsilon_n}\right) \left(\mathrm{e}^{-\mathrm{i}\Omega_n}+\braket{\Psi_\textup{nor}(k_0)|\epsilon_n}\right)^*\\ &=1+\braket{\Psi_\textup{nor}(k_0)|\epsilon_n}\mathrm{e}^{+\mathrm{i} \Omega_n}+\braket{\epsilon_n|\Psi_\textup{nor}(k_0)}\mathrm{e}^{-\mathrm{i} \Omega_n}+|\braket{\epsilon_n|\Psi_\textup{nor}(k_0)}|^2\\ &\geq 1+\braket{\Psi_\textup{nor}(k_0)|\epsilon_n}\mathrm{e}^{+\mathrm{i} \Omega_n}+\braket{\epsilon_n|\Psi_\textup{nor}(k_0)}\mathrm{e}^{-\mathrm{i} \Omega_n}. \end{align} However, \begin{align} \left| \bra{\Psi_\textup{nor}(k_0)} \Gamma_n^\dagger(T_0) \Gamma_n(T_0) \ket{\Psi_\textup{nor}(k_0)} \right|^2 &= 1 \\ \therefore \left( \bra{\Psi_\textup{nor}(k_0)} \mathrm{e}^{\mathrm{i}\Omega_n} + \bra{\epsilon_n} \right) \left( \mathrm{e}^{-\mathrm{i}\Omega_n} \ket{\Psi_\textup{nor}(k_0)} + \ket{\epsilon_n} \right) &= 1 \\ \therefore \braket{\Psi_\textup{nor}(k_0)|\epsilon_n}\mathrm{e}^{+\mathrm{i} \Omega_n}+\braket{\epsilon_n|\Psi_\textup{nor}(k_0)}\mathrm{e}^{-\mathrm{i} \Omega_n} &= -\braket{\epsilon_n|\epsilon_n} = -||\ket{\epsilon_n}||^2_2, \end{align} Thus from Eq. \eqref{eq:modulus sq of exp Gamma t}, we achieve \begin{equation} \left|\bra{\Psi_\textup{nor}(k_0)}\hat\Gamma_n(T_0)\ket{\Psi_\textup{nor}(k_0)}\right|^2 \geq 1-||\ket{\epsilon_n}||^2_2. \end{equation} Plunging into Eq. \eqref{eq: clock inital vs final fidelity in proof 1}, \begin{equation} F( \rho_c'(0),\rho_c'(T_0))\geq \sqrt{\sum_{n=1}^{d_s} \rho_{n,n}(0)\left(1-||\ket{\epsilon_n}||^2_2\right)}\geq \sqrt{\min_{n\in{1,\ldots,d_s}}\left(1-||\ket{\epsilon_n}||^2_2\right)}=\sqrt{1- \varepsilon_v^2(T_0,d)}, \end{equation} where $\varepsilon_v(\cdot,\cdot)$ is defined in Eq. \eqref{eq:ltwo epsilon nor control theorem}. Thus we conclude the Lemma using the well known relationship between Fidelity and trace distance, namey $\frac{1}{2} \|\rho-\sigma\|_1\leq \sqrt{1-F^2(\rho,\sigma)}$, for two quantum states $\rho$, $\sigma$ \end{proof} \onecolumngrid \section{Commutator and uncertainty relations for the ideal, SWP, and Gaussian clock states}\label{appendixcommutator} \subsection{Proof of Quasi-Canonical commutator} For simplicity, in this Section we will fix the dimension $d$ to be odd, and shift spectrum of $\hat H_c$ and $\hat t_c$ to be centered at zero, giving us \begin{align} \hat{H}_c &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} n \frac{2\pi}{T_0} \ketbra{E_n}{E_n}, & \hat{t}_c &= \sum_{k = -\frac{d-1}{2}}^{+\frac{d-1}{2}} k \frac{T_0}{d} \ketbra{\theta_k}{\theta_k}. \end{align} The shifting of the spectrum clearly has no physical effect on the dynamics nor the commutator of these operators since the ground state and initial time can be sifted back to zero by simply adding a term proportional to the identity operator to the r.h.s. of $\hat H_c$ and $\hat t_c$. The mean energy of the state $n_0$ before defined on $n_0\in(0,d-1)$ in this Section is now shifted to $n_0\in\left(-\frac{d}{2},\frac{d}{2}\right)$. We will also assume $k_0\in\left(-\frac{d}{2},\frac{d}{2}\right)$ in this Section for simplicity and define. \begin{align}\label{eq:alpha_c def eq} \bar\alpha &= \abs{\frac{2n_0}{d}}, & \bar\beta &= \abs{\frac{2 k_0}{d}}. \end{align} Thus both $\bar\alpha, \bar\beta \in [0,1)$. They are both measures of how close to the edge of the spectra of energy and time the state is; c.f. similar measure Def. \eqref{Distance of the mean energy from the edge}. \begin{theorem}[Quasi-Canonical commutation]\label{Quasi-Canonical commutation} For all states $\ket{\Psi_\textup{nor}(k_0)}\in \Lambda_{\sigma,n_0}$, the time operator $\hat t_c$ and Hamiltonian $\hat{H_c}$ satisfy the commutation relation \begin{align} \left[ \hat{t}_c, \hat{H}_c \right] \ket{\Psi_\textup{nor}(k_0)} &= \mathrm{i} \ket{\Psi_\textup{nor}(k_0)} + \ket{\epsilon_\textup{comm}}, \end{align} where \begin{align}\label{eq: commutation finite dim theorem} \|\ket{\epsilon_\textup{comm}}\|_2 \leq &\, \epsilon_8^{co}+\frac{1}{2}\left( \epsilon_7^{co}+\epsilon_6^{co}+\pi d\, \epsilon_5^{co} \right)+\epsilon_4^{co}+\epsilon_3^{co}+\pi d \left( \epsilon_2^{co}+\epsilon_1^{co} \right)\\ =&\left( \mathcal{O}\left(d\sigma^{1/2}\right)+\mathcal{O}\left(d\sigma^{5/2}\right) \right)\mathrm{e}^{-\frac{\pi}{4}d^2(1-\bar\alpha)^2}+ \left( \mathcal{O}\left(\frac{d^2}{\sigma^{5/2}}\right)+\mathcal{O}\left(d\sigma^{1/2}\right) \right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}(1-\bar\beta)^2}\\ &+\left( \mathcal{O}\left(\frac{d^3}{\sigma^{5/2}}\right)+\mathcal{O}\left(\frac{d^2}{\sigma^{1/2}}\right)+\mathcal{O}\left(d^2\sigma^{3/2}\right) \right)\mathrm{e}^{-\frac{\pi}{4}\frac{d^2}{\sigma^2}}, \end{align} where $\epsilon_1^{co}$ is given by Eq. \eqref{eq:ep1 comm bound}, $\epsilon_2^{co}$ by Eq. \eqref{eq:ep2 comm bound}, $\epsilon_3^{co}$ by Eq. \eqref{eq:ep3 comm bound}, $\epsilon_4^{co}$ by Eq. \eqref{eq:ep4 comm bound}, $\epsilon_5^{co}$ by Eq. \eqref{eq:ep5 comm bound}, $\epsilon_6^{co}$ by Eq. \eqref{eq:ep6 comm bound}, $\epsilon_7^{co}$ by Eq. \eqref{eq:ep7 comm bound}, and $\epsilon_8^{co}$ by Eq. \eqref{eq:ep8 comm bound}. \end{theorem} \emph{Intuition} It is well known that operators $\hat a$, $\hat b$ satisfying $[\hat a, \hat b]=\mathrm{i}$ can only exist in infinite dimensions. The above results shows that, although this is the case, such operators can exist up to a small error; --exponentially small in $d$ in when the state $\ket{\Psi_\textup{nor}(k_0)}$ is symmetric (i.e., when $\sigma=\sqrt{d}$). Also see the main text and Sections \ref{idealcommutator}, \ref{Shortcomings of the time-states} for more discussions on the topic for a physical intuition and how this result further relates our Gaussian clock states to the idealized clock case. We also observe that by definition, the l.h.s. of Eq. \ref{eq: commutation finite dim theorem} is $T_0$ independent, and thus so is the error term $\ket{\epsilon_\textup{comm}}$ \begin{remark} Although the context in \cite{sergedft} is different to the subject at hand here, by making the identifications $\hat t_c$ and $\hat H_c$ with the discrete position and momentum operators respectively in \cite{sergedft}, Theorem \ref{Quasi-Canonical commutation} proves the conjecture in \cite{sergedft} that the discrete position and momentum operators commutator relation on Gaussian states is satisfied up to a small error. \end{remark} \begin{proof} The proof will consist in going back and forth between representations of the state in the energy basis and the time basis, bounding summations over Gaussian tails while making crucial use of the Poisson summation formula to bound such approximations in the process. The identities used during the proof for the Poisson summation formula can be found in Section \ref{Poisson summation formula}. For simplicity, we calculate the commutator for $k_0$ and $n_0$ both greater than zero, the general proof is analogous. Applying the time operator on the state, \begin{align} \hat{t}_c \ket{\Psi(k_0)} &= \sumt{l} \sumS{k}{k_0} \psi(k_0;k) \ket{\theta_k}. \end{align} Since $k_0 \in [ 0, d/2 ]$, for $k \in \mathcal{S}_d(k_0)$, \begin{align} \hat{t}_c \ket{\theta_k} = \begin{cases} k \frac{T_0}{d} \ket{\theta_k} & \text{if $k < \frac{d}{2} $,} \\ (k-d) \frac{T_0}{d} \ket{\theta_k} & \text{if $k > \frac{d}{2} $.} \end{cases} \end{align} Accordingly, \begin{align} \hat{t}_c \ket{\Psi(k_0)} &= \sumS{k}{k_0} \frac{T_0}{d} k \psi(k_0;k) \ket{\theta_k} + \ket{\epsilon_1}, \\ \text{where} \; \ket{\epsilon_1} &= - \sum_{k \in \mathcal{S}_d(k_0), k > \frac{d}{2}} T_0 \psi(k_0;k) \ket{\theta_k}, \end{align} \begin{equation}\label{eq:ep1 comm bound} \epsilon_1^{co}:= \ltwo{\ket{\epsilon_1}}/T_0 < \begin{cases} \left[ 2 A \frac{e^{-\frac{\pi d}{4} (1-\beta)^2}}{1 - e^{-\pi(1-\beta)}} \right] \mbox{ if } \sigma=\sqrt{d}\\ 2 A \frac{e^{-\frac{\pi d^2}{4 \sigma^2}(1-\bar\beta))^2}}{1 - e^{-\frac{\pi d}{\sigma^2}(1-\bar\beta))}} \mbox{ otherwise} \end{cases} \end{equation} In anticipation of applying the Hamiltonian next, we move to the basis of energy eigenstates, \begin{align} \hat{t}_c \ket{\Psi(k_0)} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{k \in \mathcal{S}_d(k_0)} \frac{T_0}{d} k \psi(k_0;k) \frac{e^{-i 2\pi n k/d}}{\sqrt{d}} \ket{E_n} + \ket{\epsilon_1}. \end{align} We approximate the sum w.r.t. $k$ by the infinite sum and bound the difference, \begin{align} \hat{t}_c \ket{\Psi(k_0)} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{k \in \mathbb{Z}} \frac{T_0}{d} k \psi(k_0;k) \frac{e^{-i 2\pi n k/d}}{\sqrt{d}} \ket{E_n} + \ket{\epsilon_2} + \ket{\epsilon_1}, \\ \text{where} \; \ket{\epsilon_2} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{k \in \mathbb{Z}/\mathcal{S}_d(k_0)} \frac{T_0}{d} k \psi(k_0;k) \frac{e^{-i 2\pi n k/d}}{\sqrt{d}} \ket{E_n} \end{align} \begin{equation}\label{eq:ep2 comm bound} \epsilon_2^{co}:= \ltwo{\ket{\epsilon_2}}/T_0 < \begin{cases} A \sqrt{d} \left( 1 + \frac{1}{\pi} + \frac{\bar\beta}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} \mbox{ if } \sigma=\sqrt{d}\\ A \sqrt{d} \left( 1 + \frac{\sigma^2}{\pi d} + \frac{\bar\beta}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} \mbox{ otherwise} \end{cases} \end{equation} Applying the Poissonian summation formula on the infinite sum, \begin{align} \hat{t}_c \ket{\Psi(k_0)} &= \frac{iT_0}{2\pi} \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m\in\mathbb{Z}} \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} \ket{E_n} + \ket{\epsilon_2} + \ket{\epsilon_1}. \end{align} Applying the Hamiltonian, \begin{align} \hat{H}_c \hat{t}_c \ket{\Psi(k_0)} &= i \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m\in\mathbb{Z}} n \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} \ket{E_n} + \hat{H}_c \left( \ket{\epsilon_2} + \ket{\epsilon_1} \right). \end{align} Finally, we shift back to the basis of time-states, \begin{align} \hat{H}_c \hat{t}_c \ket{\Psi(k_0)} &= i \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m\in\mathbb{Z}} n \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} \frac{e^{+i 2\pi nl/d}}{\sqrt{d}} \ket{\theta_l} + \hat{H}_c \left( \ket{\epsilon_2} + \ket{\epsilon_1} \right). \end{align} Combining the summations over the indices $n$ and $m$, and using $\sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m\in\mathbb{Z}} f(n+md) = \sum_{s\in\mathbb{Z}} f(s)$, and noting that \begin{align} \sum_{m \in \mathbb{Z}} \abs{ m \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} } = 2 \pi A \frac{\sigma}{\sqrt{d}} \sum_{m \in \mathbb{Z}} e^{-\frac{\pi \sigma^2}{d^2}(n-n_0+md)^2} \abs{m} \abs{ \frac{\sigma^2}{d^2} (n-n_0+md) + i \frac{k_0}{d} }, \end{align} we have \begin{align} \hat{H}_c \hat{t}_c \ket{\Psi(k_0)} &= i \sum_{l \in \mathcal{S}_d(k_0)} \sum_{s\in\mathbb{Z}} s \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=s} \frac{e^{+i 2\pi sl/d}}{\sqrt{d}} \ket{\theta_l} + \ket{\epsilon_3} + \hat{H}_c \left( \ket{\epsilon_2} + \ket{\epsilon_1} \right), \\ \text{where} \; \ket{\epsilon_3} &= -id \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m\in\mathbb{Z}} m \frac{d}{dp} \tilde{\psi}(k_0;p) \bigg|_{p=n+md} \frac{e^{+i 2\pi nl/d}}{\sqrt{d}} \ket{\theta_l}, \end{align} \begin{align}\label{eq:ep3 comm bound} \epsilon_3^{co} &:= \ltwo{\ket{\epsilon_3}} \\&< \begin{cases} 2\pi A d^2 \sqrt{d} \left( (1-\bar\alpha) + \frac{1}{\pi d} \left( 2 + \frac{1}{1 - e^{-\pi d (1-\bar\alpha)}} \right) + \frac{\bar\beta}{2} \left( 1 - \bar\alpha + \frac{1}{\pi d} + \frac{1+\bar\alpha}{1 - e^{-\pi d(1-\bar\alpha)}} \right) \right) e^{-\frac{\pi d}{4}(1-\bar\alpha)^2} \mbox{ if } \sigma=\sqrt{d}\\ 2\pi A \sigma d^2 \left( \frac{\sigma^2}{d} (1-\bar\alpha) + \frac{1}{\pi d} \left( 2 + \frac{1}{1 - e^{-\pi\sigma^2(1-\bar\alpha)}} \right) + \frac{\bar\beta}{2} \left( 1 - \bar\alpha + \frac{1}{\pi\sigma^2} + \frac{1+\bar\alpha}{1 - e^{-\pi\sigma^2(1-\bar\alpha)}} \right) \right) e^{-\frac{\pi\sigma^2}{4}(1-\bar\alpha)^2} \mbox{ otherwise} \end{cases} \end{align} Applying the Poissonian summation on the infinite sum w.r.t. $s$, \begin{align} \hat{H}_c \hat{t}_c \ket{\Psi(k_0)} &= -i \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m\in\mathbb{Z}} \frac{d}{dx} \left( x \psi(k_0;x) \right)\bigg|_{x=l+md} \ket{\theta_l} + \ket{\epsilon_3} + \hat{H}_c \left( \ket{\epsilon_2} + \ket{\epsilon_1} \right). \end{align} Approximating the sum over $m$ by the $m=0$ term, \begin{align} \hat{H}_c \hat{t}_c \ket{\Psi(k_0)} &= -i \sum_{l \in \mathcal{S}_d(k_0)} \psi(k_0;l) \ket{\theta_l} + i \sum_{l \in \mathcal{S}_d(k_0)} l \frac{d}{dx} \psi(k_0;x)\bigg|_{x=l} \ket{\theta_l} + \ket{\epsilon_4} + \ket{\epsilon_3} + \hat{H}_c \left( \ket{\epsilon_2} + \ket{\epsilon_1} \right), \\ \text{where} \; \ket{\epsilon_4} &= -i \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m\in\mathbb{Z}/\{0\}} \frac{d}{dx} \left( x\psi(k_0;x) \right)\bigg|_{x=l+md} \ket{\theta_l}, \end{align} \begin{equation}\label{eq:ep4 comm bound} \epsilon_4^{co}:= \ltwo{\ket{\epsilon_4}} < \begin{cases} dA \left( d (\pi+1) (1+\bar\beta) + 2 + \frac{2}{1 - e^{-\pi}} + \bar\bar\alpha \left( d (\pi+1) + \frac{\pi\bar\beta}{1 - e^{-\pi}} \right) \right) e^{-\frac{\pi d}{4}} \mbox{ if } \sigma=\sqrt{d}\\ dA \left( \left( \frac{\pi d^2}{\sigma^2} + d \right)(1+\bar\beta) + 2 + \frac{2}{1 - e^{-\frac{\pi d}{\sigma^2}}} + \bar\alpha \left( \pi d + \sigma^2 + \frac{\pi\bar\beta}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) \right) e^{-\frac{\pi d^2}{4\sigma^2}} \mbox{ otherwise} \end{cases} \end{equation} Consider now the alternate product $\hat{t}_c \hat{H}_c \ket{\Psi(k_0)}$. To begin with, we convert the state into the energy basis, \begin{align} \ket{\Psi(k_0)} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{k \in \mathcal{S}_d(k_0)} \psi(k_0;k) \frac{e^{-i2\pi nk/d}}{\sqrt{d}} \ket{E_n}. \end{align} Converting the summation over $k$ into the infinite sum and bounding the difference, \begin{align} \ket{\Psi(k_0)} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{k \in \mathbb{Z}} \psi(k_0;k) \frac{e^{-i2\pi nk/d}}{\sqrt{d}} \ket{E_n} + \ket{\epsilon_5}, \\ \text{where} \; \ket{\epsilon_5} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{k \in \mathbb{Z}/\mathcal{S}_d(k_0)} \psi(k_0;k) \frac{e^{-i2\pi nk/d}}{\sqrt{d}} \ket{E_n}, \end{align} \begin{equation}\label{eq:ep5 comm bound} \epsilon_5^{co}:= \ltwo{\ket{\epsilon_5}} < \begin{cases} 2\sqrt{d} A \frac{ e^{-\frac{\pi d}{4}}}{1 - e^{-\pi}} \mbox{ if } \sigma=\sqrt{d}\\ 2\sqrt{d} A \frac{ e^{-\frac{\pi d^2}{4\sigma^2}}}{1 - e^{-\frac{\pi d}{\sigma^2}}} \mbox{ otherwise} \end{cases} \end{equation} Applying the Poissonian summation formula on the infinite sum, \begin{align} \ket{\Psi(k_0)} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m \in \mathbb{Z}} \tilde{\psi}(k_0;n+md) \ket{E_n} + \ket{\epsilon_5}. \end{align} We now apply the Hamiltonian, \begin{align} \hat{H}_c \ket{\Psi(k_0)} &= \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m \in \mathbb{Z}} \frac{2\pi}{T_0} n \tilde{\psi}(k_0;n+md) \ket{E_n} + \hat{H}_c \ket{\epsilon_5}. \end{align} In anticipation of applying the time operator, we shift back to the basis of time-states, \begin{align} \hat{H}_c \ket{\Psi(k_0)} &= \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m \in \mathbb{Z}} \frac{2\pi}{T_0} n \tilde{\psi}(k_0;n+md) \frac{e^{+i2\pi nl/d}}{\sqrt{d}} \ket{\theta_l} + \hat{H}_c \ket{\epsilon_5}. \end{align} Combining the sums over $n$ and $m$, \begin{align} \hat{H}_c \ket{\Psi(k_0)} &= \sum_{l \in \mathcal{S}_d(k_0)} \sum_{s \in \mathbb{Z}} \frac{2\pi}{T_0} s \tilde{\psi}(k_0;s) \frac{e^{+i2\pi sl/d}}{\sqrt{d}} \ket{\theta_l} + \ket{\epsilon_6} + \hat{H}_c \ket{\epsilon_5}, \\ \text{where} \; \ket{\epsilon_6} &= -d \sum_{l \in \mathcal{S}_d(k_0)} \sum_{n=-\frac{d-1}{2}}^{+\frac{d-1}{2}} \sum_{m \in \mathbb{Z}} \frac{2\pi}{T_0} m \tilde{\psi}(k_0;n+md) \frac{e^{+i2\pi nl/d}}{\sqrt{d}} \ket{\theta_l}, \end{align} \begin{equation}\label{eq:ep6 comm bound} \epsilon_6^{co}:= \ltwo{\ket{\epsilon_6}} T_0 < \begin{cases} 2\pi A d^2 \sqrt{d} \left( \frac{1-\bar\alpha}{2} + \frac{1}{2\pi d} + \left( \frac{1+\bar\alpha}{2} \right) \frac{1}{1 - e^{-\pi d(1-\bar\alpha)}} \right) e^{-\frac{\pi d}{4}(1-\bar\alpha)^2} \mbox{ if } \sigma=\sqrt{d}\\ 2\pi A d^2 \sigma \left( \frac{1-\bar\alpha}{2} + \frac{1}{2\pi\sigma^2} + \left( \frac{1+\bar\alpha}{2} \right) \frac{1}{1 - e^{-\pi\sigma^2(1-\bar\alpha)}} \right) e^{-\frac{\pi\sigma^2}{4}(1-\bar\alpha)^2} \mbox{ otherwise} \end{cases} \end{equation} Applying the Poissonian summation formula, \begin{align} \hat{H}_c \ket{\Psi(k_0)} &= -\frac{id}{2\pi} \frac{2\pi}{T_0} \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m \in \mathbb{Z}} \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l+md} \ket{\theta_l} + \ket{\epsilon_6} + \hat{H}_c \ket{\epsilon_5}. \end{align} Approximating the sum over $m$ by the $m=0$ term, \begin{align} \hat{H}_c \ket{\Psi(k_0)} &= -\frac{id}{T_0} \sum_{l \in \mathcal{S}_d(k_0)} \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l} \ket{\theta_l} + \ket{\epsilon_7} + \ket{\epsilon_6} + \hat{H}_c \ket{\epsilon_5}, \\ \text{where} \; \ket{\epsilon_7} &= -\frac{id}{T_0} \sum_{l \in \mathcal{S}_d(k_0)} \sum_{m \in \mathbb{Z}/\{0\}} \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l+md} \ket{\theta_l}, \end{align} \begin{equation}\label{eq:ep7 comm bound} \epsilon_7^{co}:= \ltwo{\ket{\epsilon_7}}T_0 < \begin{cases} 2\pi d A \left( 1 + \frac{1}{\pi} + \frac{\bar\alpha}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} \mbox{ if } \sigma=\sqrt{d}\\ 2\pi d A \left( \frac{d}{\sigma^2} + \frac{1}{\pi} + \frac{\bar\alpha}{1 - e^{-\frac{\pi d}{\sigma^2}}} \right) e^{-\frac{\pi d^2}{4\sigma^2}} \mbox{ otherwise} \end{cases} \end{equation} We now apply the Time operator, \begin{align} \hat{t}_c \hat{H}_c \ket{\Psi(k_0)} &= -i \sumt{l} \sumS{k}{k_0} \frac{d}{dx} \psi(k_0;x) \bigg|_{x=k} \ket{\theta_l} + \hat{t}_c \left( \ket{\epsilon_7} + \ket{\epsilon_6} + \hat{H}_c \ket{\epsilon_5} \right). \end{align} As before, we split the sum case wise, we find \begin{align} \hat{t}_c \hat{H}_c \ket{\Psi(k_0)} &= -i \sum_{l \in \mathcal{S}_d(k_0)} l \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l} \ket{\theta_l} + \ket{\epsilon_8} + \hat{t}_c \left( \ket{\epsilon_7} + \ket{\epsilon_6} + \hat{H}_c \ket{\epsilon_5} \right), \\ \text{where} \; \ket{\epsilon_8} &= id \sum_{l \in \mathcal{S}_d(k_0),l > d/2} \frac{d}{dx} \psi(k_0;x) \bigg|_{x=l} \ket{\theta_l}, \end{align} \begin{equation}\label{eq:ep8 comm bound} \epsilon_8^{co}:= \ltwo{\ket{\epsilon_8}} < \begin{cases} \pi d A \left( 1 -\bar\beta + \frac{1}{\pi} + \frac{\bar\alpha}{1 - e^{-\pi(1-\bar\beta)}} \right) e^{-\frac{\pi d}{4}(1-\bar\beta)^2} \mbox{ if } \sigma=\sqrt{d}\\ \pi d A \left( \frac{d}{\sigma^2}(1-\bar\beta) + \frac{1}{\pi} + \frac{\bar\alpha}{1 - e^{-\frac{\pi d}{\sigma^2}(1-\bar\beta)}} \right) e^{-\frac{\pi d^2}{4\sigma^2}(1-\bar\beta)^2} \mbox{ otherwise} \end{cases} \end{equation} We may thus calculate the commutator on the state, \begin{align} \left[ \hat{t}_c, \hat{H}_c \right] \ket{\Psi(k_0)} &= i \ket{\Psi(k_0)} + \ket{\epsilon_\textup{comm}} , \\ \text{where} \quad \ket{\epsilon_\textup{comm}} &= \ket{\epsilon_8} + \hat{t}_c \left( \ket{\epsilon_7} + \ket{\epsilon_6} + \hat{H}_c \ket{\epsilon_5} \right) - \left( \ket{\epsilon_4} + \ket{\epsilon_3} + \hat{H}_c \left( \ket{\epsilon_2} + \ket{\epsilon_1} \right) \right). \end{align} The operator norms of $\hat{t}_c$ and $\hat{H}_c$ are bounded by $T_0/2$ and $\pi d/T_0$ respectively, and we can thus bound the norm of the total error, expressed here for the symmetric state only, \begin{align} \ltwo{\epsilon_\textup{comm}} &< 2 \pi d A \frac{e^{-\frac{\pi d}{4} (1-\beta)^2}}{1 - e^{-\pi(1-\beta)}} + \pi d^\frac{3}{2} A \left( 1 + \frac{1}{\pi} + \frac{\bar\beta}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}}\\ &+ 2\pi A d^\frac{5}{2} \left( (1-\bar\alpha) + \frac{1}{\pi d} \left( 2 + \frac{1}{1 - e^{-\pi d (1-\bar\alpha)}} \right) + \frac{\bar\beta}{2} \left( 1 - \bar\alpha + \frac{1}{\pi d} + \frac{1+\bar\alpha}{1 - e^{-\pi d(1-\bar\alpha)}} \right) \right) e^{-\frac{\pi d}{4}(1-\bar\alpha)^2} \\ &+ dA \left( d (\pi+1) (1+\bar\beta) + 2 + \frac{2}{1 - e^{-\pi}} + \bar\alpha \left( d (\pi+1) + \frac{\pi\bar\beta}{1 - e^{-\pi}} \right) \right) e^{-\frac{\pi d}{4}}\\ &+ \pi d^\frac{3}{2} A \frac{ e^{-\frac{\pi d}{4}}}{1 - e^{-\pi}} + \pi d^\frac{5}{2} A \left( \frac{1-\bar\alpha}{2} + \frac{1}{2\pi d} + \left( \frac{1+\bar\alpha}{2} \right) \frac{1}{1 - e^{-\pi d(1-\bar\alpha)}} \right) e^{-\frac{\pi d}{4}(1-\bar\alpha)^2} \\ &+ \pi d A \left( 1 + \frac{1}{\pi} + \frac{\bar\alpha}{1 - e^{-\pi}} \right) e^{-\frac{\pi d}{4}} + \pi d A \left( 1 -\bar\beta + \frac{1}{\pi} + \frac{\bar\alpha}{1 - e^{-\pi(1-\bar\beta)}} \right) e^{-\frac{\pi d}{4}(1-\bar\beta)^2}. \end{align} For the completely symmetric state, \begin{align} \ltwo{\ket{\epsilon_{total}}} \cdot e^{+\frac{\pi d}{4}} &< 2 \pi d A \frac{1}{1 - e^{-\pi}} + \pi d^\frac{3}{2} A \left( 1 + \frac{1}{\pi} \right) + 2\pi A d^\frac{5}{2} \left( 1 + \frac{1}{\pi d} \left( 2 + \frac{1}{1 - e^{-\pi}} \right) \right) \\ &+ dA \left( d (\pi+1) + 2 + \frac{2}{1 - e^{-\pi}} \right) + \pi d^\frac{3}{2} A \frac{1}{1 - e^{-\pi}} + \pi d^\frac{5}{2} A \left( \frac{1}{2} + \frac{1}{2\pi d} + \left( \frac{1}{2} \right) \frac{1}{1 - e^{-\pi}} \right) \\ &+ 2\pi d A \left( 1 + \frac{1}{\pi} \right), \end{align} which can be simplified to \begin{align} \ltwo{\ket{\epsilon_\textup{comm}}} &= \mathcal{O}(d^{9/4})\, e^{-\frac{\pi d}{4}}, \end{align} (recall that $A$ is of the order $d^{-\frac{1}{4}}$). \end{proof} \section{Conjectures and open questions}\label{sec:conjectures} In this section we state a conjecture about the tightness of our results. This is based on numerical studies and some intuition. We will also discuss some open questions about the properties of the bounds.\\ \textit{Conjecture 1: tightness in exponential decay of clock quasi-continuity}. Based on numerical studies, we conjecture that the norm of $\ket{\epsilon}$ in Theorem \ref{gaussiancontinuity}, does not decay super-exponentially for all $t\in[0,T_0]$. In other words, there does \textit{not} exist a $\ket{\Psi_\textup{nor}(k_0)}\in\Lambda_{\sigma,n_0}$ and an $f: {\mathbbm{R}} \rightarrow {\mathbbm{R}}^+$, with $\lim_{d\rightarrow \infty} d/f(d)=0$ such that \begin{equation} \lim_{d\rightarrow \infty} \|\ket{\epsilon}\|_2 \,\mathrm{e}^{f(d)}= 0 \quad \forall t\in[0,T_0], \end{equation} where $T_0>0$ and $\ket{\epsilon}$ is defined in Theorem \ref{gaussiancontinuity}.\\ If the conjecture is indeed correct, it would prove the tightness of the exponential decay bound for symmetric states, the next related interesting questions would be whether this bound is only achievable for symmetric states and whether the factor of $\pi/4$ is optimal for completely symmetric states. Furthermore, since the family of Gaussian clock states $\ket{\Psi_\textup{nor}(k_0)}\in\Lambda_{\sigma,n_0}$ are minimum uncertainty states\footnote{Up to an exponentially decaying correction in clock dimension.}, we propose that out of all normalized states $\ket{\psi}$ in $\mathcal{H}_c$ satisfying $[\hat t_c,\hat H_c]\ket{\psi}=\mathrm{i}\ket{\psi}+\ket{\epsilon}$; their corresponding error $\|\ket{\epsilon}\|_2$ will decay, at most, exponentially fast in $d$. As such, we expect that the exponential decay in clock dimension is a fundamental limitation of finite dimensional clocks.\\ There are also related interesting questions concerning the continuous quasi-control too. For example, we know that after one period, $t=T_0$, when the potential is zero, $V_0(x)=0,\;\;\forall x\in{\mathbbm{R}}$, that the clock is returned to its initial state $\ket{\bar\Psi_\textup{nor}(k_0,\Delta)}$. We know that whenever $V_0(x)\neq 0$ that this is not true. While Theorem \ref{movig through finite time} upper bounds how small this error is, what is not so clear is how quickly the error $\|\ket{\epsilon}\|_2$ in Theorem \ref{movig through finite time} goes to zero in the limit that $\| V_0\|_2\rightarrow 0$. Such scaling is numerically challenging to estimate, and we thus do not propose an answer. e.g. it would be interesting to know whether it is power law decay or exponential decay in $1/\|V_0\|_2$.
38a7e1cf829f336f0a6ea936289037097c9b40c4
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In high energy scatterings, gluon bremsstrahlung radiation plays a crucial role in describing the rapid rise of gluon density inside hadron and nucleus. The soft gluon radiation is ingeniously encoded in the Balitsky-Fadin-Kuraev-Lipatov (BFKL) evolution equation~\cite{Balitsky:1978ic,Lipatov:1976zz,Kuraev:1976ge,Fadin:1975cb} which resums small-$x$ logarithms. As a result, the energy evolution of off-shell unintegrated parton distribution function in dilute limit is governed by the BFKL equation in contrast to the well known Dokshitzer-Gribov-Lipatov-Altarelli-Parisi (DGLAP) evolution for collinear parton densities~\cite{Gribov:1972ri,Altarelli:1977zs,Dokshitzer:1977sg}. In phenomenology, to describe particle productions in hadronic collisions in the small $x$ and low transverse momentum transfer region, we need to take into account the off-shellness of incident partons~\cite{Collins:1991ty,Catani:1990eg}, and switch from collinear factorization framework to $k_t$-factorization framework (e.g., Ref.~\cite{Kharzeev:2003wz,Kharzeev:2004yx,Blaizot:2004wu}). Study of inclusive particle productions in high energy hadron-nucleus/nucleus-nucleus collisions has attracted a lot of attention in the past few years, since this process can help to reveal nonlinear gluon dynamics on top of the linear BFKL evolution when gluon density inside large nucleus is sufficiently high. As collision energy increases, the gluon occupation number in the low-$x$ region inside nucleus becomes large due to gluon bremsstrahlung radiation. The subsequent balance between the gluon bremsstrahlung and recombination leads to the phenomenon of gluon saturation~\cite{Gribov:1984tu,Mueller:1985wy, Mueller:2001fv} or color-glass-condensate~\cite{Weigert:2005us,Gelis:2010nm,Kovchegov:2012mbw}. In principle, the so-called saturation scale $Q_s$ separates the nonlinear dense regime $Q\leq Q_s$ from dilute regime $Q\gg Q_s$ with $Q$ being the typical scale of the external probe. Since $Q_s$ is enhanced by a factor $A^{1/3}$ for heavy nuclei with $A$ being the number of nucleon, current proton-nucleus (p$A$) collision and a future Electron-Ion-Collider are expected to provide unique opportunities to investigate the gluon saturation phenomenon. In particular, the measurement of the nuclear modification factor $R_{{\rm p}A}\equiv\frac{1}{A}\frac{d\sigma_{{\rm p}A}/d^2p_{\perp}dy}{d\sigma_{{\rm pp}}/d^2p_{\perp}dy}$ at RHIC and the LHC~\cite{Arsene:2004ux,Adams:2003qm,ALICE:2012mj} reflects the relatively different strength of the saturation phenomenon in pp and p$A$ collisions and it has been described by various theoretical calculations~\cite{Kharzeev:2004yx,Albacete:2010bs,Albacete:2012xq,Lappi:2013zma,JalilianMarian:2011dt,Albacete:2014fwa} in small-$x$ physics. Furthermore, inclusive single hadron production in p$A$ collisions (${\rm p}+A\rightarrow h+X$) has been studied in the small-$x$ formalism based on the Mueller's dipole model~\cite{Mueller:1993rr,Mueller:1994gb}. In particular, at forward rapidity, the longitudinal momentum fraction of target nucleus carried by incoming parton becomes small and the saturation scale for heavy nuclei is much larger than typical hadronic scale $\Lambda_\text{QCD}$, whereas the saturation scale of the proton projectile is expected to be much smaller. Therefore, as long as we focus on the forward hadron production, the so-called hybrid treatment, in which the collinear parton coming from projectile proton scatters off dense gluons inside target nucleus with finite transverse momentum transfer, is considered to be a reasonable and simple approach~\cite{Dumitru:2002qt,Dumitru:2005kb}. In the hybrid formalism, hadron productions in p$A$ collisions are given by the convolution of short distance hard parts, collinear parton distribution functions (PDFs) for projectile proton, fragmentation functions (FFs) for produced hadron, and the wave function of target nucleus constructed from multi-point Wilson line correlators. In general, multi-point Wilson line correlators obey the so-called JIMWLK evolution equation~\cite{Jalilian-Marian:1997jx+X}. In phenomenological studies, large-$N_c$ and mean field approximations are usually utilized to reduce multi-point Wilson line correlators to products of dipole correlators which resum the small-$x$ logarithm $\alpha_s\ln1/x$ by means of the Balitsky-Kovchegov (BK) equation~\cite{Balitsky:1995ub,Kovchegov:1996ty}. In addition, inclusive hadron production in p$A$ collisions beyond leading order (LO) has been firstly studied in Refs.~\cite{Albacete:2010bs,JalilianMarian:2011dt,Altinoluk:2011qy,Fujii:2011fh} where the running coupling BK equation is employed together with the LO framework. These studies included a subset of next-to-leading order (NLO) corrections. The pursuit of complete NLO calculations began several years ago~\cite{Chirilli:2011km,Chirilli:2012jd,Stasto:2013cha,Altinoluk:2014eka,Watanabe:2015tja,Albacete:2014fwa,Ducloue:2016shw}. In Refs.~\cite{Chirilli:2011km,Chirilli:2012jd}, the systematic NLO calculations is carried out for the first time at the high energy limit. It is shown that the collinear divergences and the rapidity divergences, which arise from one-loop diagrams at NLO, are clearly separable from the partonic hard scattering part in the small-$x$ formalism for inclusive hadron production in p$A$ collisions. This allows us to establish an effective factorization in the hybrid formalism at the one-loop order. The small-$x$ factorization formula for inclusive hadron production at forward rapidity $y$ with transverse momentum $p_{\perp}$ in p$A$ collisions can be cast into~\cite{Chirilli:2012jd} \begin{align} \frac{d\sigma^{{\rm p}+A\rightarrow h+X}}{d^2p_{\perp}dy} =\sum_{i=q,g}\int \frac{dz}{z^2}x_pf_i\left(x_p,\mu\right){\cal F}^{F,A}_{x_g}(k_\perp)D_{h/i}(z,\mu)+\sum_{i,j=q,g}\frac{\alpha_s}{2\pi}\int\frac{dz}{z^2}\frac{dx}{x}\xi xf_i\left(x,\mu\right)S_{ij}D_{h/j}(z,\mu) \label{eq:factorization} \end{align} where $x_p=k_\perp e^y/\sqrt{s}$ is longitudinal momentum fraction of projectile proton carried by incoming parton, $k_\perp$ is transverse momentum of gluon coming from target nucleus with longitudinal momentum $x_g=k_\perp e^{-y}/\sqrt{s}$, and $\xi=x_p/x$. In addition, $z$ is momentum fraction of parton carried by produced hadron in final state. $f_i(x)$ is the collinear PDF of projectile proton and $D_{h/i}(z)$ is the FF from parton $i$ to hadron $h$. ${\cal F}^{F,A}_{x_g}$ is Fourier transform of the dipole amplitude in the fundamental or adjoint representation which is determined by the partonic scattering at LO. This dipole amplitude encodes all the important information on the scattering between incoming partons and target harons. $S_{ij}$ represent the partonic scattering processes at NLO convoluted with the dipole amplitude. The factorization scale $\mu$ dependence of the PDFs and the FFs is derived from the DGLAP equation due to quantum evolution, while the rapidity or energy dependence of the dipole amplitude is described by the BK equation. In Ref.~\cite{Watanabe:2015tja}, the improved numerical calculation for inclusive hadron production in p$A$ collisions at NLO is performed and it yields the inclusive hadron spectra which are in excellent agreement with both RHIC and the LHC data for various kinematical regions. Nevertheless, it is still hard to tell whether this agreement is due to the non-linear gluon saturation phenomenon or the linear BFKL dynamics, since we do not know for sure whether the saturation effect is indispensable to describe the current data in pp/p$A$ collisions or not. In order to disentangle these two effects, it is very useful to conduct a comparative NLO calculation for single hadron production in pp collisions with only linear BFKL dynamics, which would help us clearly visualize the different behaviour of linear and non-linear dynamics in the hadron $p_\perp$ spectra and eventually lead us to better understanding of the onset of gluon saturation phenomenologically. The objective of this paper is to calculate the NLO forward inclusive hadron production in pp collisions in the small-$x$ formalism in the dilute regime. As a cross check, we find that the results that we obtain for pp collisions agree with the previous calculation for p$A$ collisions (assuming large $N_c$ approximation)~\cite{Chirilli:2012jd, Watanabe:2015tja} in the large $N_c$ and dilute limit. Through this paper, we use the light cone perturbation theory~\cite{Lepage:1980fj} together with the light-cone gauge $A^+=0$ and light-cone coordinates $x^{\pm}=(x^0\pm x^3)/{\sqrt 2}$ and $x^{\mu}=(x^+,x^-,x_\perp)$ with the metric $g_{+-}=g_{-+}=1$, $g_{11}=g_{22}=-1$. The rest of this paper is organized as follows: In Sec.~II, we firstly consider inclusive hadron production in pp collisions at LO within the hybrid formalism. Next, we calculate four NLO channels in Sec.~III. We show that the rapidity divergences and the collinear divergences can be completely factorized from the partonic hard scattering part like Eq.~(\ref{eq:factorization}). We see that the factorization scale dependence of the PDF and the FF is controlled by the DGLAP equation and the energy dependence of the dipole amplitude is described by the BFKL equation. The main result of this paper is given in Eq.~(\ref{eq:master-expression}). Finally, in Sec.~IV, we summarize our calculation and discuss its future development. \section{The Leading Order} \begin{figure} \centering \includegraphics[width=12cm,angle=0]{LO.eps} \caption{Lowest order diagrams for $q+{\rm p}\rightarrow q+X$ (left) and $g+{\rm p}\rightarrow g+X$ (right) with one gluon exchange in the $t$-channel. } \label{fig:LO-diagram} \end{figure} In this section, let us first consider the LO forward single hadron production in pp collisions. For pp collision, the center of mass energy is $s=(P+P^\prime)^2\simeq P^+P^{\prime-}$ with $P$ ($P^\prime$) being momentum of projectile (target) proton. Here the projectile (target) proton beam is supposed to have large light cone plus (minus) component and then $P^-=P^{\prime+}\simeq 0$ and $P_\perp=P^{\prime}_\perp\simeq 0$. We begin with a basic process of quark scattering off proton ($q+{\rm p}\rightarrow q+X$). In the small-$x$ formalism, the differential cross section for producing a quark with momentum $k=(k^+,k_\perp)$ is given by \begin{align} \frac{d\sigma^{q+{\rm p}\rightarrow q+X}_{\rm LO}}{d^3k}=\delta(p^+-k^+)\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-ik_\perp\cdot(x_\perp-y_\perp)}S_{x_g}(x_\perp,y_\perp) \end{align} where we have averaged over spin and color of the incoming quark and summed all of the quantum number of the final state, $p^+$ is the light cone momentum of the incoming quark and $k^+$ is the momentum of the observed quark. The transverse momentum of incident parton is assumed to be $p_\perp=0_\perp$ for the sake of simplicity. The delta function is due to the momentum conservation. $S_{x_g}$ is the color singlet dipole scattering amplitude in the fundamental representation \begin{align} S_{x_g}(x_\perp,y_\perp)=\frac{1}{N_c}\langle {\rm Tr}\left[U(x_\perp)U^\dagger(y_\perp)\right]\rangle_{x_g} \end{align} with $r_\perp=x_\perp-y_\perp$ being a transverse size of the dipole (see FIG.~\ref{fig:LO-diagram}). $U(x_\perp)$ is the fundamental Wilson line at transverse coordinate $x_\perp$ which resums multiple scatterings between the quark and the gluon of target hadron in the eikonal approximation. In dilute regime, one expects \begin{align} U(x_\perp)={\cal P}\exp\left[ig\int dx^+ t^aA_a^-(x^+,x_\perp)\right] \approx 1+ig\int dx^+ t^aA_a^-(x^+,x_\perp), \end{align} where $A(x^+,x_\perp)$ is the effective background gauge field of the target proton at $x=(x^+,x_\perp)$ and $t^a$ is the generator of SU(3) group. $x_g=k^-/P^{\prime-}$ is the longitudinal momentum fraction of the target proton carried by the small-$x$ gluon attached to the quark. $\langle\rangle_{x_g}$ represents the average over all of the color configurations of the gluon inside the target proton. To obtain the cross section for ${\rm p}+{\rm p}\rightarrow q+X$, we can couple the above results with collinear quark PDF $q_f$ with flavor $f$ and write \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow q+X}_{\rm LO}}{d^3k}=\sum_f\int dx q_f(x)\frac{d\sigma^{q+{\rm p}\rightarrow q+X}_{\rm LO}}{d^3k} \end{align} with $x_p=p^+/P^+$ being the longitudinal momentum fraction of the projectile proton carried by the incoming parton. Furthermore, together with the collinear FF which converts $q$ into hadron $h$ in the final state, the differential cross section of single hadron production with $p_{h\perp}$ at rapidity $y$ is given by \begin{align} \frac{d\sigma_{\rm LO}^{{\rm p}+{\rm p}\rightarrow h/q+X}}{d^2p_{h\perp} dy}=\sum_f\int_\tau^1\frac{dz}{z^2}D_{h/q}(z)x_pq_f\left(x_p\right){\cal F}_{x_g}(k_{\perp}) \label{eq:lo-q-momentum} \end{align} with $zk_\perp=p_{h\perp}$, $\tau=zx_p$, $x_p=k_\perp e^{+y}/\sqrt{s}$, and $x_g=k_\perp e^{-y}/\sqrt{s}$. ${\cal F}_{x_g}$ is the Fourier transform of the dipole amplitude in the fundamental representation defined as \begin{align} {\cal F}_{x_g}(k_{\perp}) \equiv\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-ik_\perp\cdot(x_\perp-y_\perp)}(-T_{x_g}(x_\perp,y_\perp))=S_\perp F_{x_g}(k_\perp), \label{fundamental-amplitude} \end{align} where $F_{x_g}(k_{\perp})=\int\frac{d^2r_\perp}{(2\pi)^2}e^{-ik_\perp\cdot r_\perp}(-T_{x_g}(r_\perp))$ with $T_{x_g}=1-S_{x_g}$ being the forward scattering amplitude. We can write the cross section which is proportional to the transverse area of the target proton $S_\perp$ if the impact parameter dependence is neglected. We note that we have dropped the elastic part which is proportional to $\delta^{(2)}(k_\perp)$, since we are only interested in the inelastic production with finite $k_\perp$. Similarly, for the gluon channel at LO, one finds \begin{align} \frac{d\sigma_{\rm LO}^{{\rm p}+{\rm p}\rightarrow h/g+X}}{d^2p_{h\perp} dy}=\int_\tau^1\frac{dz}{z^2}D_{h/g}(z)x_pG\left(x_p\right)\widetilde{{\cal F}}_{x_g}(k_{\perp}) \label{eq:lo-g-momentum} \end{align} where $G$ and $D_{h/g}$ are the collinear gluon PDF and FF, respectively. The Fourier transform of the forward scattering amplitude in the adjoint representation is \begin{align} \widetilde{{\cal F}}_{x_g}(k_{\perp}) =\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-ik_\perp\cdot(x_\perp-y_\perp)}(-\widetilde{T}_{x_g}(x_\perp,y_\perp)). \end{align} with \begin{align} \widetilde{S}_{x_g}(x_\perp, y_\perp)=1-\widetilde{T}_{x_g}(x_\perp, y_\perp)=\frac{1}{N_c^2-1}\langle {\rm Tr}\left[W(x_\perp)W^\dagger(y_\perp)\right]\rangle_{x_g} \end{align} where $W(x_\perp)$ is the adjoint Wilson line defined at $x_\perp$. Using the identity \begin{align} W^{ab}(x_\perp)=2{\rm Tr}\left[t^aU(x_\perp)t^bU^\dagger(x_\perp)\right], \end{align} we can approximate the adjoint dipole amplitude in the dilute regime of target proton as \begin{align} \widetilde{S}_{x_g}(x_\perp, y_\perp)&=\frac{1}{N_c^2-1}\left[\langle{\rm Tr}\left[U(x_\perp)U^{\dagger}(y_\perp)\right]{\rm Tr}\left[U(y_\perp)U^{\dagger}(x_\perp)\right]\rangle_{x_g}-1\right]\nonumber\\ &\approx1-\frac{N_c}{C_F}T_{x_g}(x_\perp,y_\perp). \end{align} Hereafter, we will use $\widetilde{T}_{x_g}\equiv N_cT_{x_g}/C_F$ in the following caculation. \section{The Next to Leading Order} Let us now consider the corresponding NLO calculation which can be put into four different channels. Essentially, the calculations which we perform below are similar to the calculations in Ref.~\cite{Chirilli:2012jd}. However, we would like to point out that there are some subtleties in order to obtain the differential cross section for inclusive hadron production in pp collisions at NLO at finite-$N_c$. In fact, it seems that the following calculation of this process in pp collisions in the dilute regime is slightly harder than the calculation which leads to the non-linear results obtained in p$A$ collisions, since the large $N_c$ approximation was employed in the latter case. \subsection{The $q\rightarrow q$ channel} \begin{figure} \centering \includegraphics[width=16cm,angle=0]{qq-real.eps} \caption{Real diagrams at NLO for the $q\rightarrow q$ channel. Notations are the same as FIG.~\ref{fig:LO-diagram}. } \label{fig:NLO-qq} \end{figure} For $q+{\rm p}\rightarrow q+g+X$ channel as shown in FIG.~\ref{fig:NLO-qq}, the differential cross section for producing a quark with momentum $k$ and a gluon with $l$ is given by \begin{align} \frac{d\sigma^{q+{\rm p}\rightarrow q+g+X}}{d^3kd^3l} =&\alpha_s\delta(p^+-k^+-l^+)\int\frac{d^2x_\perp d^2x_\perp^\prime d^2b_\perp d^2b_\perp^\prime}{(2\pi)^8} e^{-ik_{\perp}\cdot(x_\perp-x_\perp^\prime)}e^{-il_{\perp}\cdot(b_\perp-b_\perp^\prime)}\sum_{\alpha\beta\lambda}\psi_{qg\alpha\beta}^{\lambda\ast}(u_\perp^\prime)\psi_{qg\alpha\beta}^{\lambda}(u_\perp)\nonumber\\ &\times \Bigg[C_FS_{x_g}(x_\perp,x_\perp^\prime)+C_FS_{x_g}(v_\perp,v_\perp^\prime)+\frac{1}{2N_c}S_{x_g}(x_\perp,v_\perp^\prime)+\frac{1}{2N_c}S_{x_g}(v_\perp,x_\perp^\prime)\nonumber\\ &-\frac{N_c}{2}S_{x_g}(v_\perp,b_\perp)-\frac{N_c}{2}S_{x_g}(b_\perp,v_\perp^\prime)-\frac{N_c}{2}S_{x_g}(x_\perp,b_\perp)-\frac{N_c}{2}S_{x_g}(b_\perp,x_\perp^\prime)+N_cS_{x_g}(b_\perp,b_\perp^\prime)\Bigg] \end{align} where $u_\perp=b_\perp-x_\perp$ and $u_\perp^\prime=b_\perp^\prime-x_\perp^\prime$ are the transverse distance between the produced gluon and the quark in the amplitude and the complex conjugate amplitude, respectively. $v_\perp=\xi x_\perp+(1-\xi)b_\perp$ and $v_\perp^\prime=\xi x_\perp^\prime+(1-\xi)b_\perp^\prime$ are the transverse coordinate of the incoming quark in the amplitude and the complex conjugate amplitude, respectively. $\xi=k^+/p^+$ is the longitudinal momentum fraction of the incoming quark carried by the produced quark in the final state. $\psi_{qg\alpha\beta}^{\lambda}$ is the light cone wave function which describes the quark-gluon splitting amplitude~\cite{Marquet:2007vb} \begin{align} \psi_{qg\alpha\beta}^\lambda(u_\perp)=2\pi i\sqrt{\frac{2}{(1-\xi)p^+}} \left\{ \begin{array}{c} \frac{u_\perp\cdot \varepsilon^{(1)}_\perp}{u_\perp^2}(\delta_{\alpha-\beta-}+\xi\delta_{\alpha+}\delta_{\beta+})~~~(\lambda=1)\\ \frac{u_\perp\cdot \varepsilon^{(2)}_\perp}{u_\perp^2}(\delta_{\alpha+\beta+}+\xi\delta_{\alpha-}\delta_{\beta-})~~~(\lambda=2), \end{array} \right. \end{align} where $\alpha$ and $\beta$ are a spin of the incoming quark and the outgoing quark, respectively. $\lambda$ is a polarization of the radiated gluon. The polarization vector is defined as ${\varepsilon_\perp^{(\lambda)}}=-\frac{1}{\sqrt2}(\sigma, i)$ with $\sigma=+1$ for $\lambda=1$ and $\sigma=-1$ for $\lambda=2$, respectively. Summing over $\alpha$, $\beta$, and $\lambda$, one finds \begin{align} \sum_{\alpha\beta\lambda}\psi_{qg\alpha\beta}^{\lambda\ast}(u_\perp^\prime)\psi_{qg\alpha\beta}^{\lambda}(u_\perp) =\frac{2(2\pi)^2}{p^+}\frac{1+\xi^2}{1-\xi}\frac{u_\perp^\prime\cdot u_\perp}{u_\perp^{\prime2}u_{\perp}^2}. \label{eq:splitting-kernel-qq} \end{align} Summing over all the real diagrams, the total real contribution for inclusive hadron production reads \begin{align} \frac{\alpha_s}{2\pi^2}&\int^1_\tau\frac{dz}{z^2}D_{h/q}(z)\int d^2l_{\perp}\int^1_{\frac{\tau}{z}}d\xi \frac{1+\xi^2}{1-\xi}\frac{x_p}{\xi}q_f\left(\frac{x_p}{\xi}\right)\int\frac{d^2x_\perp d^2x_\perp^\prime d^2b_\perp d^2b_\perp^\prime}{(2\pi)^4} e^{-ik_{\perp}\cdot(x_\perp-x_\perp^\prime)}e^{-il_{\perp}\cdot(b_\perp-b_\perp^\prime)}\frac{u_\perp^\prime\cdot u_\perp}{u_\perp^{\prime2}u_\perp^2}\nonumber\\ \times \Bigg[&-C_FT_{x_g}(x_\perp,x_\perp^\prime)-C_FT_{x_g}(v_\perp,v_\perp^\prime)+\frac{N_c}{2}T_{x_g}(v_\perp,b_\perp)+\frac{N_c}{2}T_{x_g}(b_\perp,v_\perp^\prime)+\frac{N_c}{2}T_{x_g}(x_\perp,b_\perp)+\frac{N_c}{2}T_{x_g}(b_\perp,x_\perp^\prime)\nonumber\\ &-\frac{1}{2N_c}T_{x_g}(x_\perp,v_\perp^\prime)-\frac{1}{2N_c}T_{x_g}(v_\perp,x_\perp^\prime)-N_cT_{x_g}(b_\perp,b_\perp^\prime)\Bigg], \label{eq:qq-real2} \end{align} where we have not yet integrated over $l_{\perp}$ since the upper limit of $\xi$ actually depends on $l_\perp$, as we will consider below. We note here that FIG.~\ref{fig:NLO-qq} $(i)$ eventually does not contribute the inelastic single hadron production thank to the unitarity constraint~\cite{Mueller:2001fv}. \begin{figure} \centering \includegraphics[width=18cm,angle=0]{qq-virtual.eps} \caption{Virtual diagrams at NLO for the $q\rightarrow q$ channel. Mirror contributions are omitted. } \label{fig:NLO-qq-virtual} \end{figure} In the meantime, the virtual diagrams shown in FIG.~\ref{fig:NLO-qq-virtual} yield the following contribution \begin{align} -2\frac{\alpha_s}{2\pi^2}&\int^1_\tau\frac{dz}{z^2}D_{h/q}(z)x_pq_f(x_p)\int_0^1d\xi\frac{1+\xi^2}{1-\xi}\int\frac{d^2v_\perp d^2v_\perp^\prime d^2u_\perp}{(2\pi^)}e^{-ik_{\perp}\cdot(v_\perp-v_\perp^\prime)}\frac{1}{u_\perp^2}\nonumber\\ &\times\Bigg[-C_FT_{x_g}(v_\perp,v_\perp^\prime)-\frac{1}{2N_c}T_{x_g}(x_\perp,v_\perp^\prime)+\frac{N_c}{2}T_{x_g}(b_\perp,v_\perp^\prime)+\frac{N_c}{2}T_{x_g}(b_\perp,x_\perp)\Bigg] \label{eq:qq-virtual} \end{align} where we multiplied a factor of 2 explicitly to reflect the mirror diagrams of FIG.~\ref{fig:NLO-qq-virtual}. We would like to emphasize that an unfamiliar contribution from FIG.~\ref{fig:NLO-qq-virtual} $(d)$ plays an important role at $\xi\rightarrow 1$ due to the unitarity relation~\cite{Mueller:2001fv}, albeit it does not contribute to the inelastic hadron production as we explain below. It is manifest that the above results have the so-called rapidity singularities at $\xi=1$ in which the radiated gluon becomes soft. To deal with the rapidity singularities we adopt the plus function which is defined as follows \begin{align} \int_\tau^1d\xi\frac{1+\xi^2}{1-\xi}F(\xi)=\int_\tau^1d\xi\frac{1+\xi^2}{(1-\xi)_+}F(\xi)+\int_0^1d\xi\frac{2}{1-\xi}F(1) \label{eq:plus-function} \end{align} for arbitrary function $F$. Making use of Eq.~(\ref{eq:plus-function}) and the identity \begin{align} \frac{x_\perp}{x_\perp^2}=\int\frac{d^2k_\perp}{2\pi i}\frac{k_\perp}{k_\perp^2}e^{ik_\perp\cdot x_\perp}, \end{align} the nonsingular part in the momentum space can be cast into \begin{align} &\frac{\alpha_s}{2\pi^2}\int_\tau^1\frac{dz}{z^2}D_{h/q}(z)\Bigg\{\int^1_{\frac{\tau}{z}} d\xi\frac{1+\xi^2}{(1-\xi)_+}\frac{x_p}{\xi}q_f\left(\frac{x_p}{\xi}\right)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\Bigg[C_F\frac{1}{(k_{\perp}-k_{g\perp})^2}+C_F\frac{1}{(k_{\perp}-\xi k_{g\perp})^2}\nonumber\\ &-N_c\frac{k_\perp\cdot(k_\perp-k_{g\perp})}{k_\perp^2(k_\perp-k_{g\perp})^2}-N_c\frac{k_\perp\cdot(k_\perp-\xi k_{g\perp})}{k_\perp^2(k_\perp-\xi k_{g\perp})^2} +\frac{1}{N_c}\frac{(k_{\perp}-k_{g\perp})\cdot(k_{\perp}-\xi k_{g\perp})}{(k_{\perp}-k_{g\perp})^2(k_{\perp}-\xi k_{g\perp})^2}\Bigg]\nonumber\\ &-\int^1_{0} d\xi\frac{1+\xi^2}{(1-\xi)_+}x_pq_f(x_p){\cal F}_{x_g}(k_{\perp})\int d^2k_{g\perp}\left[2C_F\frac{1}{k_{g\perp}^2}-N_c\frac{k_{g\perp}\cdot(k_{g\perp}+\xi k_\perp)}{k_{g\perp}^2(k_{g\perp}+\xi k_\perp)^2}+\frac{1}{N_c}\frac{k_{g\perp}\cdot(k_{g\perp}-(1-\xi) k_\perp)}{k_{g\perp}^2(k_{g\perp}-(1-\xi) k_\perp)^2}\right]\Bigg\}, \label{eq:qq-real+virtual} \end{align} where the phase space of the radiated gluon has been integrated out. Here we can take the $s\rightarrow\infty$ limit safely. It is straightforward to evaluate most of the terms in the above expression in Eq.~(\ref{eq:qq-real+virtual}) except for the term which is proportional to $\ln(1-\xi)^2$. Following Ref.~\cite{Chirilli:2012jd}, we can make use of the following identity \begin{align} \int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\frac{(k_{\perp}-k_{g\perp})\cdot(k_{\perp}-\xi k_{g\perp})}{(k_{\perp}-k_{g\perp})^2(k_{\perp}-\xi k_{g\perp})^2} =-\pi {\cal F}_{x_g}(k_\perp)\ln(1-\xi)^2+\pi \overline{I}^{(1)}_{qq} \end{align} where $\overline{I}^{(1)}_{qq}$ is defined as \begin{align} \overline{I}^{(1)}_{qq}&=\int\frac{d^2k_{g\perp}}{\pi}\left[{\cal F}_{x_g}(k_{g\perp})\frac{(k_\perp-k_{g\perp})\cdot(k_{\perp}-\xi k_{g\perp})}{(k_\perp-k_{g\perp})^2(k_{\perp}-\xi k_{g\perp})^2} -{\cal F}_{x_g}(k_{\perp})\left\{\frac{(k_\perp-k_{g\perp})\cdot(\xi k_{\perp}-k_{g\perp})}{(k_\perp-k_{g\perp})^2(\xi k_{\perp}-k_{g\perp})^2} +\frac{k_{g\perp}\cdot(k_{\perp}-k_{g\perp})}{k_{g\perp}^2(k_{\perp}-k_{g\perp})^2}\right\}\right]. \label{eq:Iqq} \end{align} Meanwhile, the virtual contribution also contains a similar term in proportion to $\ln(1-\xi)^2$. Then, by combining the real contributions and the virtual contributions together, one finds \begin{align} \int^1_{\frac{\tau}{z}}d\xi\frac{1+\xi^2}{(1-\xi)_+}\frac{x_p}{\xi}q_f\left(\frac{x_p}{\xi}\right)\ln(1-\xi)^2 -\int^1_{0}d\xi\frac{1+\xi^2}{(1-\xi)_+}x_pq_f(x_p)\ln(1-\xi)^2\nonumber\\ =\int^1_{\frac{\tau}{z}}d\xi\left(\frac{(1+\xi^2)\ln(1-\xi)^2}{1-\xi}\right)_+\frac{x_p}{\xi}q_f\left(\frac{x_p}{\xi}\right). \end{align} Regarding the virtual diagrams, each of the virtual diagrams contains UV divergence. However, by adding up all of the virtual diagrams, the UV divergences cancel between the virtual contributions. Indeed, we find \begin{align} &\int d^2k_{g\perp}\left[2C_F\frac{1}{k_{g\perp}^2}-N_c\frac{k_{g\perp}\cdot(k_{g\perp}+\xi k_\perp)}{k_{g\perp}^2(k_{g\perp}+\xi k_\perp)^2}+\frac{1}{N_c}\frac{k_{g\perp}\cdot(k_{g\perp}-(1-\xi) k_\perp)}{k_{g\perp}^2(k_{g\perp}-(1-\xi) k_\perp)^2}\right]\nonumber\\ =&\int d^2k_{g\perp}\left[\frac{N_c}{2}\frac{\xi^2 k_\perp^2}{k_{g\perp}^2(k_{g\perp}+\xi k_\perp)^2}-\frac{1}{2N_c}\frac{(1-\xi)^2 k_\perp^2}{k_{g\perp}^2(k_{g\perp}-(1-\xi) k_\perp)^2}\right]. \end{align} To evaluate the singular part, we need to consider the $\xi$-integral together with the so-called kinematical constraint. As discussed in Ref.~\cite{Mueller:2013wwa, Altinoluk:2014eka,Watanabe:2015tja}, the light cone energy conservation provides the kinematical constraint on $\xi$ as follows \begin{align} \xi\le 1-\frac{l_{\perp}^2}{x_ps}. \label{eq:xi-constraint} \end{align} Under this kinematical constraint, the $\xi$-integral gives \begin{align} \int_0^{1-\frac{l_\perp^2}{x_ps}} \frac{d\xi}{1-\xi}=\ln\frac{1}{x_g}+\ln\frac{k_{\perp}^2}{l_{\perp}^2} \label{eq:singularity-terms} \end{align} where the first term can be identified as the small-$x$ logarithm, while the second term provides us with an additional power correction. Let us define the rapidity gap between the projectile quark and the target as $Y_g=\ln(1/x_g)$. $Y_g$ goes to infinity when $\sqrt s\rightarrow\infty$, which is known as the rapidity divergence. It can be renormalized into the forward scattering amplitude by employing the BFKL equation at leading-logarithmic accuracy in $\alpha_sY_g$ as follows \begin{align} T_{x_g}(x_\perp,y_\perp)=&\;T^{(0)}_{x_g}(x_\perp,y_\perp)\nonumber\\ &+\frac{\alpha_sN_c}{2\pi^2}Y_g\int d^2b_\perp\frac{(x_\perp-y_\perp)^2}{(x_\perp-b_\perp)^2(b_\perp-y_\perp)^2} \left[T_{x_g}(x_\perp,b_\perp)+T_{x_g}(b_\perp,y_\perp)-T_{x_g}(x_\perp,y_\perp)\right], \end{align} where $T^{(0)}_{x_g}$ can be viewed as the forward scattering amplitude at LO. Next, let us consider the second logarithmic correction in Eq.~(\ref{eq:singularity-terms}). This term does not lead to large contribution when the produced quark and the produced gluon have the same order of the transverse momentum as $|k_\perp|\sim |l_\perp|$. However, we know that this logarithmic correction can be important at high $p_{h\perp}\gtrsim Q_s$~\cite{Watanabe:2015tja}. The corresponding contribution from the real diagrams can be written as \begin{align} \frac{\alpha_sN_c}{2\pi^2}\int_\tau^1\frac{dz}{z^2}D_{h/q}x_pq_f(x_p)\Bigg[&2\int\frac{d^2x_\perp d^2x_\perp^\prime}{(2\pi)^2}e^{-ik_\perp\cdot(x_\perp-x_\perp^\prime)}(-T_{x_g}(x_\perp,x_\perp^\prime))\int d^2l_\perp\frac{1}{l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}e^{il_\perp\cdot(x_\perp-x_\perp^\prime)}\nonumber\\ &-2\int d^2l_\perp\frac{k_\perp\cdot l_\perp}{k_\perp^2l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}{\cal F}_{x_g}(k_\perp-l_\perp)\Bigg]. \end{align} Using the dimensional regularization in the $\overline{\rm MS}$ scheme, the first term in the square bracket yields~\cite{Mueller:2013wwa} \begin{align} \int\frac{d^2l_\perp}{l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}e^{-il_\perp\cdot r_\perp}=\pi\Bigg[\frac{1}{\epsilon^2}-\frac{1}{\epsilon}\ln\frac{k_\perp^2}{\mu^2}+\frac{1}{2}\left(\ln\frac{k_\perp^2}{\mu^2}\right)^2-\frac{1}{2}\left(\ln\frac{k_\perp^2r_\perp^2}{c_0^2}\right)^2-\frac{\pi^2}{12}\Bigg] \end{align} where $c_0=2e^{-\gamma_E}$ with $\gamma_E$ being the Euler constant. In fact, the double pole and single pole vanish by adding another contribution associated with the virtual correction together. One can cast the virtual diagrams with the logarithmic term into \begin{align} -\frac{\alpha_sN_c}{2\pi^2}\int^1_\tau\frac{dz}{z^2}D_{h/q}(z)x_pq_f(x_p){\cal F}_{x_g}(k_\perp)\Bigg[2\int \frac{d^2l_\perp}{l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}-2\int d^2l_\perp\frac{l_\perp\cdot(l_\perp+k_\perp)}{l_\perp^2(l_\perp+k_\perp)^2}\ln\frac{k_\perp^2}{l_\perp^2}\Bigg]. \label{eq:power-correction-virtual} \end{align} As shown in the Appendix~\ref{AppendixA}, we can cast the above divergent integral into the following form and find \begin{eqnarray} && \int \frac{d^2l_\perp}{(2\pi)^2} \left[\frac{1}{l_\perp^2}-\frac{1}{(l_\perp +k_\perp)^2} +\frac{k_\perp^2}{l_\perp^2 (l_\perp +k_\perp)^2} \right] \ln\frac{k_\perp^2}{l_\perp^2} \notag \\ &=& \frac{1}{2\pi}\left[\frac{1}{\epsilon^2}-\frac{1}{\epsilon}\ln\frac{k_\perp^2}{\mu^2}+\frac{1}{2}\left(\ln\frac{k_\perp^2}{\mu^2}\right)^2-\frac{\pi^2}{12}\right] = 2 \int \frac{d^2l_\perp}{(2\pi)^2} \frac{k_\perp^2}{l_\perp^2 (l_\perp +k_\perp)^2} \ln\frac{k_\perp^2}{l_\perp^2}. \label{qqvirtual} \end{eqnarray} After adding the real correction and virtual correction together, we obtain \begin{align} \frac{\alpha_s}{2\pi}\int^1_\tau \frac{dz}{z^2}D_{h/q}(z)x_pq_f(x_p)\Bigg[-N_c\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}(-T_{x_g}(x_\perp,y_\perp))e^{-ik_\perp\cdot(x_\perp-y_\perp)}\left(\ln\frac{k_\perp^2(x_\perp-y_\perp)^2}{c_0^2}\right)^2 \nonumber\\ -\frac{2N_c}{\pi}\int d^2l_\perp \frac{k_\perp\cdot l_\perp}{k_\perp^2l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}{\cal F}_{x_g}(k_\perp-l_\perp)\Bigg]. \label{eq:doublelog-correction} \end{align} To reach Eq.~(\ref{eq:doublelog-correction}), we have used the same technique as calculations for deriving Sudakov factor~\cite{Mueller:2013wwa}. However, the inclusive hadron production has only one kinematical hard scale, therefore it does not lead to Sudakov factors~\cite{Watanabe:2015tja}. Finally, let us deal with collinear singularities in Eq.~(\ref{eq:qq-real+virtual}). There are two kinds of collinear singularities: one type of singularity corresponds to the gluon radiation from the incoming quark in the initial state depicted in FIG.~\ref{fig:NLO-qq} $(a)$ and the second one is associated with the final state gluon radiation shown in FIG.~\ref{fig:NLO-qq} $(b)$. In order to extract these collinear singularities from the real diagrams, we use the following identities \begin{align} \int d^2k_{g\perp}\frac{1}{(k_\perp-k_{g\perp})^2}{\cal F}_{x_g}(k_{g\perp}) =\;&\pi\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-ik_\perp\cdot(x_\perp-y_\perp)}(-T_{x_g}(x_\perp,y_\perp))\left(-\frac{1}{\hat\epsilon}+\ln\frac{c_0^2}{\mu^2(x_\perp-y_\perp)^2}\right) \label{eq:real-collinear-identity1}\\ \int d^2k_{g\perp}\frac{1}{(k_\perp-\xi k_{g\perp})^2}{\cal F}_{x_g}(k_{g\perp}) =\;&\frac{\pi}{\xi^2}\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-i\frac{k_\perp}{\xi}\cdot(x_\perp-y_\perp)}(-T_{x_g}(x_\perp,y_\perp))\left(-\frac{1}{\hat\epsilon}+\ln\frac{c_0^2}{\mu^2(x_\perp-y_\perp)^2}\right), \label{eq:real-collinear-identity2} \end{align} where the dimensional regularization in the $\overline{\rm MS}$ scheme has been used by setting $1/\hat\epsilon=(4\pi e^{-\gamma_E})^\epsilon/\epsilon$. For the virtual contributions, there are also useful identities \begin{align} \int d^2k_{g\perp}\frac{(\xi k_\perp)^2}{k_{g\perp}^2(k_{g\perp}+\xi k_\perp)^2}&=2\pi\left[-\frac{1}{\hat\epsilon}+\ln\frac{\xi^2k_\perp^2}{\mu^2}\right] \label{eq:virtual-collinear-identity1}\\ \int d^2k_{g\perp}\frac{((1-\xi) k_\perp)^2}{k_{g\perp}^2(k_{g\perp}-(1-\xi) k_\perp)^2}&=2\pi\left[-\frac{1}{\hat\epsilon}+\ln\frac{(1-\xi)^2k_\perp^2}{\mu^2}\right]. \label{eq:virtual-collinear-identity2} \end{align} By adding the LO result, the real contributions and the virtual contributions together, we can absorb the collinear singularity associated with the initial state radiation into the definition of the quark PDF as follows \begin{align} q_f(x_p,\mu)=q^{(0)}_f(x_p)-\frac{1}{\hat{\epsilon}}\frac{\alpha_s(\mu)}{2\pi}\int_{\frac{\tau}{z}}^1\frac{d\xi}{\xi} C_F{\cal P}_{qq}(\xi)q_f\left(\frac{x_p}{\xi}\right), \end{align} which is exactly the DGLAP evolution equation for the quark PDF. The collinear singularity associated with the final state radiation can be renormalized into the quark FF accordingly \begin{align} D_{h/q}(z,\mu)=D_{h/q}^{(0)}(z)-\frac{1}{\hat{\epsilon}}\frac{\alpha_s(\mu)}{2\pi}\int_{z}^1\frac{d\xi}{\xi} C_F{\cal P}_{qq}(\xi)D_{h/q}\left(\frac{z}{\xi}\right), \end{align} with ${\cal P}_{qq}(\xi)=\frac{1+\xi^2}{(1-\xi)_+}+\frac{3}{2}\delta(1-\xi)$. At the end of the day, all the rest of the contributions are finite. The $q\rightarrow q$ channel contribution of the differential cross section can be written as \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/q+X}_{(qq)}}{d^2p_{h\perp}dy}=\sum_f\frac{\alpha_s}{2\pi}\int^1_\tau \frac{dz}{z^2}D_{h/q}(z)\int^1_{x_p}\frac{dx}{x}\xi xq_f(x)\int\frac{d^2 x_\perp d^2 y_\perp}{(2\pi)^2}(-T_{x_g}(x_\perp, y_\perp))\left[{\cal H}_{qq}^{(1)}+\int\frac{d^2b_\perp}{(2\pi)^2}{\cal H}_{qq}^{(2)}\right], \label{eq:qq-final-coordinate} \end{align} where the finite hard scattering parts are given by \begin{align} {\cal H}_{qq}^{(1)} = &\;C_F{\cal P}_{qq}(\xi)\ln\frac{c_0^2}{\mu^2r_\perp^2}\left(e^{-ik_{\perp}\cdot r_\perp}+\frac{1}{\xi^2}e^{-i\frac{k_\perp}{\xi}\cdot r_\perp}\right) -3C_F\delta(1-\xi)e^{-ik_\perp\cdot r_\perp}\ln\frac{c_0^2}{k_\perp^2r_\perp}\nonumber\\ &-N_c\delta(1-\xi)e^{-ik_\perp\cdot r_\perp}\int_0^1d\xi^\prime\frac{1+\xi^{\prime2}}{(1-\xi^\prime)_+}\ln\xi^{\prime2} -\frac{1}{N_c}e^{-ik_\perp\cdot r_\perp}\left(\frac{(1+\xi^2)\ln(1-\xi)^2}{1-\xi}\right)_+\nonumber\\ &-N_c\delta(1-\xi)e^{-ik_\perp\cdot r_\perp}\left(\ln\frac{k_\perp^2r_\perp^2}{c_0^2}\right)^2 \label{eq:Hqq1} \end{align} and \begin{align} {\cal H}_{qq}^{(2)}=4\pi\frac{1}{N_c}e^{-ik_\perp\cdot r_\perp}\frac{1+\xi^2}{(1-\xi)_+}I_{qq}^{(1)}+4\pi N_ce^{-ik_\perp\cdot r_\perp}\frac{1+\xi^2}{(1-\xi)_+}I_{qq}^{(2)}-8N_c\pi\delta(1-\xi)\frac{b_\perp\cdot r_\perp}{b_\perp^2 r_\perp^2}\ln\frac{k_\perp^2 r_\perp^2}{c_0^2}e^{-ik_\perp\cdot(b_\perp+r_\perp)} \label{eq:Hqq2} \end{align} with $r_\perp=x_\perp-y_\perp$ and \begin{align} I_{qq}^{(1)}&=e^{-i(1-\xi)k_\perp\cdot b_\perp}\left[\frac{b_\perp\cdot(\xi b_\perp-r_\perp)}{b_\perp^2(\xi b_\perp-r_\perp)^2}-\frac{1}{b_\perp^2}\right]+e^{-ik_\perp\cdot b_\perp}\frac{1}{b_\perp^2},\label{eq:I1qq}\\ I_{qq}^{(2)}&=\frac{b_\perp\cdot r_\perp}{b_\perp^2 r_\perp^2}\left[e^{-ik_\perp\cdot b_\perp}+\frac{1}{\xi}e^{-ik_\perp\cdot b_\perp}e^{-i(\frac{1-\xi}{\xi})k_\perp\cdot r_\perp}\right]. \end{align} In addition, by using the following identities~\cite{Watanabe:2015tja} \begin{align} \int\frac{d^2r_\perp}{(2\pi)^2}\ln\frac{c_0^2}{\mu^2r_\perp^2}e^{-ik_\perp\cdot r_\perp}&=\frac{1}{\pi}\left[\frac{1}{k_\perp^2}-2\pi \delta^{(2)}(k_\perp)\int_0^\infty\frac{dl_\perp}{l_\perp}J_0\left(\frac{c_0}{\mu}l_\perp\right)\right],\\ \left(\ln\frac{k_\perp^2r_\perp^2}{c_0^2}\right)^2&=8\pi\int\frac{d^2l_\perp}{(2\pi)^2l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}\left[\theta(k_\perp-l_\perp)-e^{-il_\perp\cdot r_\perp}\right], \end{align} which gives \begin{align} \int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}(-T_{x_g}(x_\perp,y_\perp))\ln\frac{c_0^2}{\mu^2r_\perp^2}e^{-ik_\perp\cdot r_\perp} &=\frac{1}{\pi}\int\frac{d^2l_\perp}{l_\perp^2}\left[{\cal F}_{x_g}(k_\perp+l_\perp)-J_0\left(\frac{c_0}{\mu}l_\perp\right){\cal F}_{x_g}(k_\perp)\right],\\ \int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}(-T_{x_g}(x_\perp,y_\perp))\left(\ln\frac{k_\perp^2r_\perp^2}{c_0^2}\right)^2e^{-ik_\perp\cdot r_\perp} &=\frac{2}{\pi}\int \frac{d^2l_\perp}{l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}\left[\theta(k_\perp-l_\perp){\cal F}_{x_g}(k_\perp)-{\cal F}_{x_g}(k_\perp+l_\perp)\right]. \end{align} we can cast the NLO corrections into expressions in the momentum space \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/q+X}_{(qq)}}{d^2p_{h\perp}dy}=\sum_f\frac{\alpha_s}{2\pi}\int^1_\tau \frac{dz}{z^2}D_{h/q}(z)\int^1_{x_p}\frac{dx}{x} \xi xq_f(x)S_{qq}, \label{eq:qq-final-momentum} \end{align} where \begin{align} &S_{qq}=\nonumber\\ &C_F{\cal P}_{qq}(\xi)\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}^2} \Bigg[ {\cal F}_{x_g}(k_\perp+k_{g\perp})-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}(k_\perp) +\frac{1}{\xi^2}{\cal F}_{x_g}\left(\frac{k_\perp}{\xi}+k_{g\perp}\right)-\frac{1}{\xi^2}J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}\left(\frac{k_\perp}{\xi}\right)\Bigg]\nonumber\\ &-3C_F\delta(1-\xi)\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}^2} \left[{\cal F}_{x_g}(k_\perp+k_{g\perp})-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}(k_\perp)\right] +N_c\frac{1+\xi^2}{(1-\xi)_+}\overline{I}^{(2)}_{qq} \nonumber\\ & -(2C_F-N_c)\left[\frac{1+\xi^2}{(1-\xi)_+}\overline{I}^{(1)}_{qq}-\left(\frac{(1+\xi^2)\ln(1-\xi)^2}{1-\xi}\right)_+{\cal F}_{x_g}(k_\perp)\right] -N_c\delta(1-\xi){\cal F}_{x_g}(k_\perp)\int_0^1d\xi^\prime\frac{1+\xi^{\prime2}}{(1-\xi^\prime)_+}\ln \xi^{\prime2}\nonumber\\ &-\frac{2N_c}{\pi}\delta(1-\xi)\int\frac{d^2k_{g\perp}}{k_{g\perp}^2}\ln\frac{k_\perp^2}{k_{g\perp}^2}\left\{\theta(k_\perp-k_{g\perp}){\cal F}_{x_g}(k_\perp)-{\cal F}_{x_g}(k_\perp-k_{g\perp})\right\}\nonumber\\ &-\frac{2N_c}{\pi}\delta(1-\xi)\int d^2k_{g\perp} \frac{k_\perp\cdot k_{g\perp}}{k_\perp^2k_{g\perp}^2}\ln\frac{k_\perp^2}{k_{g\perp}^2}{\cal F}_{x_g}(k_\perp-k_{g\perp}), \end{align} with \begin{align} \overline{I}^{(2)}_{qq}&=\int\frac{d^2k_{g\perp}}{\pi}{\cal F}_{x_g}(k_{g\perp})\left[\frac{k_{\perp}\cdot(k_{g\perp}-k_\perp)}{k_{\perp}^2(k_{g\perp}-k_\perp)^2}+\frac{k_{\perp}\cdot(\xi k_{g\perp}-k_\perp)}{k_{\perp}^2(\xi k_{g\perp}-k_\perp)^2}\right]. \end{align} \subsection{The $g\rightarrow g$ channel} \begin{figure} \centering \includegraphics[width=16cm,angle=0]{gg-real.eps} \caption{Real diagrams at NLO for the $g\rightarrow g$ channel. } \label{fig:NLO-gg-real} \end{figure} Now let us consider the $g\rightarrow g$ channel. In the dense regime, the calculations can be more complicated, since the produced two gluons probe higher multi-point Wilson line correlators such as sextupole and quadrupole~\cite{Chirilli:2012jd}, which is suppressed by factors of $N_c$. However, in the dilute regime, we do not have to deal with such problem. For $g+{\rm p}\rightarrow g+g+X$ process shown in FIG.~\ref{fig:NLO-gg-real}, the differential cross section for producing one gluon with $k$ and another gluon $l$ is given by \begin{align} &\frac{d\sigma^{g+{\rm p}\rightarrow g+g+X}}{d^3kd^3l}=\alpha_s\delta(p^+-k^+-l^+)\int\frac{d^2x_\perp d^2x_\perp^\prime d^2b_\perp d^2 b_\perp^\prime}{(2\pi)^8}e^{-ik_{\perp}\cdot(x_\perp-x_\perp^\prime)}e^{-il_{\perp}\cdot(b_\perp-b_\perp^\prime)}\sum_{\lambda_1,\lambda_2,\lambda}\psi_{gg\lambda_1\lambda_2}^{\lambda\ast}(u_\perp^\prime)\psi_{gg\lambda_1\lambda_2}^{\lambda}(u_\perp)\nonumber\\ &\times\Bigg[N_c\widetilde{S}_{x_g}(x_\perp,x_\perp^\prime)+N_c\widetilde{S}_{x_g}(v_\perp,v_\perp^\prime)-\frac{N_c}{2}\widetilde{S}_{x_g}(v_\perp,x_\perp^\prime)-\frac{N_c}{2}\widetilde{S}_{x_g}(v_\perp,b_\perp^\prime)-\frac{N_c}{2}\widetilde{S}_{x_g}(x_\perp,v_\perp^\prime)-\frac{N_c}{2}\widetilde{S}_{x_g}(b_\perp,v_\perp^\prime)\nonumber\\ &~~~~-\frac{N_c}{2}\widetilde{S}_{x_g}(x_\perp,b_\perp^\prime)-\frac{N_c}{2}\widetilde{S}_{x_g}(b_\perp,x_\perp^\prime)+N_c\widetilde{S}_{x_g}(b_\perp,b_\perp^\prime)\Bigg]. \label{eq:gg-real} \end{align} where the color factors are computed by using FeynCalc package~\cite{Shtabovenko:2016sxi}. The light cone wave function for $g\rightarrow gg$ splitting is defined as~\cite{Dominguez:2011wm} \begin{align} \psi^\lambda_{gg\lambda_1\lambda_2}(u_\perp)=\sqrt{\frac{2\xi(1-\xi)}{p^+}}\frac{2\pi i}{u_\perp^2}\Bigg[\frac{1}{\xi}u_\perp\cdot\varepsilon^{(\lambda_1)}_{\perp}\varepsilon^{(\lambda)\ast}_{\perp}\cdot\varepsilon^{(\lambda_2)}_{\perp} &+\frac{1}{1-\xi}u_\perp\cdot\varepsilon^{(\lambda_2)}_{\perp}\varepsilon^{(\lambda)\ast}_{\perp}\cdot\varepsilon^{(\lambda_1)}_{\perp} -u_\perp\cdot\varepsilon^{(\lambda)\ast}_{\perp}\varepsilon^{(\lambda_1)}_{\perp}\cdot\varepsilon^{(\lambda_2)}_{\perp}\Bigg] \end{align} and, by summing over all of the gluon polarizations, the splitting kernel can be written as \begin{align} \sum_{\lambda_1,\lambda_2,\lambda}\psi_{gg\lambda_1\lambda_2}^{\lambda\ast}(u_\perp^\prime)\psi_{gg\lambda_1\lambda_2}^{\lambda}(u_\perp)=\frac{4(2\pi)^2}{p^+}\left[\frac{\xi}{1-\xi}+\frac{1-\xi}{\xi}+\xi(1-\xi)\right]\frac{u_\perp^\prime\cdot u_\perp}{u_\perp^{\prime2}u_\perp^2}. \end{align} For inclusive hadron production with the transverse momentum $p_{h\perp}$ at the rapidity $y$, Eq.~(\ref{eq:gg-real}) becomes \begin{align} &\frac{\alpha_s}{\pi^2}\int_\tau^1\frac{dz}{z^2}D_{h/g}(z)\int d^2l_\perp \int_{\frac{\tau}{z}}^1d\xi \left[\frac{\xi}{1-\xi}+\frac{1-\xi}{\xi}+\xi(1-\xi)\right]\frac{x_p}{\xi}G\left(\frac{x_p}{\xi}\right)\int\frac{d^2x_\perp d^2x_\perp^\prime d^2b_\perp d^2 b_\perp^\prime}{(2\pi)^4}\nonumber\\ &\times e^{-ik_{\perp}\cdot(x_\perp-x_\perp^\prime)}e^{-il_{\perp}\cdot(b_\perp-b_\perp^\prime)}\frac{u_\perp^\prime\cdot u_\perp}{u_\perp^{\prime2}u_\perp^2}\Bigg[-N_c\widetilde{T}_{x_g}(x_\perp,x_\perp^\prime)-N_c\widetilde{T}_{x_g}(v_\perp,v_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(v_\perp,x_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(v_\perp,b_\perp^\prime)\nonumber\\ &~~~~+\frac{N_c}{2}\widetilde{T}_{x_g}(x_\perp,v_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(b_\perp,v_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(x_\perp,b_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(b_\perp,x_\perp^\prime)-N_c\widetilde{T}_{x_g}(b_\perp,b_\perp^\prime)\Bigg]. \label{eq:hadron-gg-real} \end{align} \begin{figure} \centering \includegraphics[width=18cm,angle=0]{gg-virtual-a.eps} \caption{Virtual gluon loop corrections at NLO for the $g\rightarrow g$ channel. Mirror contributions are omitted. } \label{fig:NLO-gg-virtual-g} \end{figure} \begin{figure} \centering \includegraphics[width=18cm,angle=0]{gg-virtual-b.eps} \caption{Virtual quark loop corrections at NLO for the $g\rightarrow g$ channel. Mirror contributions are omitted. } \label{fig:NLO-gg-virtual-q} \end{figure} Furthermore, the gluon virtual contributions (FIG.~\ref{fig:NLO-gg-virtual-g}) yield \begin{align} &-\frac{2}{2!}\frac{\alpha_s}{\pi^2}\int^1_\tau\frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\int^1_0 d\xi \left[\frac{\xi}{1-\xi}+\frac{1-\xi}{\xi}+\xi(1-\xi)\right]\int\frac{d^2v_\perp d^2 v_\perp^\prime d^2u_\perp}{(2\pi)^2}e^{-ik_{\perp}\cdot(v_\perp-v_\perp^\prime)}\frac{1}{u_\perp^2}\nonumber\\ &\times\Bigg[-N_c\widetilde{T}_{x_g}(v_\perp,v_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(x_\perp,v_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(b_\perp,v_\perp^\prime)+\frac{N_c}{2}\widetilde{T}_{x_g}(x_\perp,b_\perp)\Bigg], \label{eq:gg-virtual-g} \end{align} where the symmetry factor $1/2!$ and a factor of 2 from the mirror diagrams have been taken into consideration. The $\xi$ dependence of Eq.~(\ref{eq:gg-virtual-g}) is symmetry under the interchange $\xi\leftrightarrow 1-\xi$ and the $\xi$ dependent part of the splitting kernel can be rewritten by $2\left[\frac{\xi}{1-\xi}+\frac{1}{2}\xi(1-\xi)\right]$. It is obvious that this splitting function now only contains the rapidity divergence at $\xi=1$. The last diagram in FIG.~\ref{fig:NLO-gg-virtual-g} does not contribute to the inelastic hadron production as is the case for the $q\rightarrow q$ channel, however, the BFKL evolution equation involves this diagram as we will show below. In addition, we should take into account quark loop corrections (FIG.~\ref{fig:NLO-gg-virtual-q}) which read \begin{align} &-2N_f\frac{\alpha_s}{2\pi^2}\int^1_\tau\frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\int^1_0 d\xi [(1-\xi)^2+\xi^2]\int\frac{d^2v_\perp d^2 v_\perp^\prime d^2u_\perp}{(2\pi)^2}e^{-ik_{\perp}\cdot(v_\perp-v_\perp^\prime)}\frac{1}{u_\perp^2}\nonumber\\ &\times\Bigg[-T_R\widetilde{T}_{x_g}(v_\perp,v_\perp^\prime)+\frac{T_R}{2}\widetilde{T}_{x_g}(x_\perp,v_\perp^\prime)+\frac{T_R}{2}\widetilde{T}_{x_g}(b_\perp,v_\perp^\prime)+\frac{T_R}{2N_c^2}\widetilde{T}_{x_g}(x_\perp,b_\perp)\Bigg], \label{eq:gg-virtual-q} \end{align} where $T_R={1}/{2}$ for SU(3) and $N_f$ is the number of active flavors in the quark loop. A factor of 2 in the front of Eq.~(\ref{eq:gg-virtual-q}) represents the mirror contributions. The light cone wave function for $g\rightarrow q\bar q$ splitting, which is used to obtain Eq.~(\ref{eq:gg-virtual-q}), is given by \begin{align} \psi^{\lambda}_{q\bar{q}\alpha\beta}(u_\perp) =2\pi i\sqrt{\frac{2}{p^+}} \left\{ \begin{array}{c} \frac{u_\perp\cdot\varepsilon_\perp^{(1)}}{u_\perp^2}\left[\xi\delta_{\alpha+}\delta_{\beta-}-(1-\xi)\delta_{\alpha-}\delta_{\beta+}\right]~~~~(\lambda=1)\\ \frac{u_\perp\cdot\varepsilon_\perp^{(2)}}{u_\perp^2}\left[\xi\delta_{\alpha-}\delta_{\beta+}-(1-\xi)\delta_{\alpha+}\delta_{\beta-}\right]~~~~(\lambda=2)\\ \end{array} \right. \end{align} and the sum of splitting kernels is \begin{align} \sum_{\alpha,\beta,\lambda}\psi_{q\bar q \alpha\beta}^{\lambda\ast}(u_\perp^\prime)\psi_{q\bar q \alpha\beta}^{\lambda}(u_\perp)=\frac{2(2\pi)^2}{p^+}\left[(1-\xi)^2+\xi^2\right]\frac{u_\perp^\prime\cdot u_\perp}{u_\perp^{\prime2}u_\perp^2}. \label{eq:splitting-kernel-qqbar} \end{align} One should keep in mind that there is no singularity in the quark loop diagrams at $\xi\rightarrow1$. The rapidity divergence in the $g\rightarrow g$ channel can be dealt in the same fashion as in the $q\rightarrow q$ channel. It is clear that only the real contributions and the gluon virtual loop corrections have the rapidity singularity. By introducing the plus function as in the $q\rightarrow q$ channel, we can separate the nonsingular part and the singular part. The nonsingular part is manifest \begin{align} \frac{\alpha_sN_c}{\pi^2}&\int^1_\tau\frac{dz}{z^2}D_{h/g}(z)\int_{\frac{\tau}{z}}^1d\xi \frac{x_p}{\xi}G\left(\frac{x_p}{\xi}\right)\frac{[1-\xi(1-\xi)]^2}{\xi(1-\xi)_+} \int d^2k_{g\perp}\widetilde{\cal F}_{x_g}(k_{g\perp})\nonumber\\ &\times\left[\frac{1}{(k_\perp-k_{g\perp})^2}+\frac{1}{(k_\perp-\xi k_{g\perp})^2}-\frac{k_\perp\cdot(k_\perp-k_{g\perp})}{k_\perp^2(k_\perp-k_{g\perp})^2}-\frac{k_\perp\cdot(k_\perp-\xi k_{g\perp})}{k_\perp^2(k_\perp-\xi k_{g\perp})^2}-\frac{(k_\perp-k_{g\perp})\cdot(k_\perp-\xi k_{g\perp})}{(k_\perp-k_{g\perp})^2(k_\perp-\xi k_{g\perp})^2}\right]\nonumber\\ -\frac{\alpha_sN_c}{\pi^2}&\int_{\tau}^1\frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\int_0^1d\xi \left[\frac{\xi}{(1-\xi)_+}+\frac{\xi(1-\xi)}{2}\right]\widetilde{\cal F}_{x_g}(k_\perp)\nonumber\\ &\times\int d^2k_{g\perp}\frac{1}{2}\left[\frac{(1-\xi)^2k_\perp^2}{k_{g\perp}^2(k_{g\perp}-(1-\xi)k_\perp)^2}+\frac{\xi^2k_\perp^2}{k_{g\perp}^2(k_{g\perp}+\xi k_\perp)^2}\right]. \end{align} As we mentioned previously, the singular part is also separated the rapidity divergent part from the logarithmic power correction. Again, the rapidity divergence can be renormalized into the definition of the wave function of the target proton as follows \begin{align} \widetilde{T}_{x_g}(x_\perp,y_\perp)=&\;\widetilde{T}^{(0)}_{x_g}(x_\perp,y_\perp)\nonumber\\ &+\frac{\alpha_sN_c}{2\pi^2}Y_g\int d^2b_\perp\frac{(x_\perp-y_\perp)^2}{(x_\perp-b_\perp)^2(b_\perp-y_\perp)^2} \left[\widetilde{T}_{x_g}(x_\perp,b_\perp)+\widetilde{T}_{x_g}(b_\perp,y_\perp)-\widetilde{T}_{x_g}(x_\perp,y_\perp)\right] \end{align} which is equivalent to the BFKL evolution equation for the fundamental dipole amplitude, since $\widetilde{T}_{x_g}$ can be simply replaced with $N_cT_{x_g}/C_F$ in the dilute regime of the target proton. The remaining power correction yields \begin{align} \frac{\alpha_sN_c}{2\pi^2}&\int_{\tau}^1\frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\Bigg[2\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-ik_\perp\cdot r_\perp}(-\widetilde{T}_{x_g}(r_\perp))\int d^2l_\perp\frac{1}{l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}e^{il_\perp\cdot r_\perp}\nonumber\\ &-2\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}e^{-ik_\perp\cdot r_\perp}(-\widetilde{T}_{x_g}(r_\perp))\int d^2l_\perp \frac{k_\perp\cdot l_\perp}{k_\perp^2 l_\perp^2}e^{il_\perp\cdot r_\perp} \Bigg]\nonumber\\ -\frac{\alpha_sN_c}{2\pi^2}&\int_{\tau}^1\frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\widetilde{\cal F}_{x_g}(k_\perp)\left[2\int d^2l_\perp\frac{1}{l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}-2\int d^2l_\perp\frac{l_\perp\cdot(l_\perp+k_\perp)}{l_\perp^2(l_\perp+k_\perp)^2}\ln\frac{k_\perp^2}{l_\perp^2}\right]. \label{eq:gg-power-correction} \end{align} Then, one finds immediately Eq.~(\ref{eq:gg-power-correction}) can be put into \begin{align} \frac{\alpha_s}{2\pi}\int^1_\tau \frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\Bigg[-N_c\int\frac{d^2 x_\perp d^2 y_\perp}{(2\pi)^2}(-\widetilde{T}_{x_g}(x_\perp, y_\perp))\left(\ln\frac{k_\perp^2r_\perp^2}{c_0^2}\right)^2e^{-ik_\perp\cdot r_\perp}\nonumber\\ -\frac{2N_c}{\pi}\int d^2l_\perp \frac{k_\perp\cdot l_\perp}{k_\perp^2l_\perp^2}\ln\frac{k_\perp^2}{l_\perp^2}\widetilde{\cal F}_{x_g}(k_\perp-l_\perp)\Bigg] \label{eq:doublelog-correction-gg} \end{align} with $r_\perp=x_\perp-y_\perp$. In the meantime, the quark virtual corrections is simply given by \begin{align} -\frac{\alpha_sN_fT_R}{2\pi^2}&\int_{\tau}^1\frac{dz}{z^2}D_{h/g}(z)x_pG(x_p)\int_0^1d\xi \left[(1-\xi)^2+\xi^2\right]\nonumber\\ &\times \widetilde{\cal F}_{x_g}(k_\perp)\int d^2k_{g\perp}\frac{1}{2}\left[\frac{(1-\xi)^2k_\perp^2}{k_{g\perp}^2(k_{g\perp}-(1-\xi)k_\perp)^2}+\frac{\xi^2k_\perp^2}{k_{g\perp}^2(k_{g\perp}+\xi k_\perp)^2}\right]. \end{align} Here, the remaining task is to extract the collinear singularities from the real and the virtual contributions, and absorb the pole singularity associated with the initial state radiation into the gluon PDF and another singularity associated with the final state radiation into the gluon FF, respectively. This can be done by using the identities Eqs.~(\ref{eq:real-collinear-identity1})--(\ref{eq:virtual-collinear-identity2}) as follows \begin{align} G(x,\mu)&=G^{(0)}(x)-\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int^1_x\frac{d\xi}{\xi}N_c{\cal P}_{gg}(\xi)G\left(\frac{x}{\xi}\right),\\ D_{h/g}(z,\mu)&=D_{h/g}^{(0)}(z)-\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int^1_z\frac{d\xi}{\xi}N_c{\cal P}_{gg}(\xi)D_{h/g}\left(\frac{z}{\xi}\right) \end{align} where the LO splitting function is defined as \begin{align} {\cal P}_{gg}(\xi)=2\left[\frac{\xi}{(1-\xi)_+}+\frac{1-\xi}{\xi}+\xi(1-\xi)\right]+\left(\frac{11}{6}-\frac{2N_fT_R}{3N_c}\right)\delta(1-\xi). \label{eq:splitting-function-gg} \end{align} Finally, for the $g\rightarrow g$ channel, the differential cross section in the coordinate space can be written by \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/g+X}_{(gg)}}{d^2p_{h\perp}dy}=\frac{\alpha_s}{2\pi}\int^1_\tau\frac{dz}{z^2}D_{h/g}(z)\int^1_{x_p}\frac{dx}{x} \xi xG(x)\int\frac{d^2x_\perp d^2y_\perp}{(2\pi)^2}(-\widetilde{T}_{x_g}(x_\perp,y_\perp))\left[{\cal H}_{gg}^{(1)}+\int\frac{d^2b_\perp}{(2\pi)^2}{\cal H}_{gg}^{(2)}\right] \label{eq:gg-final-coordinate} \end{align} where the hard parts are given by \begin{align} {\cal H}_{gg}^{(1)}=~&N_c{\cal P}_{gg}(\xi)\left(e^{-ik_{\perp}\cdot r_\perp}+\frac{1}{\xi^2}e^{-i\frac{k_\perp}{\xi}\cdot r_\perp}\right)\ln\frac{c_0^2}{\mu^2r_\perp^2} -\left(\frac{11}{3}-\frac{4N_fT_R}{3N_c}\right)N_c\delta(1-\xi)e^{-ik_\perp\cdot r_\perp}\ln\frac{c_0^2}{k_\perp^2r_\perp^2}\nonumber\\ &-N_c\delta(1-\xi)e^{-ik_\perp\cdot r_\perp}\int^1_0d\xi^\prime 2\left[\frac{\xi^\prime}{(1-\xi^\prime)_+}+\frac{1}{2}\xi^\prime(1-\xi^\prime)\right]\left[\ln\xi^{\prime2}+\ln(1-\xi^\prime)^2\right]\nonumber\\ &-N_fT_R\delta(1-\xi)e^{-ik_\perp\cdot r_\perp}\int^1_0d\xi^\prime\left[(1-\xi^\prime)^2+\xi^{\prime2}\right]\left[\ln\xi^{\prime2}+\ln(1-\xi^\prime)^2\right] -N_c\delta(1-\xi)\left(\ln\frac{k_\perp^2r_\perp^2}{c_0^2}\right)^2e^{-ik_\perp\cdot r_\perp}, \label{eq:Hgg1}\\ {\cal H}_{gg}^{(2)}=~&8\pi N_c \frac{[1-\xi(1-\xi)]^2}{\xi(1-\xi)_+}e^{-ik_\perp\cdot r_\perp}\Bigg[\frac{b_\perp\cdot r_\perp}{b_\perp^2r_\perp^2}e^{-ik_\perp\cdot b_\perp}\left(1+\frac{1}{\xi}e^{-i\left(\frac{1-\xi}{\xi}\right)k_\perp\cdot r_\perp}\right)\nonumber\\ &-\frac{1}{\xi}\frac{(x_\perp-b_\perp)\cdot(y_\perp-b_\perp)}{(x_\perp-b_\perp)^2(y_\perp-b_\perp)^2}e^{-i\left(\frac{1-\xi}{\xi}\right)k_\perp\cdot(b_\perp-y_\perp)}\Bigg] -8N_c\pi \frac{b_\perp\cdot r_\perp}{b_\perp^2 r_\perp^2}\ln\frac{k_\perp^2 r_\perp^2}{c_0^2}e^{-ik_\perp\cdot(b_\perp+r_\perp)} \label{eq:Hgg2} \end{align} with $r_\perp=x_\perp-y_\perp$. Through Fourier transform, the differential cross section for the $g\rightarrow g$ channel in the momentum space is \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/g+X}_{(gg)}}{d^2p_{h\perp}dy}=\frac{\alpha_s}{2\pi}\int^1_\tau\frac{dz}{z^2}D_{h/g}(z)\int^1_{x_p}\frac{dx}{x} \xi xG(x)S_{gg} \label{eq:gg-final-momentum} \end{align} where \begin{align} S_{gg}= &\;N_c{\cal P}_{gg}(\xi)\Bigg\{\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}}\left[\widetilde{\cal F}_{x_g}(k_{\perp}+k_{g\perp})-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right)\widetilde{\cal F}_{x_g}(k_{\perp})\right]\nonumber\\ &+\frac{1}{\xi^2}\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}}\left[\widetilde{\cal F}_{x_g}\left(\frac{k_{\perp}}{\xi}+k_{g\perp}\right)-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right)\widetilde{\cal F}_{x_g}\left(\frac{k_{\perp}}{\xi}\right)\right] \Bigg\}\nonumber\\ &-\left(\frac{11}{3}-\frac{4N_fT_R}{3N_c}\right)N_c\delta(1-\xi)\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}}\left[\widetilde{\cal F}_{x_g}(k_{\perp}+k_{g\perp})-J_0\left(\frac{c_0}{k_\perp}k_{g\perp}\right)\widetilde{\cal F}_{x_g}(k_{\perp})\right]\nonumber\\ &-\delta(1-\xi)\int^1_0d\xi^\prime \left\{N_c\left[\frac{2\xi^\prime}{(1-\xi^\prime)_+}+\xi^\prime(1-\xi^\prime)\right]+N_fT_R\left[(1-\xi^\prime)^2+\xi^{\prime2}\right]\right\}\left[\ln\xi^{\prime2}+\ln(1-\xi^\prime)^2\right]\widetilde{\cal F}_{x_g}(k_\perp)\nonumber\\ &-8\pi N_c\frac{[1-\xi(1-\xi)]^2}{\xi(1-\xi)_+} \int\frac{d^2k_{g\perp}}{(2\pi)^2}\widetilde{\cal F}_{x_g}(k_{g\perp})\left[\frac{(k_\perp-k_{g\perp})\cdot(k_{\perp}-\xi k_{g\perp})}{(k_\perp-k_{g\perp})^2(k_{\perp}-\xi k_{g\perp})^2} +\frac{k_\perp\cdot(k_{\perp}-k_{g\perp})}{k_\perp^2(k_{\perp}-k_{g\perp})^2} +\frac{k_{\perp}\cdot(k_{\perp}-\xi k_{g\perp})}{k_{\perp}^2(k_{\perp}-\xi k_{g\perp})^2}\right]\nonumber\\ &-\frac{2N_c}{\pi}\delta(1-\xi)\int\frac{d^2k_{g\perp}}{k_{g\perp}^2}\ln\frac{k_\perp^2}{k_{g\perp}^2}\left\{\theta(k_\perp-k_{g\perp})\widetilde{\cal F}_{x_g}(k_\perp)-\widetilde{\cal F}_{x_g}(k_\perp-k_{g\perp})\right\}\nonumber\\ &-\frac{2N_c}{\pi}\delta(1-\xi)\int d^2k_{g\perp} \frac{k_\perp\cdot k_{g\perp}}{k_\perp^2k_{g\perp}^2}\ln\frac{k_\perp^2}{k_{g\perp}^2}\widetilde{\cal F}_{x_g}(k_\perp-k_{g\perp}). \end{align} \subsection{The $q\rightarrow g$ channel} For the $q\rightarrow g$ channel, there is no virtual correction. The relevant diagrams are the same as in FIG.~\ref{fig:NLO-qq} but the radiated gluon is measured in final state. The differential cross section of ${q+{\rm p}\rightarrow g+q+X}$ for producing a gluon with momentum $k$ and a quark with momentum $l$ is given by \begin{align} &\frac{d\sigma^{q+{\rm p}\rightarrow g+q+X}}{d^3kd^3l} =\;\alpha_s\delta(p^+-k^+-l^+) \int\frac{d^2x_\perp d^2x_\perp^\prime d^2b_\perp d^2b_\perp^\prime}{(2\pi)^4} e^{-ik_{\perp}\cdot(x_\perp-x_\perp^\prime)}e^{-il\perp\cdot(b_\perp-b_\perp^\prime)}\sum\limits_{\alpha\beta\lambda}\psi_{gq\alpha\beta}^{\lambda\ast}(u_\perp^\prime)\psi_{gq\alpha\beta}^{\lambda}(u_\perp)\nonumber\\ \times& \Bigg[C_FS_{x_g}({x_\perp,x_\perp^\prime})+C_FS_{x_g}({v_\perp,v_\perp^\prime}) -\frac{N_c}{2}S_{x_g}({b_\perp,v_\perp^\prime})-\frac{N_c}{2}S_{x_g}({v_\perp,b_\perp^\prime})-\frac{N_c}{2}S_{x_g}({b_\perp,x_\perp^\prime})-\frac{N_c}{2}S_{x_g}({x_\perp,b_\perp^\prime})\nonumber\\ &+C_F\widetilde{S}_{x_g}({b_\perp,b_\perp^\prime})+\frac{1}{2N_c}S_{x_g}({x_\perp,v_\perp^\prime})+\frac{1}{2N_c}S_{x_g}({v_\perp,x_\perp^\prime}) \Bigg] \label{eq:qg-real} \end{align} where the splitting kernel is \begin{align} \sum\limits_{\alpha\beta\lambda}\psi_{gq\alpha\beta}^{\lambda\ast}(u_\perp^\prime)\psi_{gq\alpha\beta}^{\lambda}(u_\perp) =\frac{2(2\pi)^2}{p^+}\frac{1+(1-\xi)^2}{\xi}\frac{u_\perp^\prime\cdot u_\perp}{u_\perp^{\prime2}u_\perp^2}. \end{align} By convoluting with the quark PDF and the gluon FF, Eq.~(\ref{eq:qg-real}) is cast into \begin{align} &\sum_f\frac{\alpha_s}{2\pi^2}\int^1_\tau\frac{dz}{z^2}D_{h/g}(z)\int^1_{\frac{\tau}{z}} d\xi{\cal P}_{gq}(\xi)\frac{x_p}{\xi}q_f\left(\frac{x_p}{\xi}\right)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\Bigg[C_F\frac{1}{(k_{\perp}-\xi k_{g\perp})^2}+C_F\frac{N_c}{C_F}\frac{1}{(k_{g\perp}-k_{\perp})^2}\nonumber\\ &-N_c\frac{(\xi k_{g\perp}-k_{\perp})\cdot(k_{g\perp}-k_{\perp})}{(\xi k_{g\perp}-k_{\perp})^2(k_{g\perp}-k_{\perp})^2}-N_c\frac{k_{\perp}\cdot(k_{\perp}-k_{g\perp})}{k_{\perp}^2(k_{\perp}-k_{g\perp})^2}+\frac{1}{N_c}\frac{k_{\perp}\cdot(k_{\perp}-\xi k_{g\perp})}{k_{\perp}^2(k_{\perp}-\xi k_{g\perp})^2} \Bigg] \label{eq:gq-momentum} \end{align} where the splitting function at LO is defined as \begin{align} {\cal P}_{gq}(\xi)=\frac{1+(1-\xi)^2}{\xi}. \label{eq:splitting-function-gq} \end{align} The lower limit of the $\xi$-integral is constrained by the kinematics and then the $\xi$ never reaches 0. The remaining task in this channel is to extract the collinear divergences from Eq.~(\ref{eq:gq-momentum}). As we mentioned in the previous sections, the collinear divergences associated with the initial state and the final state radiations can be renormalized into the definition of the gluon PDF and the quark FF, respectively \begin{align} G(x,\mu)&=G^{(0)}(x)-\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int^1_{x}\frac{d\xi}{\xi} C_F\sum_f{\cal P}_{gq}(\xi)q_f\left(\frac{x}{\xi}\right)\\ D_{h/q}(z,\mu)&=D_{h/q}^{(0)}(z)-\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int^1_{z}\frac{d\xi}{\xi} C_F{\cal P}_{gq}(\xi)D_{h/g}\left(\frac{z}{\xi}\right). \end{align} In the end, it is easy to obtain the differential cross section in the coordinate space as \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/g+X}_{(gq)}}{d^2p_{h\perp}dy} =\sum_f\frac{\alpha_s}{2\pi}\int_\tau^1\frac{dz}{z^2}D_{h/g}(z)\int_{x_p}^1\frac{dx}{x} \xi xq_f(x)\int\frac{d^2x_\perp d^2y_{\perp}}{(2\pi)^2}(-T_{x_g}(x_\perp,y_\perp))\left[{\cal H}^{(1)}_{gq}+\int\frac{d^2b_\perp}{(2\pi)^2}{\cal H}^{(2)}_{gq}\right] \label{eq:gq-final-coordinate} \end{align} where \begin{align} {\cal H}^{(1)}_{gq}=&\;C_F\frac{1}{\xi^2}{\cal P}_{gq}(\xi)e^{-i\frac{k_\perp}{\xi}\cdot r_\perp}\ln\frac{c_0^2}{\mu^2 r_\perp^2}+N_c{\cal P}_{gq}(\xi)e^{-ik_\perp \cdot r_\perp}\ln\frac{c_0^2}{\mu^2 r_\perp^2}, \label{eq:Hgq1}\\ {\cal H}^{(2)}_{gq}=&-4\pi N_c\frac{1}{\xi}{\cal P}_{gq}(\xi)\Bigg[e^{-i\frac{k_\perp}{\xi}\cdot(y_\perp-b_\perp)-ik_\perp\cdot r_\perp}\frac{(b_\perp-y_\perp)\cdot r_\perp}{(b_\perp-y_\perp)^2 r_\perp^2} +e^{-i\frac{k_\perp}{\xi}\cdot(b_\perp-y_\perp)-ik_\perp\cdot(x_\perp-b_\perp)}\frac{(b_\perp-x_\perp)\cdot(b_\perp-y_\perp)}{(b_\perp-x_\perp)^2(b_\perp-y_\perp)^2}\Bigg] \nonumber\\ & +\frac{4\pi}{N_c}\frac{1}{\xi^2}{\cal P}_{gq}(\xi)e^{-i\frac{k_\perp}{\xi}\cdot r_\perp-i\frac{k_\perp}{\xi}\cdot(y_\perp-b_\perp)}\frac{(b_\perp-y_\perp)\cdot r_\perp}{(b_\perp-y_\perp)^2 r_\perp^2}. \label{eq:Hgq2} \end{align} The last term in ${\cal H}^{(2)}_{gq}$ contributes to inclusive hadron production only in pp collisions but it is subleading compared to the other terms when we take the large-$N_c$. In the momentum space, one obtain easily \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/g+X}_{(gq)}}{d^2p_{h\perp}dy} =\sum_f\frac{\alpha_s}{2\pi}\int_\tau^1\frac{dz}{z^2}D_{h/g}(z)\int_{x_p}^1\frac{dx}{x} \xi xq_f(x)S_{gq} \label{eq:gq-final-momentum} \end{align} with \begin{align} S_{gq}= &\;\frac{C_F}{\pi}\frac{1}{\xi^2}{\cal P}_{gq}(\xi)\int \frac{d^2k_{g\perp}}{k_{g\perp}^2}\left[{\cal F}_{x_g}\left(\frac{k_\perp}{\xi}+k_{g\perp}\right)-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}\left(\frac{k_\perp}{\xi}\right)\right] \nonumber\\ & +\frac{N_c}{\pi}{\cal P}_{gq}(\xi)\int \frac{d^2k_{g\perp}}{k_{g\perp}^2}\left[{\cal F}_{x_g}\left(k_\perp+k_{g\perp}\right)-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}\left(k_\perp\right)\right]\nonumber\\ &-\frac{N_c}{\pi}{\cal P}_{gq}(\xi)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\left[\frac{k_\perp\cdot(k_\perp-k_{g\perp})}{k_\perp^2(k_\perp-k_{g\perp})^2}+\frac{(k_\perp-\xi k_{g\perp})\cdot(k_\perp-k_{g\perp})}{(k_\perp-\xi k_{g\perp})^2(k_\perp-k_{g\perp})^2}\right] \nonumber\\ & +\frac{1}{N_c\pi}{\cal P}_{gq}(\xi)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\frac{k_\perp\cdot(k_\perp-\xi k_{g\perp})}{k_\perp^2(k_\perp-\xi k_{g\perp})^2}. \end{align} \subsection{The $g\rightarrow q$ channel} \begin{figure} \centering \includegraphics[width=16cm,angle=0]{gq-real.eps} \caption{Real diagrams at NLO for the $g\rightarrow q$ channel. } \label{fig:NLO-gq-real} \end{figure} At last, we consider the $g\rightarrow q$ channel as shown in FIG.~\ref{fig:NLO-gq-real} which has no virtual corrections either. The differential cross section of $g+{\rm p}\rightarrow q+\bar q+X$ producing a quark with momentum $k$ and an anti-quark with momentum $l$ is \begin{align} \frac{d\sigma^{g+{\rm p}\rightarrow q+\bar q+X}}{d^3kd^3l}=\;&\alpha_s\delta(p^+-k^+-l^+)\int\frac{d^2x_\perp d^2x_\perp^\prime d^2b_\perp d^2b_\perp^\prime}{(2\pi)^8}e^{-ik_{\perp}\cdot(x_\perp-x_\perp^\prime)}e^{-il_{\perp}\cdot(b_\perp-b_\perp^\prime)}\sum_{\alpha,\beta,\lambda}\psi_{q\bar q\alpha\beta}^{\lambda\ast}(u_\perp^\prime)\psi_{q\bar q\alpha\beta}^{\lambda}(u_\perp)\nonumber\\ &\times\Bigg[T_R\widetilde{S}_{x_g}(v_\perp,v_\perp^\prime)-\frac{T_R}{2}\widetilde{S}_{x_g}(v_\perp,x_\perp^\prime)-\frac{T_R}{2}\widetilde{S}_{x_g}(v_\perp,b_\perp^\prime)-\frac{T_R}{2}\widetilde{S}_{x_g}(x_\perp,v_\perp^\prime)-\frac{T_R}{2}\widetilde{S}_{x_g}(b_\perp,v_\perp^\prime)\nonumber\\ &+\frac{T_RC_F}{N_c}\widetilde{S}_{x_g}(x_\perp,x_\perp^\prime)+\frac{T_RC_F}{N_c}\widetilde{S}_{x_g}(b_\perp,b_\perp^\prime)+\frac{T_R}{2N_c^2}\widetilde{S}_{x_g}(b_\perp,x_\perp^\prime)+\frac{T_R}{2N_c^2}\widetilde{S}_{x_g}(x_\perp,b_\perp^\prime)\Bigg], \end{align} where the splitting kernel is given in Eq.~(\ref{eq:splitting-kernel-qqbar}). By integrating over the phase space of the anti-quark and convoluting with the gluon PDF and the quark FF, one can cast the differential cross section in the momentum space into \begin{align} &\frac{\alpha_s}{2\pi^2}T_R\int_\tau^1\frac{dz}{z^2}D_{h/q}(z)\int_{\frac{\tau}{z}}^1d\xi {\cal P}_{qg}(\xi)\frac{x_p}{\xi}G\left(\frac{x_p}{\xi}\right)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\Bigg[\frac{2N_c^2}{d_A}\frac{1}{(k_\perp-\xi k_{g\perp})^2}+\frac{1}{(k_\perp-k_{g\perp})^2}\nonumber\\ &-\frac{2N_c^2}{d_A}\frac{(k_\perp-k_{g\perp})\cdot(k_\perp-\xi k_{g\perp})}{(k_\perp-k_{g\perp})^2(k_\perp-\xi k_{g\perp})^2}-\frac{2N_c^2}{d_A}\frac{k_\perp\cdot(k_\perp-\xi k_{g\perp})}{k_\perp^2(k_\perp-\xi k_{g\perp})^2}+\frac{2}{d_A}\frac{k_\perp\cdot(k_\perp-k_{g\perp})}{k_\perp^2(k_\perp-k_{g\perp})^2}\Bigg], \label{eq:qg-momentum} \end{align} where $d_A=N_c^2-1$. The first two terms in the square bracket correspond to the $q\bar q$ splitting after and before the incoming gluon scattering off the target, respectively. These terms contain the collinear divergences which can be subtracted from Eq.~(\ref{eq:qg-momentum}) and put into the definition of the quark PDF and the gluon FF as follows \begin{align} q_f(x,\mu)&=q_f^{(0)}(x)-\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int_x^1\frac{d\xi}{\xi}T_R{\cal P}_{qg}(\xi)G\left(\frac{x}{\xi}\right),\\ D_{h/g}(z,\mu)&=D_{h/g}^{(0)}(z)-\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int_z^1\frac{d\xi}{\xi}T_R{\cal P}_{qg}(\xi)D_{h/q}\left(\frac{z}{\xi}\right) \end{align} with the LO splitting function \begin{align} {\cal P}_{qg}(\xi)=(1-\xi)^2+\xi^2. \label{eq:splitting-function-qg} \end{align} For this channel, caution should be taken when the dimensional regularization is used, since the number of gluon polarization is shifted to $2(1-\epsilon)$. Therefore, the average of gluon polarization yields $\frac{1}{2(1-\epsilon)}$, and thus $-1/\hat\epsilon$ is replaced with $-1/\hat\epsilon-1$ in Eqs.~(\ref{eq:real-collinear-identity1})--(\ref{eq:virtual-collinear-identity2}) eventually. Following the same procedure as shown above, it is straightforward to obtain the differential cross section in the coordinate space \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/q+X}_{(qg)}}{d^2p_{h\perp}dy} =&\;\frac{\alpha_s}{2\pi}\int_\tau^1\frac{dz}{z^2}D_{h/q}(z)\int_{x_p}^1\frac{dx}{x} \xi xG\left(x\right)\int\frac{d^2 x_\perp d^2 y_\perp}{(2\pi)^2}(-T_{x_g}(x_\perp,y_\perp))\left[{\cal H}^{(1)}_{qg}+\int\frac{d^2 b_\perp}{(2\pi)^2}{\cal H}^{(2)}_{qg}\right], \label{eq:qg-final-coordinate} \end{align} with the finite hard parts \begin{align} {\cal H}^{(1)}_{qg}=&\;T_R {\cal P}_{qg}(\xi)e^{-ik_{\perp}\cdot r_\perp}\left[\ln\frac{c_0^2}{\mu^2 r_\perp^2}-1\right] +\frac{1}{\xi^2}\frac{T_R N_c}{C_F}{\cal P}_{qg}(\xi)e^{-i\frac{k_{\perp}}{\xi}\cdot r_\perp}\left[\ln\frac{c_0^2}{\mu^2 r_\perp^2}-1\right], \label{eq:Hqg1}\\ {\cal H}^{(2)}_{qg}=&-\frac{4\pi}{\xi}\frac{T_RN_c}{C_F}{\cal P}_{qg}(\xi)\Bigg[e^{-i\frac{k_{\perp}}{\xi}\cdot(y_\perp-b_\perp)-ik_{\perp}\cdot(b_\perp-x_\perp)}\frac{(x_\perp-b_\perp)\cdot(y_\perp-b_\perp)}{(x_\perp-b_\perp)^2(y_\perp-b_\perp)^2}+e^{-i\frac{k_{\perp}}{\xi}\cdot r_\perp-ik_{\perp}\cdot(y_\perp-b_\perp)}\frac{(b_\perp-y_\perp)\cdot r_\perp}{(b_\perp-y_\perp)^2 r_\perp^2}\Bigg]\nonumber\\ &+4\pi \frac{T_R}{N_cC_F}{\cal P}_{qg}(\xi)e^{-ik_{\perp}\cdot r_\perp-ik_{\perp}\cdot(y_\perp-b_\perp)}\frac{(b_\perp-y_\perp)\cdot r_\perp}{(b_\perp-y_\perp)^2 r_\perp^2}. \label{eq:Hqg2} \end{align} The last term in ${\cal H}^{(2)}_{qg}$ is subleading $N_c$ contribution. In the momentum space, Eq.~(\ref{eq:qg-final-coordinate}) becomes \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h/q+X}_{(qg)}}{d^2p_{h\perp}dy} =\frac{\alpha_s}{2\pi}\int_\tau^1\frac{dz}{z^2}D_{h/q}(z)\int_{x_p}^1\frac{dx}{x} \xi xG\left(x\right)S_{qg}, \label{eq:qg-final-momentum} \end{align} with \begin{align} S_{qg}&=T_R{\cal P}_{qg}(\xi)\left\{\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}^2}\left[{\cal F}_{x_g}(k_{\perp}+k_{g\perp})-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}(k_{\perp})\right]-{\cal F}_{x_g}(k_{\perp})\right\}\nonumber\\ &+\frac{1}{\xi^2}\frac{T_RN_c}{C_F}{\cal P}_{qg}(\xi)\left\{\frac{1}{\pi}\int\frac{d^2k_{g\perp}}{k_{g\perp}^2}\left[{\cal F}_{x_g}\left(\frac{k_{\perp}}{\xi}+k_{g\perp}\right)-J_0\left(\frac{c_0}{\mu}k_{g\perp}\right){\cal F}_{x_g}\left(\frac{k_{\perp}}{\xi}\right)\right]-{\cal F}_{x_g}\left(\frac{k_{\perp}}{\xi}\right)\right\}\nonumber\\ &-\frac{T_RN_c}{\pi C_F}{\cal P}_{qg}(\xi)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\left[\frac{(k_{\perp}-k_{g\perp})\cdot(k_{\perp}-\xi k_{g\perp})}{(k_{\perp}-k_{g\perp})^2(k_{\perp}-\xi k_{g\perp})^2}+\frac{k_{\perp}\cdot(k_{\perp}-\xi k_{g\perp})}{k_{\perp}^2(k_{\perp}-\xi k_{g\perp})^2}\right]\nonumber\\ &+\frac{T_R}{\pi N_cC_F}{\cal P}_{qg}(\xi)\int d^2k_{g\perp}{\cal F}_{x_g}(k_{g\perp})\frac{k_{\perp}\cdot(k_{\perp}-k_{g\perp})}{k_{\perp}^2(k_{\perp}-k_{g\perp})^2}. \end{align} \section{Summary} In this paper, we have calculated the inclusive hadron production at forward rapidity in pp collisions in the dilute regime with finite $N_c$. Using the dimensional regularization with the $\overline{\rm MS}$ scheme, we have shown that the collinear divergences and the rapidity divergences can be separated from the hard scattering parts and renormalized into the PDFs, the FFs, or the wave function of the target proton. Here, let us summarize all of the results by adding up the LO differential cross section and all four channels of the NLO corrections together. In the momentum space, the master formula can be written as \begin{align} \frac{d\sigma^{{\rm p}+{\rm p}\rightarrow h+X}}{d^2p_{h\perp}dy} =\sum_f\int_{\tau}^1\frac{dz}{z^2}\int_{x_p}^1\frac{dx}{x}\xi \left(xq_f\left(x,\mu\right),xG\left(x,\mu\right)\right) \begin{pmatrix} S^{(0)}_{qq}+\frac{\alpha_s}{2\pi}S_{qq} & \frac{\alpha_s}{2\pi}S_{gq}\\ \frac{\alpha_s}{2\pi}S_{qg} & S^{(0)}_{gg}+\frac{\alpha_s}{2\pi}S_{gg} \end{pmatrix} \begin{pmatrix} D_{h/q}(z,\mu)\\ D_{h/g}(z,\mu) \end{pmatrix} \label{eq:master-expression} \end{align} where $S_{qq}^{(0)}=\delta(1-\xi){\cal F}_{x_g}(k_\perp)$ and $S_{gg}^{(0)}=\delta(1-\xi)\widetilde{\cal F}_{x_g}(k_\perp)$. As shown in the Appendix~\ref{AppendixB}, basically, the hard coefficients for all four channels derived in this paper are equivalent to those obtained in p$A$ collisions in Ref.~\cite{Chirilli:2012jd} after taking the dilute limit as well as the large $N_c$ limit. The PDFs depend on the factorization scale and obey the DGLAP evolution equation \begin{align} \begin{pmatrix} q_f(x,\mu)\\ G(x,\mu) \end{pmatrix} = \begin{pmatrix} q_f^{(0)}(x)\\ G^{(0)}(x) \end{pmatrix} -\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int_{z}^1\frac{d\xi}{\xi} \begin{pmatrix} C_F{\cal P}_{qq}(\xi)&T_R{\cal P}_{qg}(\xi)\\ \sum_fC_F{\cal P}_{gq}(\xi)&N_c{\cal P}_{gg}(\xi) \end{pmatrix} \begin{pmatrix} q_f(x/\xi)\\ G(x/\xi) \end{pmatrix} . \label{eq:DGLAP-PDF} \end{align} Similarly, the FFs obey \begin{align} \begin{pmatrix} D_{h/q}(z,\mu)\\ D_{h/g}(z,\mu) \end{pmatrix} = \begin{pmatrix} D^{(0)}_{h/q}(z)\\ D^{(0)}_{h/g}(z) \end{pmatrix} -\frac{1}{\hat\epsilon}\frac{\alpha_s(\mu)}{2\pi}\int_{z}^1\frac{d\xi}{\xi} \begin{pmatrix} C_F{\cal P}_{qq}(\xi)&C_F{\cal P}_{gq}(\xi)\\ \sum_fT_R{\cal P}_{qg}(\xi)&N_c{\cal P}_{gg}(\xi) \end{pmatrix} \begin{pmatrix} D_{h/q}(z/\xi)\\ D_{h/g}(z/\xi) \end{pmatrix} . \label{eq:DGLAP-FF} \end{align} The rapidity divergences can be dealt with the BFKL equation. In the leading logarithmic approximation, the BFKL equation can be written as \begin{align} \frac{\partial T_{x_g}(r_\perp)}{\partial Y_g}= \frac{\overline{\alpha}_s}{2\pi}\int d^2b_\perp\frac{r_\perp^2}{(r_\perp-b_\perp)^2b_\perp^2} \left[T_{x_g}(r_\perp-b_\perp)+T_{x_g}(b_\perp)-T_{x_g}(r_\perp)\right], \end{align} with $\overline{\alpha}_s=\alpha_s N_c/\pi$. The factorization scale $\mu$ dependence is largely cancelled between the the hard parts and the PDFs and the FFs. Putting in the one-loop running coupling constant in the hard parts and using the NLO DGLAP equation and the NLO BFKL equation, Eq.~(\ref{eq:master-expression}) provides the complete results at NLO for inclusive hadron production in pp collisions in the dilute regime. It would be very interesting to evaluate Eq.~(\ref{eq:master-expression}) numerically for pp and p$A$ collisions and compare with the results obtained from the non-linear formulation~\cite{Watanabe:2015tja}. As to the initial condition for the BFKL evolution, we could use either use the dipole-dipole scattering amplitude, or models that we used for the heavy nuclei such as the McLerran-Venugopalan~\cite{McLerran:1993ni} or GBW model~\cite{GolecBiernat:1998js} in the dilute regime. Nevertheless, once the same initial conditions are set for both the BFKL equation and the BK equation, we anticipate that Eq.~(\ref{eq:master-expression}) can reveal a difference between the small-$x$ formalism in pp and p$A$ collisions for hadron spectra in the low transverse momentum region. This interesting comparison can provides us the precise and important information about the role of the non-linear gluon dynamics in these collisions, and thus help us observe the onset the saturation effects. We leave this issue for future study. \section*{Acknowledgements} The authors are grateful to A.~H.~Mueller and F.~Yuan for useful discussions. This work is supported by the NSFC under Grant No.~11575070.
a74d3aec36ad333b27640de97ae746c0dab0fb31
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Toroidal Alfvén eigenmodes (TAEs) are discrete frequency MHD waves that exist in the toroidicity induced gap of the Alfvén continuum spectra in toroidal magnetized plasmas. TAEs are typically excited by an ensemble of energetic ions (e.g.\ coming from auxiliary heating or from fusion reactions) with an inverted energy distribution along the characteristic curves of wave--particle interaction in momentum space. If these waves are excited to large amplitudes, they might eject a large fraction of energetic ions from the plasma before the ions transfer their energy to the bulk plasma, causing a significant reduction of heating efficiency of fast ions~\cite{hei,won}. It is therefore of great importance to understand the significance of TAEs in future devices, such as ITER. Accurate modeling is required that can resolve the nonlinear evolution of the wave--particle interactions. Many hybrid MHD--kinetic Monte Carlo codes developed for this purpose are orbit following, requiring temporal resolutions well below bounce time scales of the resonant energetic particles. These are many orders of magnitude shorter than the time scales for the relevant dynamics of long-lived eigenmodes in large tokamaks, such as ITER. Relatively simple equations of motion that can resolve the relevant time scales for wave--particle interaction more efficiently can be acquired by using action--angle coordinates~\cite{kau} of the equilibrium system for the phase space of energetic particles. Particles in the unperturbed equilibrium system then follow straight lines in configuration space at constant velocities, and their canonical momenta are constants of motion. This coordinate system has the advantage that particles exactly remain on their guiding center orbits in the unperturbed system, independently of the time step length. In order to get satisfactory convergence of the Monte Carlo codes describing nonlinear wave--particle interactions, $\delta f$ methods are often used. Although the method is computationally advantageous, it makes the code more difficult to use in conjunction with existing Monte Carlo codes that use full-$f$ methods or that use a different background distribution. FOXTAIL (``\underline{FO}urier series e\underline{X}pansion of fas\underline{T} particle--\underline{A}lfvén eigenmode \underline{I}nteraction''-mode\underline{L}) is a new hybrid magnetohydrodynamic--kinetic model that both uses action--angle coordinates for particle phase space and a full-$f$ Monte Carlo method to represent the resonant energetic particle distribution. It is based on a model by Berk \emph{et al.}~\cite{bb1}, which is derived from a Lagrangian formulation of the wave--particle interaction. The use of action--angle variables can give scenarios where the shortest time scale needed to be resolved in FOXTAIL is on the order of $\omega_\mathrm{p}^{-1}$, where $\omega_\mathrm{p}$ is the precession frequency of the energetic particles interacting strongly with the eigenmodes. A simplification used in FOXTAIL is that the spatial structures of the eigenmode wave fields are taken to be constant in time. This limitation means that FOXTAIL is unable to model, e.g., energetic particle modes. There also exist scenarios where the time evolution of TAE eigenfunctions is of importance (see e.g.\ Ref.~\cite{wan}). Such scenarios can be modeled by other existing codes, not having the limitation of static eigenfunctions, such as the hybrid MHD--gyrokinetic code, HMGC~\cite{hm1,hm2,hm3}, and the gyrokinetic toroidal code, GTC~\cite{lin,zha,xia}. The structure of this paper is the following: Section~\ref{sec:mod} presents the mathematical, the physical and the technical background of FOXTAIL, including derivations of the equations used in the different parts of the code and the overall structure of the code. Section~\ref{sec:coi} describes how some of the resulting equations are numerically implemented. Section~\ref{sec:num} presents some of the possible applications of FOXTAIL, including quantitative comparisons with the corresponding one-dimensional bump-on-tail approximation of the system and numerical studies of the Chirikov criterion in scenarios with two eigenmodes. Section~\ref{sec:con} summarizes the paper. \section{Model description}\label{sec:mod} \subsection{Physics overview} The equations used in FOXTAIL are based on a Lagrangian formulation of the wave--particle system \cite{bb1}. Each simulation is formulated as an initial value problem, starting from an energetic particle distribution in a background equilibrium plasma and a set of eigenmodes with a static spatial structure and a dynamical amplitude and phase. Eigenmodes are treated as weak perturbations of the equilibrium, excluding direct mode--mode interaction. The nonlinear coupling of eigenmodes is taken into account only by the energy and momentum exchange between the modes via the energetic particles. The physical process of the considered wave--particle interactions is essentially absorption and stimulated emission, and the interactions are energy and momentum conserving. The total energy of the eigenmode is proportional to the amplitude squared. Energy conservation is ensured by the equation \begin{equation}\label{eq:enc} \sum_\mathrm{particles}\dot{W}_\mathrm{wave} + \sum_\mathrm{eigenmodes}C\re(A \dot{A}^*) = 0, \end{equation} where $\dot{W}_\mathrm{wave}$ is the time derivative of the kinetic particle energy, as accelerated by the wave field, $A$ is the complex amplitude of the eigenmode, and $C$ is the ratio between the eigenmode energy and $|A|^2/2$. A Lagrangian formulation of the wave--particle interaction is presented in section~\ref{sec:cse}, which is shown to be consistent with eq.~\eqref{eq:enc}. The acceleration of a particle by an Alfvén eigenmode convects the momentum of the particles along curves in the phase space $(W, P_\phi, \mu)$ according to \begin{equation}\label{eq:cha} \frac{\mathrm{d} W}{\omega} = \frac{\mathrm{d} P_\phi}{n},~\mathrm{d}\mu = 0, \end{equation} where $W$ is the particle kinetic energy, $P_\phi$ is the toroidal canonical momentum, $\mu$ is the magnetic moment, $n$ is the toroidal mode number of the eigenmode and $\omega$ is the eigenmode frequency. The magnetic moment is unperturbed, since the Alfvén eigenmodes are low frequency waves ($\omega \ll \omega_\mathrm{c}$). The curves in $W,P_\phi$-space specified by eq.~\eqref{eq:cha} are referred to as the characteristic curves of wave--particle interaction. Within certain parameter limits, FOXTAIL is equivalent with a one-dimensional bump-on-tail model describing the wave--particle interaction (cf.\ Ref.~\cite{bb2}). Different characteristic curves are indistinguishable in this limit, and the momentum of the 1D model then corresponds to a variable indexing the location of the particle along the characteristic curves. From the theory of the 1D bump-on-tail model, it is apparent that an inverted energy distribution along the characteristic curves at the wave--particle resonance can excite eigenmodes via a process analogous to Landau damping. \begin{figure}[t!] \centering \includegraphics[width=75mm]{Fig1.pdf} \caption{Flowchart of the FOXTAIL code. As input, FOX takes the mass and charge of the energetic particle species and a grid in the space spanned by the adiabatic invariants of the equilibrium motion. On this grid, the orbits are solved using the input equilibrium configuration. External routines are used to calculate spatial wave field structures and frequencies of a chosen set of Alfvén eigenmodes. Eigenmode data and orbit data is sent to a routine that integrates all guiding center orbits along the wave fields of the eigenmodes in order to obtain a set of interaction coefficients characterizing the particle response with respect to the wave. A set of particle, eigenmode and wave--particle interaction data is collected and sent to the TAIL code, along with an initial distribution of markers and initial complex amplitudes of the eigenmodes. TAIL then solves the nonlinear time evolution of the markers and the considered eigenmodes.}\label{fig:fch} \end{figure} \subsection{Overview of the FOXTAIL code} The FOXTAIL code is essentially split into two parts, ``FOX'' and ``TAIL'', as illustrated in Fig.~\ref{fig:fch}. TAIL is the numerical dynamics solver of the eigenmodes and the energetic particles, solving the wave--particle system as an initial value problem. FOX can be viewed as a preprocessor of TAIL, calculating orbital, eigenmode and interaction related data for a given set of eigenmodes and energetic particle species in a defined equilibrium configuration. The distribution of energetic particles is represented using a finite set of markers with predefined weights. \subsection{Action--angle coordinates}\label{sec:aac} The phase space of markers in TAIL is described using action--angle coordinates of the equilibrium system \cite{kau}. The equations of motion of the system become particularly simple in this choice of coordinates, which contributes to the computational efficiency of the solver. In the absence of wave field perturbations, the ``action'' coordinates (i.e.\ the momentum coordinates of the canonical action--angle coordinate system) are constants of motion, whereas the ``angle'' coordinates (configuration space coordinates) evolve with constant velocities. The presence of eigenmodes perturbs this simple dynamics, and the adiabatic invariants are convected according to eq.~\eqref{eq:cha}. The angle coordinates of the system are given by \begin{equation}\label{eq:ang}\left\{\begin{array}{*3{>{\displaystyle\vspace{1mm}}l}} \tilde{\alpha} = \alpha + \int_0^\theta\mathrm{d}\theta'\frac{\bar{\omega}_\mathrm{c}(\bm{P}) - \omega_\mathrm{c}(\bm{P}, \theta')}{\dot{\theta}(\bm{P}, \theta')}, \\ \tilde{\theta} = \int_0^\theta\mathrm{d}\theta'\frac{\omega_\mathrm{B}(\bm{P})}{\dot{\theta}(\bm{P}, \theta')}, \\ \tilde{\phi} = \phi + \int_0^{\theta}\mathrm{d}\theta'\frac{\omega_\mathrm{p}(\bm{P}) - \dot{\phi}(\bm{P}, \theta')}{\dot{\theta}(\bm{P}, \theta')}, \end{array}\right.\end{equation} where $\alpha$ is the gyro-angle, $\theta$ and $\phi$ are the poloidal and toroidal angles, respectively, $\bm{P} \equiv (P_\alpha, P_\theta, P_\phi)$ are the action coordinates, canonical to the angles $(\tilde{\alpha}, \tilde{\theta}, \tilde{\phi})$. Furthermore, $\omega_\mathrm{c}$ ($\bar{\omega}_\mathrm{c}$) is the (time averaged) gyro-frequency and $\omega_\mathrm{B}$ and $\omega_\mathrm{p}$ are the bounce and precession frequencies, respectively. The integrals of eq.~\eqref{eq:ang} are evaluated on the poloidal coordinates along the guiding center orbits. FOXTAIL uses the adiabatic invariants $\bm{J} \equiv (\mu, \Lambda, P_\phi)$ as momentum coordinates ($\Lambda = \mu B_0/W$ is the normalized magnetic moment, where $B_0$ is the on-axis magnetic field strength). These momentum coordinates can be expressed as functions of $\bm{P}$. The angle coordinates $(\tilde{\alpha}, \tilde{\theta}, \tilde{\phi})$ are referred to as the \emph{transformed} gyro-angle, poloidal angle and toroidal angle, respectively. In the equilibrium system, where the momentum coordinates $\bm{J}$ are constant, the transformed angles evolve at a constant velocity $(\bar{\omega}_\mathrm{c}, \omega_\mathrm{B}, \omega_\mathrm{p})$. The dynamics in the FOXTAIL model is averaged over gyration time scales, and consequently $\tilde{\alpha}$ is an ignorable coordinate of the system. The transformed poloidal angle, $\tilde{\theta}$, can be viewed as an index of the location of the particle in the guiding center orbit, where $\tilde{\theta}: 0 \to 2\pi$ is a complete period of the guiding center orbit in the poloidal plane ($\tilde{\theta} = 0$ is defined as the point where the outer leg of the orbit intersects with the equatorial plane). In section~\ref{sec:wav}, it is shown that a Fourier expansion of the instant acceleration of the particle in the wave field is a convenient representation of the wave--particle interactions. Complications arise for the $\tilde{\theta}$ coordinate close to the boundary $\omega_\mathrm{B} = 0$, where the particle asymptotically approaches one of the turning points. In this limit, all points along the orbit besides the turning points are represented by infinitely narrow intervals in $\tilde{\theta}$, and the representation of the wave--particle interaction using Fourier series expansions in $\tilde{\theta}$-space becomes invalid. These complications can potentially be resolved either by ad hoc boundary conditions or by more sophisticated coordinate transformations close to this boundary. However, all of the numerical studies presented in this paper consider scenarios where the simulated energetic particle distributions are sufficiently far from the boundary. \subsection{Orbit solver} FOX contains subroutines that solves the 3D motion of particles in a given equilibrium field configuration on a grid in $\bm{J}$-space. The equilibrium configuration and the particle orbits are described in $\psi,\theta,\phi$-space, where $\psi$ is the poloidal magnetic flux per radian. For each orbit on the $\bm{J}$-grid, the time evolution of $\psi$, $\theta$ and $\phi_\mathrm{s}$ is calculated, where $\phi_\mathrm{s} \equiv \phi - \tilde{\phi} + \omega_\mathrm{p}\tilde{\theta}/\omega_\mathrm{B}$ is the shifted toroidal coordinate ($\phi_\mathrm{s}$ is $\phi$ shifted such that $\phi_\mathrm{s} = 0$ coincides with $\tilde{\theta} = 0$). The equilibrium configuration is taken to be axisymmetric. Assuming MHD force balance and nested magnetic flux surfaces, the equilibrium magnetic field can be expressed as \begin{equation} \bm{B} = F(\psi)\nabla\phi + \nabla\phi\times\nabla\psi. \end{equation} The guiding center motion consists of a parallel motion and a drift motion, where the drift is given by the combined $\bm{E}\times\bm{B}$, $\nabla B$ and curvature drifts according to \begin{equation} \bm{v}_\mathrm{d} = \frac{\bm{E}\times\bm{B}}{B^2} + \frac{\mu(2 B_0 - \Lambda B)}{Z e \Lambda B^3}\bm{B}\times\nabla B. \end{equation} By combining $W = m v_\parallel^2/2 + \mu B$ with $P_\phi = m v_\parallel F/B + m v_{\mathrm{d},\phi} - Z e\psi$ and eliminating $v_\parallel$, it can be shown that the coordinates of the guiding center orbits in the poloidal plane follow the equation \begin{equation}\label{eq:gco} f(\psi, \theta, \bm{J}) = 0, \end{equation} where \ba f(\psi, \theta, \bm{J}) &\equiv 1 - \frac{\Lambda B(\psi, \theta)}{B_0} - \frac{\Lambda B^2(\psi, \theta)}{2 m \mu B_0 F^2(\psi)} \nonumber\\ &\quad\, \times [P_\phi + Z e\psi - m v_{\mathrm{d}, \phi}(\psi, \theta, \mu, \Lambda)]^2, \ea \ba v_{\mathrm{d}, \phi}(\psi, \theta, \mu, \Lambda) &= \frac{E^\psi(\psi, \theta)}{B^2(\psi, \theta)} - \frac{\mu[2 B_0 - \Lambda B(\psi, \theta)]}{Z e \Lambda B^3(\psi, \theta)}\nonumber\\ &\quad\, \times[\nabla B(\psi, \theta)]^\psi \ea is the covariant toroidal component of the drift velocity, $E^\psi$ and $[\nabla B]^\psi$ are the contravariant $\psi$-components of $\bm{E}$ and $\nabla B$, respectively, and $m$ is the particle mass. Once the projections of the orbits on the poloidal plane are solved on the chosen $\bm{J}$-grid using eq.~\eqref{eq:gco}, the corresponding time coordinates are calculated according to \begin{equation}\label{eq:tps} t(\psi) = \int_{\psi_0}^\psi\frac{\mathrm{d}\psi}{\dot{\psi}}, \end{equation} where \begin{equation}\label{eq:dps} \dot{\psi} = \frac{J F E_\theta - g_{\theta\theta}E_\phi}{J^2 B^2} - \frac{\mu F(2 B_0 - \Lambda B)}{Z e \Lambda J B^3}\dd{B}{\theta}, \end{equation}\begin{equation} J \equiv \left(\dd{\bm{r}}{\psi}\times\dd{\bm{r}}{\theta}\right)\cdot\dd{\bm{r}}{\phi}, \end{equation} and $\psi(t = 0) = \psi_0$ is defined as the point where the outer leg of the guiding center orbit intersects the equatorial plane ($\tilde{\theta} = 0$). Similarly, the $\phi_\mathrm{s}$ coordinates are calculated from the toroidal velocity according to \begin{equation}\label{eq:tph} \phi_\mathrm{s}(t) = \int_0^t\mathrm{d} t'\:\dot{\phi}(\bm{J}, t'), \end{equation} where \begin{equation}\label{eq:dph} \dot{\phi}(\bm{J}, t) = \frac{P_\phi + Z e\psi(\bm{J}, t)}{m R^2(\bm{J}, t)}. \end{equation} When calculating the wave--particle interaction coefficients, the poloidal velocity is required at each point of the orbit (see eq.~\eqref{eq:vig}). It is given by \ba\label{eq:dth} \dot{\theta} &= \frac{P_\phi + Z e\psi - m v_{\mathrm{d},\phi}}{m J F} - \frac{J F E_\psi + g_{\psi\theta}E_\phi}{J^2 B^2} \nonumber\\ &\quad\, + \frac{\mu F(2 B_0 - \Lambda B)}{Z e \Lambda J B^3}\dd{B}{\psi}. \ea \subsection{Wave field description}\label{sec:wav} The present version of FOXTAIL describes the dynamics of low frequency, \emph{shear} eigenmodes, such as the toroidal Alfvén eigenmodes (TAEs), but the model can be extended to describe eigenmodes with, e.g., compressible components as well. Neglecting plasma resistivity, the general electric wave field can be represented by the two scalar potentials $\Phi$ and $\Psi$ according to \begin{equation}\label{eq:esc} \delta\bm{E} = -\nabla_\perp\Phi + \frac{\bm{B}\times\nabla\Psi}{B}, \end{equation} where the first term is associated with magnetic shear, and the second term is associated with magnetic compression. When parallel gradients in the plasma are negligible in comparison to perpendicular gradients, the excitation of the two scalar potentials $\Phi$ and $\Psi$ is almost decoupled, and it is sufficient to describe the shear Alfvén wave using the first term \cite{bb1}. In an axisymmetric toroidal plasma, the scalar potential of each eigenmode (indexed by $i$) can be written on the form \ba \Phi_i(\bm{r}, t) &= \re\sum_m C_i(t)\mathrm{e}^{\mathrm{i}\chi_i(t)}\Phi_{i, m}(\psi) \nonumber\\ &\quad\, \times\mathrm{e}^{\mathrm{i}(n_i \phi - m\theta - \omega_i t)}, \label{eq:phi} \ea where $C_i$ and $\chi_i$ are the slowly varying amplitude and phase of the eigenmode, respectively ($\dot{C}_i/C_i \ll \omega_i$, $\dot{\chi_i} \ll \omega_i$), $n_i$ is the toroidal mode number and $\omega_i$ is the eigenmode frequency. The electric wave field therefore given by \begin{equation} \delta\bm{E} = \re\sum_i C_i\mathrm{e}^{\mathrm{i}\chi_i}\tilde{\bm{E}}_i\mathrm{e}^{\mathrm{i}(n_i\phi - \omega_i t)}, \end{equation} where \ba \tilde{\bm{E}}_i &= \sum_m\mathrm{e}^{-\mathrm{i} m\theta}\bigg(\left[\mathrm{i}\Phi_{i,m}g_{\psi\theta}G_{i,m} - \dr{\Phi_{i, m}}{\psi}\right]\nabla\psi + \mathrm{i}\Phi_{i,m} \nonumber\\ &\quad\,\times\Big[(g_{\theta\theta}G_{i,m} + m)\nabla\theta + (J F G_{i,m} - n_i)\nabla\phi\Big]\bigg), \label{eq:tie} \ea $G_{i,m} \equiv (n_i J F - m R^2)/(J^2 R^2 B^2)$. \subsection{Fourier series expansion of fast particle--Alfvén eigenmode interaction} The acceleration of the energetic particle in the wave field is described by the equation \begin{equation} \dot{W} = Z e \bm{v}\cdot\delta\bm{E} \approx Z e\langle\bm{v}\cdot\delta\bm{E}\rangle_\mathrm{g}, \end{equation} where $\langle\cdot\rangle_\mathrm{g}$ averages over the gyro-motion. Initial versions of FOXTAIL only consider the lowest order averaging over the gyro-motion, but gyro-kinetic corrections to the averaging may be included in later versions. The averaged $\bm{v}\cdot\delta\bm{E}$ can be written as \begin{equation} \langle\bm{v}\cdot\delta\bm{E}\rangle_\mathrm{g} = \langle\bm{v}\rangle_\mathrm{g}\cdot\langle\delta\bm{E}\rangle_\mathrm{g} + \frac{\mu}{Z e}\dd{\langle\delta B_\parallel\rangle_{S_\mathrm{g}}}{t}, \end{equation} where $\langle\delta B_\parallel\rangle_{S_\mathrm{g}}$ averages the parallel magnetic wave field over the surface $S_\mathrm{g}$ enclosed by the gyro-ring (generated by the span of the gyro-angle. For shear waves, the $\delta B_\parallel$ term can be neglected, and we are left with \begin{equation} \dot{W} = Z e\langle\bm{v}\rangle_\mathrm{g}\cdot\langle\delta\bm{E}\rangle_\mathrm{g} = \re\sum_i C_i\mathrm{e}^{\mathrm{i}\chi_i} V_i \mathrm{e}^{\mathrm{i}(n_i\phi - \omega_i t)}, \end{equation} where \begin{equation}\label{eq:vig} V_i(\bm{J}, t) = Z e(\dot{\psi}\tilde{E}_\psi + \dot{\theta}\tilde{E}_\theta + \dot{\phi}\tilde{E}_\phi), \end{equation} $\dot{\psi}(\bm{J}, t)$, $\dot{\theta}(\bm{J}, t)$ and $\dot{\phi}(\bm{J}, t)$ are the guiding center velocities, given by eqs.~\eqref{eq:dps}, \eqref{eq:dth} and \eqref{eq:dph}, respectively, and $\tilde{E}_\psi(\psi(t), \theta(t))$, $\tilde{E}_\theta(\psi(t), \theta(t))$ and $\tilde{E}_\phi(\psi(t), \theta(t))$ are given by eq.~\eqref{eq:tie}. All guiding center coordinates $(\psi, \theta, \phi_\mathrm{s})$ and their velocities can be written as functions of $\bm{J}$ and $\tilde{\theta}$. It is then possible to write $\dot{W}$ on a very compact form using action--angle coordinates: \begin{equation}\label{eq:dwi} \dot{W} = \re\sum_{i,\ell}C_i\mathrm{e}^{\mathrm{i}\chi_i}V_{i,\ell}\mathrm{e}^{\mathrm{i}(\ell\tilde{\theta} + n_i\tilde{\phi} - \omega_i t)}, \end{equation} where \ba V_{i,\ell}(\bm{J}) &= \frac{Z e}{2\pi}\int_0^{2\pi}\mathrm{d}\tilde{\theta}\:V_i(\tilde{\theta}, \bm{J}) \nonumber\\ &\quad\, \times\exp\bigg[\mathrm{i}\bigg(n_i\left[\phi_\mathrm{s}(\tilde{\theta}, \bm{J}) - \frac{\omega_\mathrm{p}(\bm{J})}{\omega_\mathrm{B}(\bm{J})}\tilde{\theta}\right] - \ell\tilde{\theta}\bigg)\bigg]\label{eq:vil} \ea The coefficients $V_{i, \ell}$ are the Fourier series expansion of the wave--particle interaction that named the FOX code. It can be seen in eq.~\eqref{eq:dwi} that a wave--particle resonance is characterized by the condition $\ell\omega_\mathrm{B} + n_i\omega_\mathrm{p} \approx \omega_i$. The model is very efficient in the sense that one can select the relevant modes $i$ and coefficients $\ell$ that are close to resonant with an ensemble of energetic particles and neglect all other modes and non-resonant Fourier coefficients. If the relevant coefficients only have $\ell = 0$, $\tilde{\theta}$ becomes an ignorable coordinate of the system, and the shortest time scales that need to be resolved in simulations using TAIL are the precession time scales, $\omega_\mathrm{p}^{-1}$. As was mentioned in section~\ref{sec:aac}, complications arise for the choice of action--angle coordinates when describing the wave--particle interaction in regions where $\omega_\mathrm{B} \approx 0$. This is explicitly seen in eq.~\eqref{eq:vil}, where $\omega_\mathrm{p}/\omega_\mathrm{B} \to \infty$, and the integrand oscillates infinitely fast in $\tilde{\theta}$. $V_i(\tilde{\theta}, \bm{J})$ is bounded, and constant for all $\tilde{\theta}$ in this limit except for the infinitely narrow intervals that do not represent the turning points. For this reason, it can be understood that $V_{i,\ell}$ tends to zero towards the $\omega_\mathrm{B} = 0$ boundary surface. Note that the sum of eq.~\eqref{eq:dwi} over all integers $\ell \in \mathbb{Z}$ should remain finite, given that particles can be accelerated across these surfaces by wave fields. \subsection{Lagrangian formulation of wave--particle interaction}\label{sec:cse} A model for describing the dynamics of the momentum variables $(\mu, \Lambda, P_\phi)$ and the amplitudes and phases of the eigenmodes remains to be found. Such a model can be derived from a Lagrangian formulation of the wave--particle system. We consider additions to the equilibrium Lagrangian by power series of the eigenmode amplitude $C_i$. Expressed in action--angle variables, the zeroth order Lagrangian reads \cite{whi} \begin{equation}\label{eq:la0} \mathcal{L}_0 = \sum_k \left[P_{\alpha,k}\dot{\tilde{\alpha}}_k + P_{\theta,k}\dot{\tilde{\theta}}_k + P_{\phi, k}\dot{\tilde{\phi}}_k - \mathcal{H}_0(\bm{P}_k)\right], \end{equation} where $k$ is the particle index, and $\mathcal{H}_0$, satisfying \begin{equation} \dd{\mathcal{H}_0}{\bm{P}} = (\bar{\omega}_\mathrm{c}, \omega_\mathrm{B}, \omega_\mathrm{p}), \end{equation} is the equilibrium Hamiltonian. A first order Lagrangian consistent with eq.~\eqref{eq:dwi} is \begin{equation}\label{eq:la1} \mathcal{L}_1 = \im\sum_{i,k,\ell}\frac{C_i}{\omega_i}\mathrm{e}^{\mathrm{i}\chi_i}V_{i,\ell}(\bm{J}_k)\mathrm{e}^{\mathrm{i}(\ell\tilde{\theta}_k + n_i\tilde{\phi}_k - \omega_i t)}. \end{equation} In a low-$\beta$ plasma, the second order Lagrangian for shear Alfvén eigenmodes can be expressed on the form \cite{bb1} \begin{equation}\label{eq:l20} \mathcal{L}_2 = -\sum_i\frac{\dot{\chi}_i C_i^2}{2\mu_0\omega_i}\int\mathrm{d} V\:\frac{|\tilde{\bm{E}}_i(\bm{r})|^2}{v_\mathrm{A}^2(\bm{r})} \end{equation} when neglecting rapidly oscillating terms (on time scales $\omega_i^{-1}$), where $\mu_0$ is the vacuum permeability. Note that the system is invariant under the transformation $C_i\mathrm{e}^{\mathrm{i}\chi_i} \to \kappa_i C_i\mathrm{e}^{\mathrm{i}\chi_i}$, $\Phi_{i,m} \to \Phi_{i,m}/\kappa_i$ for any constant $\kappa_i$. A normalization can be imposed by the condition \begin{equation} \int\mathrm{d} V\:\frac{|\tilde{\bm{E}}_i(\bm{r})|^2}{\mu_0 v_\mathrm{A}^2(\bm{r})} = 1 \end{equation} for each mode $i$. Then the second order Lagrangian reduces to \begin{equation}\label{eq:la2} \mathcal{L}_2 = -\sum_i\frac{\dot{\chi}_i C_i^2}{2\omega_i}. \end{equation} Assuming that the amplitudes $C_i$ are small enough, such that $|\mathcal{L}_1| \ll |\mathcal{H}_0|$, the corresponding Hamiltonian for the wave--particle system is $\mathcal{H} = \mathcal{H}_0 - \mathcal{L}_1$, with the canonically conjugate pairs $(\tilde{\alpha}_k\,; P_{\alpha, k})$, $(\tilde{\theta}_k\,; P_{\theta, k})$, $(\tilde{\phi}_k\,; P_{\phi, k})$ and $(-\chi_i\,; C_i^2/2\omega_i)$. From this, one can derive all the remaining equations of motion of the wave--particle system self-consistently. Defining the complex amplitude $A_i \equiv C_i\mathrm{e}^{\mathrm{i}\chi_i}$ and using that $\mu = Z e P_\alpha / m$, the equations of motion of the wave--particle system is \begin{equation}\label{eq:ta1}\left\{\begin{array}{*3{>{\displaystyle\vspace{1mm}}l}} \dot{\mu}_k = 0, & \dot{\tilde{\theta}}_k = \omega_\mathrm{B}(\bm{J}_k), \\ \dot{\Lambda}_k = -\frac{\Lambda_k^2}{\mu_k B_0}\re\sum_i A_i U_{i,k}, & \dot{\tilde{\phi}}_k = \omega_\mathrm{p}(\bm{J}_k), \\ \dot{P}_{\phi,k} = \re\sum_i\frac{n_i}{\omega_i}A_i U_{i,k}, & \dot{A}_i = -\sum_k U_{i,k}^*, \end{array}\right.\end{equation} where \begin{equation} U_{i,k} \equiv \sum_\ell V_{i,\ell}(\bm{J}_k)\mathrm{e}^{\mathrm{i}(\ell\tilde{\theta}_k + n_i\tilde{\phi}_k - \omega_i t)}. \end{equation} \subsection{Additional operators}\label{sec:col} The effects of particle collisions are not covered by the theory presented in the preceding sections. These effects can be added explicitly to the system, e.g.\ by using a diffusion operator in momentum space \cite{bb3} combined with a momentum drag \cite{lil} derived from the Fokker--Planck operator, which act directly on the energetic particle distribution. Adding diffusion to the system transforms the system of ordinary differential equations in eq.~\eqref{eq:ta1} to a system of stochastic differential equations. The general drag--diffusion operator can be written on It\=o form according to \begin{equation}\label{eq:col}\left\{\begin{array}{*3{>{\displaystyle\vspace{1mm}}l}} \mathrm{d}\mu_\mathrm{c} = a_\mu\mathrm{d} t + \bm{b}_\mu\cdot\mathrm{d}\bm{W}_t, \\ \mathrm{d}\Lambda_\mathrm{c} = a_\Lambda\mathrm{d} t + \bm{b}_\Lambda\cdot\mathrm{d}\bm{W}_t, \\ \mathrm{d} P_{\phi,\mathrm{c}} = a_{P_\phi}\mathrm{d} t + \bm{b}_{P_\phi}\cdot\mathrm{d}\bm{W}_t, \end{array}\right.\end{equation} where $a_{\bm{J}}$ is associated with drag, $\bm{b}_{\bm{J}}$ is associated with diffusion and the components of $\bm{W}_t$ are independent Wiener processes in time. Both $a_{\bm{J}}$ and $\bm{b}_{\bm{J}}$ are functions of $\bm{J}$ and $\tilde{\theta}$ in general, but orbit averaged (i.e.\ $\tilde{\theta}$ averaged) versions of the operators may be used for simplicity (see e.g.\ Ref.~\cite{eri}). There are other processes external to the energetic particle--Alfvén eigenmode system which may be included. Depending on the source of the energetic particle distribution (e.g.\ neutral beam injection, cyclotron resonance heating or nuclear fusion), operators can be added that continuously supply particles and/or energy to the system. Particle sources are typically modeled by dynamical weights and statistical redistribution of markers. Losses from Bremsstrahlung and cyclotron radiation can be modeled by additions to the $a_{\bm{J}}$ terms in eq.~\eqref{eq:col}. An additional $-\gamma_{\mathrm{d},i} A_i$ term can be added to the amplitude equation ($\dot{A}_i$ in eq.~\eqref{eq:ta1}) in order to model various damping mechanisms on the eigenmode amplitudes, where $\gamma_{\mathrm{d}, i}$ is the damping rate of the $i$:th mode. This damping can, e.g., come from Landau damping in the interaction with thermal particles or damping due to mode conversion. The $-\gamma_{\mathrm{d},i} A_i$ term of the amplitude equation is here referred to as \emph{explicit} wave damping, unlike the Landau damping coming from the interaction with the energetic particle distribution, which arises implicitly from the model equations. \subsection{Lowest order corrections to the angle perturbation}\label{sec:loc} When going from the Lagrangians of eqs.~\eqref{eq:la0}, \eqref{eq:la1} and \eqref{eq:la2} to eq.~\eqref{eq:ta1}, it was assumed that $|\mathcal{L}_1| \ll |\mathcal{H}_0|$, which allowed one to neglect corrections of the canonical momenta coming from the implicit dependence of $\mathcal{L}_1$ with respect to $\dot{\tilde{\theta}}$ and $\dot{\tilde{\phi}}$. For a small enough $\mathcal{L}_1$, the final equations for $\dot{\tilde{\theta}}$ and $\dot{\tilde{\phi}}$ are given only by the derivatives of the equilibrium Hamiltonian $\mathcal{H}_0$ with respect to $P_\theta$ and $P_\phi$, respectively. However, the above assumptions might not be valid for large amplitude eigenmodes and for processes affecting the energetic particles or the eigenmodes that are not direct wave--particle interaction, and one may have to consider the lowest order correction to $\dot{\tilde{\theta}}$ and $\dot{\tilde{\phi}}$ due to momentum perturbation. The correction can be understood from the fact that $\tilde{\theta}$ maps differently to locations on the guiding center orbit for different $\bm{J}$. Without the correction, one pushes the particle forwards or backwards along the orbit due to different mapping of $\tilde{\theta}$. Assuming a minimal perturbation of the guiding center position in $R,Z$-space while perturbing $\bm{J}$ by the amount $\mathrm{d}\bm{J}$, it can be shown that $\tilde{\theta}$ should be corrected by the amount $\partial\tilde{\theta}/\partial\bm{J}\cdot\mathrm{d}\bm{J}$, where \begin{equation}\label{eq:dtj} \dd{\tilde{\theta}}{\bm{J}} = -\frac{\omega_\mathrm{B}}{\dot{R}^2 + \dot{Z}^2}\left(\dot{R}\dd{R}{\bm{J}} + \dot{Z}\dd{Z}{\bm{J}}\right), \end{equation} $R$ is the distance from the symmetry axis and $Z$ is the vertical guiding center position. Using that $\tilde{\phi} = \phi - \phi_\mathrm{s} + \omega_\mathrm{p}\tilde{\theta}/\omega_\mathrm{B}$, it can be shown that \ba\label{eq:dpj} \dd{\tilde{\phi}}{\bm{J}} &= \frac{\dot{\phi} - \omega_\mathrm{p}}{\dot{R}^2 + \dot{Z}^2}\left(\dot{R}\dd{R}{\bm{J}} + \dot{Z}\dd{Z}{\bm{J}}\right). \nonumber\\ &\quad\, + \dd{}{\bm{J}}\left(\frac{\omega_\mathrm{p}}{\omega_\mathrm{B}}\tilde{\theta} - \phi_\mathrm{s}\right). \ea All the derivatives with respect to $\bm{J}$ on the right hand sides of eqs.~\eqref{eq:dtj} and \eqref{eq:dpj} are evaluated while keeping $\tilde{\theta}$ constant. When the $\bm{J}$ perturbations of any process is stochastic ($\bm{b}_{\bm{J}}$ of eq.~\eqref{eq:col} is nonzero), the angle coordinates become stochastic as well when including the $\partial\tilde{\theta}/\partial\bm{J}$ and $\partial\tilde{\phi}/\partial\bm{J}$ corrections. This is typically how phase decorrelation is introduced to the system. Such an operator was analyzed for the one-dimensional bump-on-tail model in Ref.~\cite{tho}. \subsection{Summary of the TAIL model equations}\label{sec:sum} To summarize, the model equations used by TAIL are \begin{equation}\label{eq:ta2}\left\{\begin{array}{*3{>{\displaystyle\vspace{1mm}}l}} \mathrm{d}\tilde{\theta}_k = \omega_\mathrm{B}(\bm{J}_k)\mathrm{d} t + \dd{\tilde{\theta}(\tilde{\theta}_k, \bm{J}_k)}{\bm{J}}\cdot\mathrm{d}\bm{J}_k, \\ \mathrm{d}\tilde{\phi}_k = \omega_\mathrm{p}(\bm{J}_k)\mathrm{d} t + \dd{\tilde{\phi}(\tilde{\theta}_k, \bm{J}_k)}{\bm{J}}\cdot\mathrm{d}\bm{J}_k, \\ \mathrm{d} \mu_k = \mathrm{d}\mu_\mathrm{c}(\bm{J}_k), \\ \mathrm{d}\Lambda_k = -\frac{\Lambda_k^2}{\mu_k B_0}\re\sum_i A_i U_{i,k}\mathrm{d} t + \mathrm{d}\Lambda_\mathrm{c}(\bm{J}_k), \\ \dot{P}_{\phi,k} = \re\sum_i\frac{n_i}{\omega_i}A_i U_{i,k} + \mathrm{d} P_{\phi,\mathrm{c}}(\bm{J}_k), \\ \dot{A}_i = -\sum_k w_k U_{i,k}^* - \gamma_{\mathrm{d}, i} A_i, \end{array}\right.\end{equation} where $k$ is now the \emph{marker} index with weight $w_k$, and $(\mathrm{d}\mu_\mathrm{c}, \mathrm{d}\Lambda_\mathrm{c}, \mathrm{d} P_{\phi,\mathrm{c}})$ represents additional differential operators acting on the momentum space of markers. The total wave--particle energy of the system can be defined as \ba W_\mathrm{tot} &\equiv \sum_k w_k W_k + \sum_i\frac{|A_i|^2}{2} \nonumber\\ &= \sum_k w_k \frac{\mu_k B_0}{\Lambda_k} + \sum_i\frac{|A_i|^2}{2}. \ea In the absence of explicit wave damping and particle sources and sinks, it can easily be shown that \begin{equation} \dot{W}_\mathrm{tot} = -\sum_k\frac{w_k \mu_k B_0}{\Lambda_k^2}\dot{\Lambda}_k + \re\sum_i A_i\dot{A}_i^* = 0, \end{equation} which is consistent with the condition for energy conservation in eq.~\eqref{eq:enc}. \section{Code implementation}\label{sec:coi} As input, FOX takes a file that contains all information characterizing the equilibrium configuration on a format compatible with Integrated Tokamak Modelling standards \cite{imb}. All scalar fields, such as $B$, $J$ and $F$, are specified on a grid in $\psi,\theta$-space. The user defines an equidistant grid in $\bm{J}$-space, where all guiding center orbits are to be solved in the given equilibrium. Equation~\eqref{eq:gco} is then solved for each $\bm{J}$ on the grid by bilinear interpolation of $f$ in $\psi,\theta$-space. An example of such a solution is shown in Fig.~\ref{fig:orb}. The solution method is optimal for wide orbits, whereas thinner orbits require a large enough $\psi$ resolution of the equilibrium. Once the guiding center orbit points are found in the poloidal plane, the time dependence of the orbit is calculated using eq.~\eqref{eq:tps}. Numerically, the integration is performed by assuming $\ddot{\psi}$ to be constant between adjacent points of the guiding center orbit. This assumption generates the approximation \begin{equation}\label{eq:tpa} t_j = \int_{\psi_0}^{\psi_j}\frac{\mathrm{d}\psi}{\dot{\psi}} \approx 2\sum_{k = 1}^j\frac{\psi_k - \psi_{k - 1}}{\dot{\psi}_k + \dot{\psi}_{k - 1}}, \end{equation} where $(\psi_j, \theta_j)$ is the $j$:th $\psi,\theta$-point along the orbit and $\dot{\psi}_j$ is $\dot{\psi}$ evaluated at $(\psi_j, \theta_j)$. Equation~\eqref{eq:tpa} becomes singular at the points where $\dot{\psi}_j = -\dot{\psi}_{j - 1}$. Close to such a singularity, or when the $\psi,\theta$-grid is coarse, large or even non-monotonous time coordinates may result. When these events are identified,\footnote{``Large'' time steps are identified by FOX using the condition $|\dot{\psi}_j| > C|\psi_{j + 1} - \psi_j|/|t_{j + 1} - t_j|$, where the constant $C = 4$.} FOX successively eliminates points along the guiding center orbit until all time steps are small and monotonous. The $\phi_\mathrm{s}$ coordinates of the orbit are then solved simply by using the trapezoidal method on eq.~\eqref{eq:tph}. \begin{figure}[t] \centering \includegraphics[width=58mm]{Fig2.pdf} \caption{Guiding center orbit in the poloidal plane solved by FOX in an ITER equilibrium, using a deuterium ion with $\mu = \SI{4}{MeV/T}$, $\Lambda = 0.85$ and $P_\phi = \SI{5}{eVs}$ (\SI{1}{eV} = \SI{1.6022e-19}{J}).}\label{fig:orb} \end{figure} The eigenfunctions of the wave field, $\Phi_{i,m}(\psi)$ and the corresponding mode frequencies $\omega$ are solved by using the analytical approximations presented in Ref.~\cite{can}. Future versions of FOXTAIL intend to use the MHD eigenmode analyzer code MISHKA~\cite{mi1,mi2,mi3,mi4} to solve the TAEs numerically. Besides ideal MHD effects, MISHKA also considers effects from finite ion Larmor radii, ion and electron drift, neoclassical ion viscosity and bootstrap current, indirect energetic ion effects and the collisionless skin effect. The interaction coefficients $V_{i,\ell}$ are calculated from eq.~\eqref{eq:vil} on each point on the $\bm{J}$-grid also by using the trapezoidal method. The user specifies which mode--Fourier coefficient pairs to be calculated by FOX. All bounce frequencies, precession frequencies, interaction coefficients, mode frequencies and toroidal mode numbers are then collected in a single output file. A separate output file is generated that contains all initial conditions used in a specific TAIL simulation, which contains the initial energetic particle distribution, flags on the mode--Fourier coefficient pairs to be active in the simulation, initial mode amplitudes, etc. In the absence of collisions, the TAIL model equations of eq.~\eqref{eq:ta1} defines a system of ordinary differential equation, which is solved numerically using the standard 4$^\mathrm{th}$ order Runge-Kutta method. When momentum diffusion is present, the model equations instead become a system of stochastic differential equations, which can be modeled numerically, e.g.\ by using an It\=o--Taylor numerical scheme~\cite{klo}. \section{Numerical studies}\label{sec:num} \subsection{Comparison with the 1D bump-on-tail model}\label{sec:com} One of the possible applications of FOXTAIL is to determine whether a one-dimensional bump-on-tail approximation of the system is sufficient to capture the essential wave--particle dynamics, such as growth rate and saturation amplitude. A higher computational efficiency typically follows from the lower dimensionality of the approximation. However, the 1D bump-on-tail model is only valid within certain parameter regimes. Section~\ref{sec:com} presents a quantitative numerical study of these regimes. \subsubsection{Theoretical comparison} A simple bump-on-tail model, neglecting collisions, explicit wave damping and particle sources and sinks, can be written on the form~\cite{tho} \begin{equation}\label{eq:bot} \dr{\xi_k}{\tau} = u_k,~\dr{u_k}{\tau} = \re(\tilde{A}\mathrm{e}^{\mathrm{i}\xi_k}),~\dr{\tilde{A}}{\tau} = -\sum_k\mathrm{e}^{-\mathrm{i}\xi_k}, \end{equation} where $\tau$ is the time coordinate, $\xi$ is the particle phase, $u$ is the particle momentum and $\tilde{A}$ is the complex eigenmode amplitude. Now, consider the FOXTAIL model, only including a single mode $i$ and a single Fourier coefficient $\ell$. We define the particle phase as \begin{equation} \xi_k \equiv \ell\tilde{\theta}_k + n_i\tilde{\phi}_k - \omega_i t, \end{equation} and the relative wave--particle frequency as \begin{equation} \Omega(\bm{J}_k) \equiv \dot{\xi}_k = \ell\omega_\mathrm{B}(\bm{J}_k) + n_i\omega_\mathrm{p}(\bm{J}_k) - \omega_i. \end{equation} Next, we define a new momentum coordinate system $\bm{K} \equiv (\mu, S, W)$, where both $\mu$ and \begin{equation} S \equiv W - \frac{\omega_i}{n_i} P_\phi \end{equation} are constants of motion as particles move along the wave--particle characteristic curves of mode $i$. Assuming that the energetic particle distribution is located in a neighborhood around the wave--particle resonance $\bm{K}_\mathrm{res}$ (such that $\Omega(\bm{K}_\mathrm{res}) = 0$) where $\Omega' = \partial\Omega/\partial W$ and the interaction coefficient $V_{i,\ell}$ is constant in $\bm{K}$, the coordinate substitution \begin{equation}\label{eq:f2b}\left\{\begin{array}{*3{>{\displaystyle\vspace{1mm}}l}} \tau \equiv t/\tilde{t}, & \tilde{A} \equiv \tilde{t}^2\Omega'(\bm{K}_\mathrm{res}) V_{i,\ell}(\bm{K}_\mathrm{res}) A_i, \\ u_k \equiv \tilde{t}\Omega(\bm{K}_k), & \tilde{t} \equiv \left(\Omega'|V_{i,\ell}|^2\right)_{\bm{K}_\mathrm{res}}^{-1/3}, \end{array}\right.\end{equation} transforms the FOXTAIL model exactly to the 1D bump-on-tail model of eq.~\eqref{eq:bot}. Note that the substitutions of eq.~\eqref{eq:f2b} can be made for any FOXTAIL scenario, as long as one chooses a relevant eigenmode--Fourier coefficient pair and a resonant point $\bm{K}_\mathrm{res}$. According to the theory of the 1D bump-on-tail model, the linear growth rate of the mode ($|\tilde{A}(\tau)| \approx \tilde{A}_0\mathrm{e}^{\gamma_\mathrm{L}\tau}$) is~\cite{tho} \begin{equation}\label{eq:gam} \gamma_\mathrm{L} = \frac{\pi}{2}\left.\dr{\tilde{F}_0}{u}\right|_{u = 0}, \end{equation} where $\tilde{F}_0$ is the initial distribution function in $u$. The growth rate in time $t$ can be approximated by \begin{equation}\label{eq:gaf} \gamma_\mathrm{L} = \frac{\pi}{2}\frac{|V_{i,\ell}(\bm{K}_\mathrm{res})|^2}{\Omega'(\bm{K}_\mathrm{res})}\left.\dr{F_0}{W}\right|_{W_\mathrm{res}}, \end{equation} where $F_0$ is the $W$ distribution of energetic particles along the characteristic curves of wave--particle interaction ($F_0(W) = \int \mathrm{d}\mu\,\mathrm{d} S\,f_0(\bm{K})$). In the 1D bump-on-tail model, particles deeply trapped by the wave field oscillate in $\xi,u$-space around the resonance with the frequency\footnote{The frequency of eq.~\eqref{eq:omb} is in real time, $t$. For the frequency in the normalized time, $\tau$, the expression should be multiplied by $\tilde{t}$.} \begin{equation}\label{eq:omb} \omega_\mathrm{b} \approx \frac{\sqrt{|\tilde{A}|}}{\tilde{t}} = \sqrt{\left|\Omega'(\bm{K}_\mathrm{res})V_{i,\ell}(\bm{K}_\mathrm{res}) A_i\right|}. \end{equation} By comparing the numerical growth rates and other system properties of the 1D bump-on-tail model and FOXTAIL, one can make estimations of the effects of having non-constant $\Omega'$ and interaction coefficients, and of having multiple modes and/or Fourier coefficients. \subsubsection{Simulation parameters}\label{sec:sim} In this study, we consider an ITER equilibrium configuration with an energetic particle distribution consisting of $^3$He$^{2+}$ ions. For simplicity, we neglect collisions, explicit wave damping and particle sources and sinks. In such scenarios, the amplitudes of the modes are expected to grow exponentially until saturation at an amplitude proportional to the growth rate squared \cite{lev}, assuming that the modes interact more or less independently with the energetic particle distribution. Since the magnetic moment is not a dynamical variable in this system, the dimensionality of the problem can be reduced by considering an energetic particle distribution on a $\mu = \mathrm{const.}$ surface. An ad hoc energetic particle distribution functions is constructed, with $\mu = \SI{0.5}{MeV/T}$ (\SI{1}{eV} = \SI{1.6022e-19}{J}). The distribution in $W$ and $S$ are chosen to be Gaussians localized near the two resonances defined by $(n_i, \ell) = (5, 1)$ and $(n_i, \ell) = (6, 1)$. \begin{figure}[t!]\centering \includegraphics[width=88mm]{Fig3.pdf} \caption{The orbit categories, the bounce and the precession frequencies of trapped $^3$He$^{2+}$ ions on a $\Lambda,P_\phi$-grid with $\mu = \SI{0.5}{MeV/T}$. Category 0 orbits cross the last flux surface. For other categories, the convention of Ref.~\cite{hed} is used.}\label{fig:fbp} \end{figure} \begin{figure}[t!]\centering \includegraphics[width=88mm]{Fig4.pdf} \caption{The interaction coefficient $|V_{i,\ell}|$ plotted in $\bm{J}$-space, with $\mu = \SI{0.5}{MeV/T}$. The bold purple curves are the resonant curves $\ell\omega_\mathrm{B} + n_i\omega_\mathrm{p} = \omega_i$ for $\ell = 1$ and $i = 1$ (Fig.~\ref{fig:vil}.a), $i = 2$ (Fig.~\ref{fig:vil}.b). The thin black curves are the same resonant curves, but for $\ell = \{-2, -1, 0, 2\}$. The dashed curves are the wave--particle characteristic curves for the corresponding mode.}\label{fig:vil} \end{figure} \begin{figure}[t!]\centering \includegraphics[width=88mm]{Fig5.pdf} \caption{The energetic particle distributions of this study is placed close to the $\ell = 1$ resonant curves. Figure~\ref{fig:dis}.a shows the 3$\sigma$ curves for the Gaussian distributions of the energetic particles (97\,\% of the particles are contained within the 3$\sigma$ curves). The solid black curve is the ``narrow'' initial distribution, and the dashed black curve is the ``wide'' initial distribution. The solid purple curve is the $(i, \ell) = (1, 1)$ resonance, and the dashed purple curve is the $(i, \ell) = (2, 1)$ resonance. The reference point (blue dot) used for transforming to the 1D bump-on-tail coordinate system is placed on the $(1, 1)$ resonance. Both initial distributions have the same corresponding 1D bump-on-tail distribution, shown in Fig.~\ref{fig:dis}.b, where $N_\mathrm{EP}$ is the total number of energetic particles. The Gaussian bump-on-tail distribution is centered at $W - W_\mathrm{res} = \SI{3}{keV}$, and it has a width of $\sigma = \SI{3}{keV}$.}\label{fig:dis} \end{figure} The chosen $\bm{J}$-grid to calculate the guiding center orbits is in the region of trapped particles (category V- and VII-orbits in Fig.~\ref{fig:fbp}.a) on the $\mu = \SI{0.5}{MeV/T}$ surface. The corresponding bounce and precession frequencies, presented in Figs.~\ref{fig:fbp}.b and \ref{fig:fbp}.c, respectively, are found from the time dependence of the guiding center orbits, which are solved using eqs.~\eqref{eq:tps} and \eqref{eq:tph}. Two TAEs are chosen for this study: the first mode with a toroidal mode number $n = 5$, and the second one with $n = 6$. The energetic particle distribution is placed close to the resonance $(i, \ell) = (1, 1)$, i.e., the surface defined by $\omega_\mathrm{B} + 5\omega_\mathrm{p} = \omega_1$, where $\omega_1$ is the frequency of the first TAE. Figure~\ref{fig:vil} shows the calculated interaction coefficients $V_{i,\ell}(\bm{J})$ for $(i, \ell) = (1, 1)$, Fig.~\ref{fig:vil}.a, and $(i, \ell) = (2, 1)$, Fig.~\ref{fig:vil}.b. The frequencies of the two modes are \SI{38.2}{kHz} and \SI{40.2}{kHz}, respectively. Since the precession frequency is small compared to the bounce frequency at the $(1, 1)$ resonance, the $(1, 1)$ and the $(2, 1)$ resonances are almost the same in $\Lambda,P_\phi$-space. A reference point is chosen on the $(1, 1)$ resonance, as shown in Fig.~\ref{fig:dis}.a. Around this point, the bounce and precession frequencies are approximated to first order in $W,P_\phi$-space, and $V_{1,1}$ is approximated to zeroth order, when transforming from the FOXTAIL to the 1D bump-on-tail coordinate system. Two different initial distribution functions are used in these studies: one localized around the reference point and one that is more spread along the resonance. The two distributions transform to the same distribution in the 1D bump-on-tail model, shown in Fig.~\ref{fig:dis}.b. The initial distributions are chosen such that there is a positive derivative of the energy distribution at the resonance, giving a positive linear growth rate according to eq.~\eqref{eq:gam}. An energetic particle distribution consisting of $2.5\cdot 10^{16}$ $^3$He$^{2+}$ ions are distributed on $2.5\cdot 10^5$ markers. The markers are spread out in phase space using quasi-random low-discrepancy sequences (a Sobol' sequence~\cite{sob} combined with the Matou\v{s}ek scrambling method~\cite{mat}). They are placed as a Gaussian in $W$-space \emph{centered} around the resonance, and then the marker weights are set such that they represent the \emph{shifted} Gaussian as in Fig.~\ref{fig:dis}.b. This is done in order to improve the statistics of markers around the resonance as compared to a scenario with equal weights. \begin{table*}\centering \begin{tabular}{cccccccc} \toprule Case & Eigenmodes & Fourier coefficients & $\sigma_S$ [keV] & $\sigma_W$ [keV] & $V_{2,1}$ factor & $\Delta\omega/2\pi$ [kHz] & $N_\mathrm{part}$ [$10^{16}$] \\ \midrule \#1 & $i = \{1,2\}$ & $\ell = \{-2, \ldots, 2\}$ & 5.7 & 3.0 & 1 & 2.0 & 2.5 \\ \#2 & $i = 1$ & $\ell = 1$ & 5.7 & 3.0 & -- & -- & 2.5 \\ \#3 & $i = \{1,2\}$ & $\ell = \{-2, \ldots, 2\}$ & 22.7 & 3.0 & 1 & 2.0 & 2.5 \\ \#4 & $i = 1$ & $\ell = 1$ & 22.7 & 3.0 & -- & -- & 2.5 \\ \#5 & $i = \{1, 2\}$ & $\ell = 1$ & 5.7 & 3.0 & 10.1 & 2.0 & 2.5 \\ \#6 & $i = 2$ & $\ell = 1$ & 5.7 & 3.0 & 10.1 & -- & 2.5 \\ \#7 & $i = 1$ & $\ell = 1$ & 5.7 & 4.0 & -- & -- & 3.0 \\ \#8 & $i = \{1, 2\}$ & $\ell = 1$ & 5.7 & 4.0 & 13.2 & 6.1 & 3.0 \\ \#9 & $i = 2$ & $\ell = 1$ & 5.7 & 4.0 & 13.2 & -- & 3.0 \\\bottomrule \end{tabular}\caption{Summary of the initial parameters used in the FOXTAIL simulations presented in Figs.~\ref{fig:bot} -- \ref{fig:na2}, where $\sigma_S$ and $\sigma_W$ are the energy widths of the Gaussian energetic particle distribution along the resonance curve and along the characteristic curves for wave--particle interaction, respectively, $\Delta\omega$ is the frequency separation between the two modes and $N_\mathrm{part}$ is the amount of energetic particles that the markers represent.}\label{tab:par} \end{table*} \subsubsection{Numerical comparison with the 1D bump-on-tail model}\label{sec:bot} Here, different FOXTAIL scenarios are compared with the corresponding 1D bump-on-tail scenario by varying initial parameters such as the width of the initial energetic particle distribution along the resonances and the number of eigenmodes and Fourier coefficients to include in the simulations. Besides the 1D scenario, four FOXTAIL scenarios are presented. In the first two scenarios (referred to as cases \#1 and \#2), a narrow initial distribution is used (see Fig.~\ref{fig:dis}.a), whereas the two latter scenarios (cases \#3 and \#4) use the wide initial distribution. Furthermore, cases \#1 and \#3 include both of the TAEs presented in section~\ref{sec:sim} and a range of Fourier coefficients $-2 \leq \ell \leq 2$ for both eigenmodes. Cases \#2 and \#4 only include the first eigenmode ($n = 5$) and a single Fourier coefficient $\ell = 1$. For a complete list of initial parameter values used in the FOXTAIL simulations presented in this paper, see Table~\ref{tab:par}. Figure~\ref{fig:bot} shows the amplitude evolutions of the first eigenmode for the bump-on-tail simulation and FOXTAIL simulations \#1 -- 4. The amplitude of the second mode never grows larger than $\approx$ 0.8\,\% of the amplitude of the first mode after saturation in case \#1, and up to 5\,\% in case \#3. It can immediately be seen that the two FOXTAIL scenarios with a narrow initial distribution (cases \#1 and \#2) agrees well with the corresponding 1D bump-on-tail scenario, both in growth rate and in saturation amplitude. On the other hand, the wider distribution (cases \#3 and \#4) gives a different growth rate and saturation level of the amplitude. This is presumably due to the fact that the wide distribution spans over regions where the interaction coefficient $V_{1,1}$ is considerably lower than in the reference point, giving a lower growth rate on average. Including multiple eigenmodes and Fourier coefficients in the simulations seem to have negligible effects on the system in the considered scenarios, both for the narrow and the wide initial distributions. The reason for why the second mode does not influence the system considerably is because it has an expected growth rate of approximately 100 times lower than the first mode, which is due to the comparably lower values of the interaction coefficient $V_{2,1}$ in the region of the initial distribution function (recall that the growth rate scales as $|V_{i,\ell}|^2$, see eq.~\eqref{eq:gaf}). When comparing growth rates of the different scenarios, it should be noted that $\gamma_\mathrm{L}$ of the 1D bump-on-tail scenario is approximately 81\,\% of the analytical growth rate of eq.~\eqref{eq:gam}, whereas the growth rates of cases \#1 and \#2 are 77\,\% of the analytical growth rate, and 68\,\% for cases \#3 and \#4. The reason why the growth rate of the 1D bump-on-tail scenario is considerably lower than the analytical one is primarily because of the finite extension of the energetic particle distribution along the characteristic curves (this issue was analyzed in more detail in Ref.~\cite{tho}). \begin{figure}[t!]\centering \includegraphics[width=88mm]{Fig6.pdf} \caption{Amplitude evolution of the first mode $(n = 5)$. $\omega_\mathrm{b} \propto \sqrt{|A_i|}$ is the bounce frequency of particles deeply trapped by the wave field, and $\gamma_\mathrm{L}$ is the analytical linear growth rate of the mode, calculated from the value of $|V_{i,\ell}|$ and $\partial\Omega/\partial W$ in the reference point. The solid black curve uses the 1D bump-on-tail model, which is analogous to FOXTAIL using a single mode--Fourier coefficient pair (in this case $(i, \ell) = (1, 1)$) and constant $V_{i,\ell}$ and $\Omega'$ in $\Lambda,P_\phi$-space. Cases \#1 and \#2 use a narrow initial energetic distribution function along the $(i, \ell) = (1, 1)$ resonance curve in $\Lambda,P_\phi$-space, whereas cases \#3 and \#4 use a wide initial distribution. Cases \#1 and \#3 include both TAEs and Fourier coefficients $-2 \leq \ell \leq 2$, whereas cases \#2 and \#4 only include the $(i, \ell) = (1, 1)$ mode--Fourier coefficient pair. See Table~\ref{tab:par} for a list of parameter values used in all FOXTAIL simulations.}\label{fig:bot} \end{figure} \subsection{Numerical multi-mode studies} The presence of multiple eigenmodes proved to have negligible effect on the system in the scenarios presented in section~\ref{sec:bot} due to the low interaction coefficient of the second mode in the considered part of $\Lambda,P_\phi$-space, compared to the interaction coefficient of the first mode. Scenarios with significant multimode dynamics can be constructed by adding an ad hoc scaling factor to the interaction coefficient of the second mode. Multiplying $V_{2,1}$ by a factor of 10.1 gives approximately the same linear growth rate of the two modes. This has been done for cases \#5 and \#6, presented in Fig.~\ref{fig:nar}, along with the previous case \#2. The three scenarios are the same, except that in cases \#2 and \#6 the eigenmodes are simulated individually, whereas in case \#5 both modes are included to the system. Such a comparison allows one to isolate the nonlinear effects of mode interaction via the energetic particle distribution. \begin{figure}[t!]\centering \includegraphics[width=88mm]{Fig7.pdf} \caption{Figure~\ref{fig:nar}.a shows the mode amplitude evolution for a set of FOXTAIL simulations using the narrow initial distribution function. Case \#4 includes both TAEs, and the Fourier coefficient $\ell = 1$ for each mode. Only the evolution of the first mode is presented. Case \#5 is the same as \#4, but the interaction coefficient $V_{2,1}$ is scaled up by a factor of 10.1, such that the linear growth rates of the two modes approximately match. Case \#6 is the same as \#5, but the first mode is deactivated in the simulation. See Table~\ref{tab:par} for a list of parameter values used in all FOXTAIL simulations. Figure~\ref{fig:nar}.b shows the corresponding 1D bump-on-tail distribution of the above simulations (the final distribution is at $t = \SI{10}{ms}$). Figure \ref{fig:nar}.c tests the Chirikov criterion for case \#5 by dividing the average resonance width in $W$-space and divide by the average energy separation between the resonances along the two characteristic curves.}\label{fig:nar} \end{figure} \begin{figure}[t!]\centering \includegraphics[width=88mm]{Fig8.pdf} \caption{The same as Fig.~\ref{fig:nar}, but for slightly different scenarios. The frequency separation between the two modes is increased by a factor of 3, the width of the Gaussian energetic particle distribution along the characteristic curve of the first mode is increased from $\sigma = \SI{3}{keV}$ to $\sigma = \SI{4}{kev}$ and the number of particles is increased from $2.5\times 10^{16}$ to $3.0\times 10^{16}$. The scaling factor of the interaction coefficient $V_{2,1}$ is increased from 10.1 to 13.2 in order to match the linear growth rates of the two modes. The amplitude evolutions of the second mode ($A_2$) in Fig.~\ref{fig:na2}.a are smoothened in order to remove high frequency amplitude oscillations coming from the interactions with off-resonant particles and from statistical fluctuations. $\Delta F(W)$ of Fig.~\ref{fig:na2}.b is the difference between the final and the initial corresponding bump-on-tail distributions.}\label{fig:na2} \end{figure} Comparing the multimode scenario with the scenarios where the modes are included individually, it can be seen that the indirect interaction between the modes via the energetic particles has significant macroscopic effects on the system. This can partly be understood as a consequence of stochastization of particle trajectories in phase space due to resonance overlap of the two eigenmodes. Stochastization of trajectories causes a locally enhanced transport of energetic particles around the resonances, allowing the eigenmodes to exhaust more energy from the energetic particle distribution (see e.g.\ stochastization from resonance overlap, \cite{bb4,bb5}, and from phase decorrelation, \cite{tho}). This results in a wider portion of the energetic particle distribution being flattened around the resonances, compared to the individual eigenmode cases, as seen in Fig.~\ref{fig:nar}.b. The resonance width of an eigenmode can be estimated as the separatrix width of the unperturbed mode (i.e.\ in the absence of other modes) along the characteristic curve in $W$-space, using the 1D bump-on-tail approximation. When the resonance widths of the two modes are comparable, the resonance-overlap parameter~\cite{chi} (English translation: Ref.~\cite{ch2}), estimated as the average resonance width of the two modes divided by their distance in phase space, is an approximate measure of the level of stochastization of particle trajectories. The Chirikov criterion for stochastization is satisfied when the resonance-overlap parameter is larger than unity. The full separatrix width in $W$-space, $W_\mathrm{sep}$, is $4\omega_\mathrm{b}/\Omega'(\bm{K}_\mathrm{res})$, with $\omega_\mathrm{b}$ given by eq.~\eqref{eq:omb}. As seen in Fig.~\ref{fig:nar}.c, the Chirikov criterion is well satisfied for case \#5 after $t \gtrsim \SI{1.3}{ms}$. Slightly different scenarios are tested in the simulations presented in Fig.~\ref{fig:na2}. Cases \#7, \#8 and \#9 are the same as cases \#2, \#5 and \#6, respectively, except that the frequency separation between the two eigenmodes is artificially increased by a factor of 3, and the width of the energetic distribution function along the characteristic curves of the first mode is increased to $\sigma = \SI{4}{keV}$ instead of \SI{3}{keV}. The scaling factor of the interaction coefficient is also adjusted such that the linear growth rates of the two modes approximately match. As can be seen in Fig.~\ref{fig:na2}.c, the Chirikov criterion is never satisfied for case \#8, although the resonance-overlap parameter is of the order of unity. Comparing the amplitude evolutions in Figs.~\ref{fig:nar}.a and \ref{fig:na2}.a, the two modes of case \#8 evolves more similarly to the corresponding individual mode scenarios than case \#5 does. This is especially seen in Fig.~\ref{fig:na2}.b, where the modes of case \#8 flattens two separate regions of the energetic distribution function, matching the flattening regions of the individual mode scenarios. \section{Summary}\label{sec:con} This paper presents the theoretical framework of the FOXTAIL code, which is used to describe the nonlinear interaction between Alfvén eigenmodes and energetic particles in toroidal geometries. FOXTAIL is a hybrid magnetohydrodynamic--kinetic model based on a model developed by Berk \emph{et al.}~\cite{bb1}, where each simulation is formulated as an initial value problem. Eigenmodes are treated as perturbations of the equilibrium system, with temporally constant eigenfunctions and dynamic complex amplitudes that vary on time scales longer than the inverse mode frequency. The energetic particle distribution is modeled by a finite set of markers in an action--angle phase space of the unperturbed system. The use of action--angle coordinates rather than conventional toroidal coordinates simplifies the equations of motion of the individual markers, and it allows for efficient resolution of time scales relevant for resonant eigenmode--particle interaction in numerical simulations. The particle response with respect to the wave field is quantized by a Fourier series expansion of the kinetic energy change $q \bm{v}\cdot\delta\bm{E}$ along the transit period of the unperturbed guiding center orbit, where $q$ is the particle charge, $\bm{v}$ is the guiding center velocity and $\delta\bm{E}$ is the electric wave field at the guiding center position. A Lagrangian formulation of the wave--particle system, consistent with the derived particle response with respect to the wave, is used to derive equations for the eigenmode amplitudes and phases. The resulting system of equations describing direct wave--particle interaction has a phase space with four particle dimensions and two eigenmode dimensions (amplitude and phase). When including mechanisms that perturb the magnetic moment of energetic particles, the particle phase space extends to five dimensions. When splitting the interaction in the Fourier terms along the transit period, each term contributes to resonant nonlinear interaction mainly in a narrow region around surfaces in the adiabatic invariant space, referred to as resonant surfaces. These surfaces are given by $\ell\omega_\mathrm{B} + n_\phi\omega_\mathrm{p} = \omega_\mathrm{mode}$, where $\ell$ is the Fourier index of interaction, $\omega_\mathrm{B}$ is the bounce frequency, $n_\phi$ is the toroidal mode number of the eigenmode, $\omega_\mathrm{p}$ is the precession frequency and $\omega_\mathrm{mode}$ is the eigenmode frequency. The width of the relevant region around the resonant surfaces depends on the amplitude of the eigenmode, the strength of wave--particle interaction at the resonant surfaces (quantified by the Fourier coefficients of interaction) and the variation of bounce and precession frequencies of particles along the characteristic curves of wave--particle interaction (the curves in adiabatic invariant space along which a given eigenmode accelerates particles). The presented multi-dimensional model can be approximated with a conventional 1D bump-on-tail model. For the 1D approximation to be valid, three approximate criteria must be met: \begin{itemize} \item No more than one eigenmode--Fourier coefficient pair interacts significantly with the energetic particle distribution. \item The complex Fourier coefficient of interaction is approximately constant in adiabatic invariant space throughout the resonant part of the energetic particle distribution. \item The bounce and precession frequencies of the energetic particles vary approximately linearly in kinetic energy--toroidal canonical momentum space across the region of the resonance where the energetic particle distribution is located. \end{itemize} All these conditions can be quantitatively evaluated with FOXTAIL. Effects of the fulfillment of the Chirikov criterion in scenarios with two active eigenmodes have been studied numerically using FOXTAIL. It has been verified that eigenmodes can be treated independently in scenarios where the criterion is not satisfied. On the other hand, when the resonance-overlap parameter becomes larger than unity, indirect mode--mode interaction via the energetic particle distribution becomes significant, and a larger portion of the inverted energetic particle distribution becomes flattened in energy space due to stochastization of particle trajectories in phase space. \section*{Acknowledgments} This work was supported by the Swedish research council (VR) contract 621-2011-5387. \bibliographystyle{elsarticle-num}
190aa66251db8882ee96e7743007257d39d8fb51
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction}\label{sec:intro} A major challenge towards self-organizing networks (SON) is the joint optimization of multiple SON use cases by coordinately handling multiple configuration parameters. Widely studied SON use cases include coverage and capacity optimization (CCO), mobility load balancing (MLB) and mobility robustness optimization (MRO)\cite{3GPP36902}.\cosl{We need a reference here} However, most of these works study an isolated single use case and ignore the conflicts or interactions between the use cases \cite{giovanidis2012dist,razavi2010self}. In contrast, this paper considers a joint optimization of two strongly coupled use cases: CCO and MLB. The objective is to achieve a good trade-off between coverage and capacity performance, while ensuring a load-balanced network. The SON functionalities are usually implemented at the network management layer and are designed to deal with \lq\lq long-term\rq\rq \ network performance. Short-term optimization of individual users is left to lower layers of the protocol stack. To capture long-term global changes in a network, we consider a cluster-based network scenario, where users served by the same base station (BS) with similar SINR distribution are adaptively grouped into clusters. Our objective is to jointly optimizing the following variables: \begin{itemize} \item Cluster-based BS assignment and power allocation. \item BS-based antenna tilt optimization and power allocation. \end{itemize} The joint optimization of assignment, antenna tilts, and powers is an inherently challenging problem. The interference and the resulting performance measures depend on these variables in a complex and intertwined manner. Such a problem, to the best of the authors' knowledge, has been studied in only a few works. For example, in \cite{klessig2012improving} a problem of jointly optimizing antenna tilt and cell selection to improve the spectral and energy efficiency is stated, however, the solution derived by a structured searching algorithm may not be optimal. In this paper, we propose a robust algorithmic framework built on a utility model, which enables fast and near-optimal uplink solutions and sub-optimal downlink solutions\cosl{Do we know that this is near-optimal?} by exploiting three properties: 1) the monotonic property and fixed point of the monotone and strictly subhomogenoues (MSS) functions \footnote{Many literatures use the term {\it interference function} for the functions satisfy three condotions, positivity, monotonicity and scalability \cite{yates95}. Positivity is shown to be a consequence of the other two properties \cite{leung2004convergence}, and we use the term {\it strctly subhomogeneous} in place of scalable from a constraction mapping point of view in keeping with some related literature \cite{nuzman2007contraction}.}, 2) decoupled property of the antenna tilt and BS assignment optimization in the uplink network, and 3) uplink-downlink duality. The first property admits global optimal solution with fixed-point iteration for two specific problems: utility-constrained power minimization and power-constrained max-min utility balancing \cite{vucic2011fixed,stanczak2009fundamentals,schubert2012interference,yates95}. The second and third properties enable decomposition of the high-dimensional optimization problem, such as the joint beamforming and power control proposed in \cite{BocheDuality06,schubert2005iterative,huang2013joint,he2012multi}. Our distinct contributions in this work can be summarized as follows:\\ 1) We propose a max-min utility balancing algorithm for capacity-coverage trade-off optimization over a joint space of antenna tilts, BS assignments and powers. The utility defined as a convex combination of the average SINR and the worst-case SINR implies the balanced performance of capacity and coverage. Load balancing is improved as well due to a uniform distribution of the interference among the BSs.\\ 2) The proposed utility is formulated based on the MSS functions, which allows us to find the optimal solution by applying fixed-point iterations.\\ 3) Note that antenna tilts are BS-specific variables, while assignments are cluster-specific, we develop two optimization problems with the same objective functions, formulated either as a problem of per-cluster variables or as a problem of per-base variables. We propose a two-step optimization algorithm in the uplink to iteratively optimize the per BS variables (antenna tilts and BS power budgets) and the cluster-based variables (assignments and cluster power). Since both problems aim at optimizing the same objective function, the algorithm is shown to be convergent.\\ 4) The decoupled property of antenna tilt and assignment in the uplink decomposes the high-dimensional optimization problem and enables more efficient optimization algorithm. We then analyze the uplink-downlink duality by using the Perron-Frobenius theory\cite{meyer2000matrix}, and propose an efficient sub-optimal solution in the downlink by utilizing optimized variables in the dual uplink. \section{System Model}\label{sec:Model} We consider a multicell wireless network composed of a set of BSs $\set{N}:=\{1,\ldots, N\}$ and a set of users $\set{K}:=\{1,\ldots, K\}$. Using fuzzy C-means clustering algorithm \cite{bezdek1984fcm}, we group users with similar SINR distributions\footnote{We assume the Kullback-Leibler divergence as the distance metric.} and served by the same BS into clusters. The clustering algorithm is beyond the scope of this paper. Let the set of user clusters be denoted by $\set{C}:=\{1,\ldots,C\}$, and let $\bm{A}$ denote a $C\times K$ binary user/cluster assignment matrix whose columns sum to one. The BS/cluster assignment is defined by a $N\times C$ binary matrix $\bm{B}$ whose columns also sum to one. Throughout the paper, we assume a frequency flat channel. The average/long-term downlink path attenuation between $N$ BSs and $K$ users are collected in a channel gain matrix $\bm{H}\in {\field{R}}^{N\times K}$. We introduce the cross-link gain matrix $\bm{V}\in{\field{R}}^{K\times K}$, where the entry $v_{lk}(\theta_j)$ is the cross-link gain between user $l$ served by BS $j$, and user $k$ served by BS $i$, i.e., between the transmitter of the link $(j, l)$ and the receiver of the link $(i, k)$. Note that $v_{lk}(\theta_j)$ depends on the antenna downtilt $\theta_j$. Let the BS/user assignment matrix be denoted by $\bm{J}$ so that we have $\bm{J}:=\bm{B}\bm{A}\in\{0,1\}^{N\times K}$, and $\bm{V}:=\bm{J}^T\bm{H}$. We denote by $\bm{r}:=[r_1, \ldots, r_N]^T$, $\bm{q}:=[q_1, \ldots, q_C]^T$ and $\bm{p}:=[p_1, \ldots, p_K]^T$the BS transmission power budget, the cluster power allocation and the user power allocation, respectively. % \subsection{Inter-cluster and intra-cluster power sharing factors} \label{subsec:powFactor} We introduce the inter-cluster and intra-cluster power sharing factors to enable the transformation between two power vectors with different dimensions. Let $\bm{b}:=[b_1, \ldots, b_C]^T$ denote the serving BSs of clusters $\{1, \ldots, C\}$. We define the vector of the inter-cluster power sharing factors to be $\bm{\beta}:=[\beta_1, \ldots, \beta_C]^T$, where $\beta_c:=q_c/r_{b_c}$. With the BS/cluster assignment matrix $\bm{B}$, we have $\bm{q}:=\ma{B}_{\ve{\beta}}^T \bm{r}$, where $\ma{B}_{\ve{\beta}}:=\bm{B}\mathop{\mathrm{diag}}\{\bm{\beta}\}$. Since users belonging to the same cluster have similar SINR distribution, we allocate the cluster power uniformly to the users in the cluster. The intra-cluster sharing factors are represented by $\bm{\alpha}:=[\alpha_1, \ldots, \alpha_K]^T$ with $\alpha_k=1/|\set{K}_{c_k}|$ for $k\in\set{K}$, where $\set{K}_{c_k}$ denotes the set of users belonging to cluster $c_k$, while $c_k$ denotes the cluster with user $k$. We have $\bm{p}:=\ma{A}_{\ve{\alpha}}^T\bm{q}$, where $\ma{A}_{\ve{\alpha}}:=\bm{A}\mathop{\mathrm{diag}}\{\bm{\alpha}\}$. The transformation between BS power $\bm{r}$ and user power $\bm{p}$ is then $\bm{p}:=\bm{T}\bm{r}$ where the transformation matrix $\bm{T}:=\ma{A}_{\ve{\alpha}}^T\ma{B}_{\ve{\beta}}^T$. % \subsection{Signal-to-interference-plus-noise ratio}\label{subsec:SINR} Given the cross-link gain matrix $\bm{V}$, the downlink SINR of the $k$th user depends on all powers and is given by \begin{equation} \operator{SINR}_k^{(\text{d})}:=\frac{p_k \cdot v_{kk}(\theta_{n_k})}{\sum_{l\in\set{K}\setminus k} p_l \cdot v_{lk}(\theta_{n_l})+\sigma_k^2}, k\in\set{K} \label{eqn:DL_SINR} \end{equation} where $n_k$ denotes the serving BS of user $k$, $\sigma_k^2$ denotes the noise power received in user $k$. Likewise, the uplink SINR is \begin{equation} \operator{SINR}_k^{(\text{u})}:=\frac{p_k \cdot v_{kk}(\theta_{n_k})}{\sum_{l\in\set{K}\setminus k} p_l \cdot v_{kl}(\theta_{n_k})+\sigma_k^2}, k\in\set{K} \label{eqn:UL_SINR} \end{equation} % Assuming that there is no self-interference, the cross-talk terms can be collected in a matrix \begin{equation} [\tilde{\ma{V}}]_{lk}:= \begin{cases} v_{lk}(\theta_{n_l}), & l\neq k\\ 0, & l=k \end{cases}. \label{eqn:PsiMat} \end{equation} Thus the downlink interference received by user $k$ can be written as $I_k^{(\text{d})}:=[\tilde{\bm{V}}^T\bm{p}]_k$, while the uplink interference is given by $I_k^{(\text{u})}:=[\tilde{\bm{V}}\bm{p}]_k$. A crucial property is that the uplink SINR of user $k$ depends on the BS assignment $n_k$ and the single antenna tilt $\theta_{n_k}$ alone, while the downlink SINR depends on the BS assignment vector $\bm{n}:=[n_1,\ldots, n_K]^T$, and the antenna tilt vector $\bm{\theta}:=[\theta_1, \ldots, \theta_N]^T$. The decoupled property of uplink transmission has been widely exploited in the context of uplink and downlink multi-user beamforming \cite{BocheDuality06}\cosl{Reference} and provides a basis for the optimization algorithm in this paper. % The notation used in this paper is summarized in Table \ref{tab:CovCap_notation}. \begin{table}[t] \centering \caption{NOTATION SUMMARY} \begin{tabular}{|c|c|} \hline ${\emenge{N}}$ & set of BSs \\ ${\emenge{K}}$ & set of users \\ ${\emenge{C}}$ & set of user clusters\\ $\bm{A}$ & cluster/user assignment matrix\\ $\bm{B}$ & BS/cluster assignment matrix\\ $\bm{J}$ & BS/user assignment matrix\\ $c_k$ & cluster that user $k$ is subordinated to\\ ${\emenge{K}}_{c}$ & set of users subordinated to cluster $c$\\ $\bm{H}$ & channel gain matrix\\ $\bm{V}$ & interference coupling matrix\\ $\tilde{\bm{V}}$ & interference coupling matrix without intra-cell interference\\ $\tilde{\bm{V}}_{\bm{b}}$ & interference coupling matrix depending on BS assignments $\bm{b}$\\ $\tilde{\bm{V}}_{\bm{\theta}}$ & interference coupling matrix depending on antenna tilts $\bm{\theta}$\\ $\bm{r}$ & BS power budget vector\\ $\bm{q}$ & cluster power vector\\ $\bm{p}$ & user power vector\\ $\bm{\alpha}$ & intra-cluster power sharing factors\\ $\bm{\beta}$ & inter-cluster power sharing factors\\ $\bm{A}_{\bm{\alpha}}$ & transformation from $\bm{q}$ to $\bm{p}$, $\bm{p}:=\bm{A}_{\bm{\alpha}}^T\bm{q}$\\ $\bm{B}_{\bm{\beta}}$ & transformation from $\bm{r}$ to $\bm{q}$, $\bm{q}:=\bm{B}_{\bm{\beta}}^T\bm{r}$\\ $\bm{T}$ & transformation from $\bm{r}$ to $\bm{p}$, $\bm{p}:=\bm{T}\bm{r}$\\ $\bm{\theta}$ & BS antenna tilt vector\\ $\bm{b}$ & serving BSs of clusters\\ $b_c$ & serving BS of cluster $c$\\ $\bm{n}$ & serving BSs of the users\\ $n_k$ & serving BS of user $k$\\ $\bm{\sigma}$ & noise power vector\\ $P^{\text{max}}$ & sum power constraint\\ \hline \end{tabular} \label{tab:CovCap_notation} \end{table} \section{Utility Definition and Problem Formulation}\label{sec:ProbForm} As mentioned, the objective is a joint optimization of coverage, capacity and load balancing. We capture coverage by the worst-case SINR, while the average SINR is used to represent capacity. A cluster-based utility $U_c(\bm{\theta},\bm{r},\bm{q},\bm{b})$ is introduced as the combined function of the worst-case SINR and average SINR, depending on BS power allocation $\bm{r}$, antenna downtilt $\bm{\theta}$ , cluster power allocation $\bm{q}$ and BS/cluster assignment $\bm{b}$.\footnote{The reader should note that user-specific variables $(\bm{p},\bm{n})$ can be derived directly from cluster-specific variables $\bm{q}$ and $\bm{b}$, provided that cluster/user assignment $\bm{A}$ and intra-cluster power sharing factor $\bm{\alpha}$ are given.} To achieve the load balancing by distributing the clusters to the BSs such that their utility targets can be achieved \footnote {The assignment of clusters also distributes the interference among the BSs.}, we formulate the following objective $$\max_{(\bm{r},\bm{\theta},\bm{q},\bm{b})}\min_{c\in\set{C}} \frac{U_c(\bm{r},\bm{\theta},\bm{q},\bm{b})}{\gamma_c}$$ where $\gamma_c$ is the predefined utility target for cluster $c$. The BS variables $(\bm{r},\bm{\theta})$ and cluster variables $(\bm{q}, \bm{b})$ are optimized by iteratively solving\\ 1) Cluster-based BS assignment and power allocation $\max_{(\bm{q},\bm{b})}\max_{c\in\set{C}} U_c(\bm{q},\bm{b})/\gamma_c$ given the fixed $(\hat{\bm{r}},\hat{\bm{\theta}})$ \\ 2) BS-based antenna tilt optimization and power allocation $\max_{(\bm{r},\bm{\theta})}\max_{c\in\set{C}} U_c(\bm{r},\bm{\theta})/\gamma_c$ given the fixed $(\hat{\bm{q}},\hat{\bm{b}})$. In the following we introduce the utility definition and problem formulation for the cluster-based and the BS-based problems respectively. We start with the problem statement and algorithmic approaches for the uplink. We then discuss the downlink in Section \ref{sec:Duality}. % \subsection{Cluster-Based BS Assignment and Power Allocation}\label{subsec:clusterOpt} Assume the per-BS variables $(\hat{\bm{r}}, \hat{\bm{\theta}})$ are fixed, let the interference coupling matrix depending on BS assignment $\bm{b}$ in \eqref{eqn:PsiMat} be denoted by $\V_{\ve{b}}$. We first define two utility functions indicating capacity and coverage per cluster respectively, then we introduce the joint utility as a combination of the capacity and coverage utility. After that we define the cluster-based max-min utility balancing problem based on the joint utility. % \subsubsection{Average SINR Utility (Capacity)}\label{subsubsec:LB_A} With the intra-cluster power sharing factor introduced in Section \ref{subsec:powFactor}, we have $\bm{p}:=\ma{A}_{\ve{\alpha}}^T \bm{q}$. Define the noise vector $\bm{\sigma}:=[\sigma_1^2, \ldots, \sigma_K^2]^T$, the average SINR of all users in cluster $c$ is written as \begin{align} \bar{U}_c^{(\text{u},1)}&(\bm{q}, \bm{b}) := \frac{1}{|\set{K}_c|} \sum_{k\in\set{K}_c}\operator{SINR}_k^{(\text{u})}\nonumber\\ &= \frac{1}{|\set{K}_c|} \sum_{k\in\set{K}_c}\frac{q_c \alpha_k v_{kk}}{\left[\V_{\ve{b}} \ma{A}_{\ve{\alpha}}^T \bm{q}+\bm{\sigma}\right]_k}\nonumber\\ &\geq \frac{1}{|\set{K}_c|}\frac{q_c \sum_{k\in\set{K}_c} \alpha_k v_{kk}}{\sum_{k\in\set{K}_c} \left[\V_{\ve{b}} \ma{A}_{\ve{\alpha}}^T \bm{q}+\bm{\sigma}\right]_k} =U_c^{(\text{u},1)}(\bm{q}, \bm{b}) \label{eqn:CL_cap_1} \end{align} The uplink capacity utility of cluster $c$ denoted by $U_c^{(\text{u},1)}$ is measured by the ratio between the total useful power and the total interference power received in the uplink in the cluster. Utility $U_c^{(\text{u},1)}$ is used instead of $\bar{U}_c^{(\text{u},1)}$ because of two reasons: First, it is a lower bound for the average SINR. Second, it has certain monotonicity properties (introduced in Section \ref{sec:OPAlgor}) which are useful for optimization. Introducing the cluster coupling term $\overline{\ma{G}}_{\ve{b}}^{(\text{u})}:=\bm{\Psi}\bm{A}\V_{\ve{b}}\ma{A}_{\ve{\alpha}}^T$, where $\bm{\Psi}:=\mathop{\mathrm{diag}}\{|\set{K}_1|/g_1, \ldots, |\set{K}_c|/g_C\}$ and $g_c:=\sum_{k\in \set{K}_c}\alpha_k v_{kk}$ for $c\in\set{C}$; and the noise term $\overline{\bm{z}}:=\bm{\Psi}\bm{A}\bm{\sigma}$, the capacity utility is simplified as \begin{align} U_c^{(\text{u},1)}(\bm{q}, \bm{b})&:=\frac{q_c}{\set{J}_c^{(\text{u},1)}(\bm{q}, \bm{b})}\label{eqn:CL_cap_2}\\ \mbox{where } \set{J}_c^{(\text{u},1)}(\bm{q}, \bm{b})&:=\left[\overline{\ma{G}}_{\ve{b}}^{(\text{u})}\bm{q}+\overline{\bm{z}}\right]_c. \label{eqn:CL_cap_inter} \end{align} % \subsubsection{Worst-Case SINR Utility (Coverage)} Roughly speaking, the coverage problem arises when a certain number of the SINRs are lower than the predefined SINR threshold. Thus, improving the coverage performance is equivalent to maximizing the worst-case SINR such that the worst-case SINR achieves the desired SINR target. We then define the uplink coverage utility for each cluster as \begin{align} U_c^{(\text{u},2)}(\bm{q},\bm{b})&:=\min_{k\in\set{K}_c}\operator{SINR}_k^{(\text{u})}=\min_{k\in\set{K}_c} \frac{q_c\alpha_k v_{kk}}{\left[\V_{\ve{b}} \ma{A}_{\ve{\alpha}}^T \bm{q}+\bm{\sigma}\right]_k}\nonumber\\ &= \frac{q_c}{\max_{k\in\set{K}_c}\left[ \bm{\Phi}\V_{\ve{b}} \ma{A}_{\ve{\alpha}}^T \bm{q}+\bm{\Phi}\bm{\sigma}\right]_k} \label{eqn:CL_cov_1} \end{align} where $\bm{\Phi}:=\mathop{\mathrm{diag}}\{1/\alpha_1 v_{11}, \ldots, 1/\alpha_K v_{KK}\}$. We define a $C \times K$ matrix $\bm{X}:=[\bm{x}_1|\ldots|\bm{x}_C]^T$, where $\bm{x}_c:=\bm{e}^j_K$ and $\bm{e}^j_i$ denotes an $i$-dimensional binary vector which has exact one entry (the j-th entry) equal to 1. Introducing the term $\underline{\ma{G}}_{\ve{b}}^{(\text{u})}:=\bm{\Phi}\V_{\ve{b}} \ma{A}_{\ve{\alpha}}^T$, and the noise term $\underline{\bm{z}}:=\bm{\Phi}\bm{\sigma}$, the coverage utility is given by \begin{align} U_c^{(\text{u},2)}(\bm{q},\bm{b})&:=\frac{q_c}{\set{J}_c^{(\text{u},2)}(\bm{q}, \bm{b})}\label{eqn:CL_cov_2}\\ \mbox{where } \set{J}_c^{(\text{u},2)}(\bm{q}, \bm{b}) & := \max_{\bm{x}_c:=\bm{e}_K^j, j\in\set{K}_c} \left[\bm{X}\underline{\ma{G}}_{\ve{b}}^{(\text{u})}\bm{q}+\bm{X}\underline{\bm{z}}\right]_c. \label{eqn:CL_cov_inter} \end{align} % \subsubsection{Joint Utility and Cluster-Based Max-Min Utility Balancing}\label{eqn:LB_maxmin} The joint utility $U_c^{(\text{u})}(\bm{q}, \bm{b})$ is defined as \begin{align} U_c^{(\text{u})}(\bm{q}, \bm{b})&:=\frac{q_c}{\set{J}_c^{\ul}(\bm{q}, \bm{b})}\label{eqn:LB_utility_1}\\ \mbox{where }\set{J}_c^{\ul}(\bm{q}, \bm{b})&:= \mu\set{J}_c^{(\text{u},1)}(\bm{q}, \bm{b})+(1-\mu)\set{J}_c^{(\text{u},2)}(\bm{q}, \bm{b})\label{eqn:LB_utility_2}. \end{align} In other words, the joint interference function $\set{I}_c^{(\text{u})}$ is a convex combination of $\set{I}_c^{(\text{u},1)}$ in \eqref{eqn:CL_cap_inter} and $\set{I}_c^{(\text{u},2)}$ in \eqref{eqn:CL_cov_inter}. The cluster-based power-constrained max-min utility balancing problem in the uplink is then provided by \begin{problem}[Cluster-Based Utility Balancing] \begin{equation} C^{(\text{u})}(P^{\text{max}})=\max_{\bm{q}\geq 0, \bm{b}\in \set{N}^C} \min_{c\in\set{C}} \frac{U_c^{(\text{u})}(\bm{q}, \bm{b})}{\gamma_c}, \mbox{s.t. } \|\bm{q}\|\leq P^{\text{max}} \label{eqn:LB_OP} \end{equation} Here, $\|\cdot\|$ is an arbitrary monotone norm, i.e., $\bm{q}\leq\bm{q}'$ implies $\|\bm{q}\|\leq\|\bm{q}'\|$, $P^{\text{max}}$ denotes the total power constraint. According to the joint utility in \eqref{eqn:LB_utility_1},\eqref{eqn:LB_utility_2}, the algorithm optimizes the performance of capacity when we set the tuning parameter $\mu=1$ (utility is equivalent to the capacity utility in \eqref{eqn:CL_cap_2}), while with $\mu=0$ it optimizes the performance of coverage (utility equals to the coverage utility in \eqref{eqn:CL_cov_2}). By tuning $\mu$ properly, we can achieve a good trade-off between the performance of coverage and capacity. \label{prob:LB} \end{problem} % \subsection{BS-Based Antenna Tilt Optimization and Power Allocation}\label{subsec:AO} Given the fixed $(\hat{\bm{q}},\hat{\bm{b}})$, we compute the intra-cluster power allocation factor $\bm{\beta}$, given by $\beta_c:=\hat{q}_c/\sum_{c\in\set{C}_{b_c}}\hat{q}_c$ for $c\in\set{C}$. We denote the cross-link coupling matrix depending on $\bm{\theta}$ by $\V_{\ve{\theta}}$. In the following we formulate the BS-based max-min utility balancing problem such that it has the same physical meaning as the problem stated in \eqref{eqn:LB_OP}. We then introduce the BS-based joint utility interpreted by $(\bm{r}, \bm{\theta})$. \subsubsection{BS-Based Max-Min Utility Balancing}\label{subsubsec:AO_maxmin} To be consistent with our objective function $C^{(\text{u})}(P^{\text{max}})$ in \eqref{eqn:LB_OP}, we transform the cluster-based optimization problem to the BS-based optimization problem: % \begin{problem}[BS-Based Utility Balancing] \begin{align} C^{(u)}&(P^{\text{max}})=\max\limits_{\bm{r}\geq 0, \bm{\theta}\in\Theta^N} \min\limits_{c\in\set{C}} \frac{U_c^{(\text{u})}(\bm{r},\bm{\theta})}{\gamma_c}\nonumber\\ &=\max\limits_{\bm{r}\geq 0, \bm{\theta}\in\Theta^N} \min\limits_{n\in\set{N}}\left(\min\limits_{c\in\set{C}_n}\frac{U_c^{(\text{u})}(\bm{r},\bm{\theta})}{\gamma_c}\right)\nonumber\\ & = \max\limits_{\bm{r}\geq 0, \bm{\theta}\in\Theta^N} \min\limits_{n\in\set{N}} \widehat{U}_n^{(\text{u})}(\bm{r},\bm{\theta}), \mbox{ s.t. } \|\bm{r}\|\leq P^{\text{max}} \label{eqn:maxmin_AO} \end{align} \label{prob:AO} \end{problem} where $\Theta$ denotes the predefined space for antenna tilt configuration. \subsubsection{BS-Based Joint Utility}\label{subsubsec:AO_joinyUtility} It is shown in \eqref{eqn:maxmin_AO} that the cluster-based problem is transformed to the BS-based problem by defining \begin{align} \widehat{U}_n^{(\text{u})}(\bm{r},\bm{\theta})&:=\min_{c\in\set{C}_n}\frac{U_c^{(\text{u})}(\bm{r},\bm{\theta})}{\gamma_c}= \frac{r_n}{\widehat{\set{J}}_n^{\ul}(\bm{r}, \bm{\theta})}\label{eqn:AO_utility_1}\\ \widehat{\set{J}}_n^{\ul}(\bm{r}, \bm{\theta}) &:= \max_{c\in\set{C}_n} \frac{\gamma_c}{\beta_c} \set{J}_c^{\ul}(\bm{r}, \bm{\theta}), \label{eqn:AO_utility_2} \end{align} where $\set{J}_c^{\ul}(\bm{r}, \bm{\theta})$ is obtained from $\set{J}_c^{\ul}(\bm{q}, \bm{b})$ in \eqref{eqn:LB_utility_2} by substituting $\bm{q}$ with $\bm{q}:=\ma{B}_{\ve{\beta}}^T\bm{r}$, and $\tilde{\ma{V}}_{\bm{b}}$ with $\tilde{\ma{V}}_{\bm{\theta}}$. Note that \eqref{eqn:AO_utility_1} is derived by applying the inter-cluster sharing factor such that $r_n:=q_c/\beta_c$ for $n=b_c$. Due to lack of space we omit the details of the individual per BS capacity and coverage utilities corresponding to the cluster-based utilities \eqref{eqn:CL_cap_1} and \eqref{eqn:CL_cov_1}. % % % % \section{Optimization Algorithm}\label{sec:OPAlgor} We developed our optimization algorithm based on the fixed-point iteration algorithm proposed by Yates \cite{yates95}, by exploiting the properties of the monotone and strictly subhomogeneous functions. \subsection{MSS function and Fixed-Point Iteration}\label{subsec:contraction} The vector function $\bm{f}: {\field{R}}_+^K\mapsto {\field{R}}_+^K$ of interest has the following two properties: \begin{itemize} \item {\it Monotonicity}: $\bm{x}\leq \bm{y}$ implies $\bm{f}(\bm{x})\leq\bm{f}(\bm{y})$,. \item {\it Strict subhomogeneity}: for each $\alpha>1, \bm{f}(\alpha \bm{x})<\alpha\bm{f}(\bm{x})$. \end{itemize} A function satisfying the above two properties is referred to be {\it monotonic and strict subhomogeneous (MSS)}. When the strict inequality is relaxed to weak inequality, the function is said to be {\it monotonic and subhomogeneous (MS)}. \begin{theorem}\cite{nuzman2007contraction} Suppose that $\bm{f}: {\field{R}}_+^K\mapsto {\field{R}}_+^K$ is MSS and that $\bm{h}=\bm{x}/l(\bm{x})$, where $l:{\field{R}}_+^K \mapsto {\field{R}}_+$ is MS. For each $\theta>0$, there is exactly one eigenvector $\bm{v}$ and the associated eigenvalue $\lambda$ of $\bm{f}$ such that $l(\bm{v})=\theta$. Given an arbitrary $\theta$, the repeated iterations of the function \begin{equation} \bm{g}(\bm{x})=\theta \bm{f}(x)/l(\bm{f(x)}) \label{eqn:fixedpointiteration} \end{equation} converge to a unique fixed point such that $l(\bm{v})=\theta$. \label{Theoremmapping} \end{theorem} The fixed point iteration in \eqref{eqn:fixedpointiteration} is used to obtain the solution of the following max-min utility balancing problem \begin{equation} \max_{\bm{p}}\min_{k\in\set{K}} U_k(\bm{p}), \mbox{ s.t. } \|\bm{p}\|\leq P^{\text{max}} \label{eqn:prob_maxmin_1} \end{equation} where the utility function can be defined as $U_k(\bm{p}):= p_k/f_k(\bm{p})$. \subsection{Joint Optimization Algorithm}\label{subsec:JointOptAlgor} We aim on jointly optimizing both problems, by optimizing $(\bm{q}, \bm{b})$ in Problem \ref{prob:LB} and $(\bm{r},\bm{\theta})$ in Problem \ref{prob:AO} iteratively with the fixed-point iteration. In the following we present some properties that are required to solve the problem efficiently and to guarantee the convergence of the algorithm. \subsubsection{Decoupled Variables in Uplink} In uplink the variables $\bm{b}$ and $\bm{\theta}$ are decoupled in the interference functions \eqref{eqn:LB_utility_2} and \eqref{eqn:AO_utility_2}, i.e., $\set{J}_c^{\ul}(\bm{q}, \bm{b}):=\set{J}_c^{\ul}(\bm{q}, b_c)$ and $\widehat{\set{J}}_n^{\ul}(\bm{r}, \bm{\theta}):=\widehat{\set{J}}_n^{\ul}(\bm{r}, \theta_n)$. Thus, we can decompose the BS assignment (or tilt optimization) problem into sub-problems that can be independently solved in each cluster (or BS), and the interference functions can be modified as functions of the power allocation only: \begin{align} \set{J}_c^{\ul}(\bm{q})&:=\min_{b_c\in\set{N}} \set{J}_c^{\ul}(\bm{q}, b_c)\label{eqn:modi_inter_1}\\ \widehat{\set{J}}_n^{\ul}(\bm{r})&:=\min_{\theta_n\in\Theta} \widehat{\set{J}}_n^{\ul}(\bm{r}, \theta_n) \label{eqn:modi_inter_2} \end{align} \subsubsection{Standard Interference Function} The modified interference function \eqref{eqn:modi_inter_1} and \eqref{eqn:modi_inter_2} are \textit{standard}. Using the following three properties: 1) an affine function $\bm{\set{I}}(\bm{p}):=\bm{V}\bm{p}+\bm{\sigma}$ is standard, 2) if $\bm{\set{I}}(\bm{p})$ and $\bm{\set{I}}'(\bm{p})$ are standard, then $\beta\bm{\set{I}}(\bm{p})+(1-\beta)\bm{\set{I}}'(\bm{p})$ are standard, and 3) If $\bm{\set{I}}(\bm{p})$ and $\bm{\set{I}}'(\bm{p})$ are standard, then $\bm{\set{I}}^{\text{min}}(\bm{p})$ and $\bm{\set{I}}^{\text{max}}(\bm{p})$ are standard, where $\bm{\set{I}}^{\text{min}}(\bm{p})$ and $\bm{\set{I}}^{\text{max}}(\bm{p})$ are defined as $\set{I}_j^{\text{min}}(\bm{p}):=\min\{\set{I}_j(\bm{p}), \set{I}_j'(\bm{p})\}$ and $\set{I}_j^{\text{max}}(\bm{p}):=\max\{\set{I}_j(\bm{p}), \set{I}_j'(\bm{p})\}$ respectively \cite{yates95}, we can easily prove that \eqref{eqn:modi_inter_1} and \eqref{eqn:modi_inter_2} are standard interference functions. Substituting \eqref{eqn:modi_inter_1} and \eqref{eqn:modi_inter_2} in Problem \ref{prob:LB} and Problem \ref{prob:AO}, define $U_c^{(\text{u})}(\bm{q}):=q_c/\set{I}_c^{(\text{u})}(\bm{q})$ and $U_n^{(\text{u})}(\bm{r}):=r_n/\widehat{\set{J}}_n^{\ul}(\bm{r})$, we can write both problems in the general framework of the max-min fairness problem \eqref{eqn:prob_maxmin_1}: \begin{itemize} \item[]Problem 1. $\max_{\bm{q}\geq 0}\min_{c\in\set{C}} U_c^{(\text{u})}(\bm{q})/\gamma_c, \|\bm{q}\|\leq P^{\text{max}}$. \item[]Problem 2. $\max_{\bm{r}\geq 0}\min_{n\in\set{N}} U_n^{(\text{u})}(\bm{r}), \|\bm{r}\|\leq P^{\text{max}}$ \end{itemize} % The property of the decoupled variables in uplink and the property of utilities based on the standard interference functions enable us to solve each problem efficiently with two iterative steps: 1) find optimum variable $b_c$ (or $\theta_n$) for each cluster $c$ (or each BS $n$) independently, 2) solve the max-min balancing power allocation problem with fixed-point iteration. % \subsubsection{Connections between The Two Problems} Problem \ref{prob:LB} and Problem \ref{prob:AO} have the same objective $C^{(\text{u})}(P^{\text{max}})$ as stated in \eqref{eqn:LB_OP} and \eqref{eqn:maxmin_AO}, i.e., given the same variables $(\hat{\bm{q}}, \hat{\bm{b}}, \hat{\bm{r}}, \hat{\bm{\theta}})$, using \eqref{eqn:AO_utility_1}, we have $\min_{c\in\set{C}} U_c^{(\text{u})}/\gamma_c=\min_{n\in\set{N}} \widehat{U}_n^{(\text{u})}$. Both problems are under the same sum power constraint. However, the convergence of the two-step iteration requires two more properties: 1) the BS power budget $\bm{r}$ derived by solving Problem \ref{prob:AO} at the previous step should not be violated by the cluster power allocation $\bm{q}$ found by optimizing Problem \ref{prob:LB}, and 2) when optimizing Problem \ref{prob:AO}, the inter-cluster power sharing factor $\bm{\beta}$ should be consistent with the derived cluster power allocation $\bm{q}$ in Problem \ref{prob:LB}. To fulfill the first requirement, we introduce the per BS power constraint $P_n^{\text{max}}$ for Problem \ref{prob:AO} equivalent to the BS power budget $r_n$ in Problem \ref{prob:LB}. We also propose a scaled version of fixed point iteration similar to the one proposed in \cite{nuzman2007contraction} to iteratively scale the cluster power vector and achieve the max-min utility boundary under per BS power budget constraints, as stated below. \begin{equation} q_c^{(t+1)} =\frac{\gamma_c\set{I}_c^{(\text{u})}(\bm{q}^{(t)})}{\|\bm{B}\bm{\set{I}}^{(\text{u})}(\bm{q}^{(t)}) \oslash {\bm{P}^{\text{max}}}^{(t)}\|_{\infty}} \label{eqn:FP_LB} \end{equation} where $\oslash$ denote the element-wise division of vectors, $\|\cdot\|_{\infty}$ denotes the maximum norm, ${\bm{P}^{\text{max}}}^{(t)}:=\bm{r}^{(t)}$. To fulfill the second requirement, once $\bm{q}^{(n+1)}$ is derived, the power sharing factors $\bm{\beta}$ need to be updated for solving Problem \ref{prob:AO} at the next step, given by \begin{equation} \bm{\beta}^{(n+1)}:=\bm{Q}^{-1}\bm{B}^T\bm{r}^{(n)}, \mbox{where } \bm{Q}=\mathop{\mathrm{diag}}\{\bm{q}^{(n+1)}\} \label{eqn:FP_LB_beta} \end{equation} The scaled fixed-point iteration to optimize Problem \ref{prob:AO} is provided by \begin{equation} r_n^{(t+1)}= \frac{P^{\text{max}}}{\|\bm{\widehat{\set{I}}}^{(\text{u})}(\bm{r}^{(t)})\|}\cdot \widehat{\set{I}}_n^{(\text{u})}(\bm{r}^{(t)}) \label{eqn:FP_AO_1} \end{equation} % The joint optimization algorithm is given in Algorithm \ref{alg:optim-algor}. % \begin{algorithm}[t]\label{alg:optim-algor} \caption{Joint Optimization of Problem \ref{prob:LB} and \ref{prob:AO}} \begin{algorithmic}[1] \STATE broadcast the information required for computing $\bm{V}$, predefined constraint $P^{\text{max}}$ and thresholds $\epsilon_1,\epsilon_2,\epsilon_3$ \STATE arbitrary initial power vector $\bm{q}^{(t)}>0$ and iteration step $t:=0$ \REPEAT[joint optimization of Problem \ref{prob:LB} and \ref{prob:AO}] \REPEAT[fixed-point iteration for every cluster $c\in\set{C}$] \STATE broadcast $\bm{q}^{(t)}$ to all base stations \FOR{all assignment options $b_c \in \set{N}$} \STATE compute $\set{I}_c^{(\text{u})}(\bm{q}^{(t)}, b_c)$ with \eqref{eqn:LB_utility_2} \ENDFOR \STATE compute $\set{I}_c^{(\text{u})}(\bm{q}^{(t)})$ with \eqref{eqn:modi_inter_1} and update $b_c^{(t+1)}$ \STATE update $q_c^{(t+1)}$ with \eqref{eqn:FP_LB} \STATE $t := t+1$ \UNTIL{convergence: $\bigl| q_c^{(t+1)} - q_c^{(t)}\bigr| / q_c^{(t)} \leq \epsilon_1$} \STATE update $\bm{\beta}^{(t)}$ with \eqref{eqn:FP_LB_beta} % \REPEAT[fixed-point iteration for every BS $n\in\set{N}$] \STATE broadcast $\bm{r}^{(t)}$ to all base stations \FOR{all antenna tilt options $\theta_n \in \Theta$} \STATE compute $\widehat{\set{I}}_n^{(\text{u})}(\bm{r}^{(t)}, \theta_n)$ with \eqref{eqn:AO_utility_2} \ENDFOR \STATE compute $\widehat{\set{I}}_n^{(\text{u})}(\bm{r}^{(t)})$ with \eqref{eqn:modi_inter_2} and update $\theta_n^{(t+1)}$ \STATE update $r_c^{(n+1)}$ with \eqref{eqn:FP_AO_1} \STATE $t := t+1$ \UNTIL{convergence: $\bigl| r_n^{(t+1)} - r_n^{(t)}\bigr| / r_n^{(t)} \leq \epsilon_2$} \STATE update ${P_n^{\text{max}}}^{(t)}:=r_n^{(t)}$ \STATE compute $l^{(t+1)}:=\min_{n\in\set{N}} \widehat{U}^{(\text{u})}_n(\bm{r}^{(n+1)})$ \UNTIL{convergence: $|l^{(t+1)}-l^{(t)}|/l^{(t)}\leq\epsilon_3$} \end{algorithmic} \end{algorithm} % \section{Uplink-Downlink Duality}\label{sec:Duality} We state the joint optimization problem in uplink in Section \ref{sec:ProbForm} and propose an efficient solution in Section \ref{sec:OPAlgor} by exploiting the decoupled property of $\bm{V}$ over the variables $\bm{\theta}$ and $\bm{b}$. The downlink problem, due to the coupled structure of $\bm{V}^T$, is more difficult to solve. As extended discussion we want to address the relationship between the uplink and the downlink problem, and to propose a sub-optimal solution for downlink which can be possibly found through the uplink solution. Let us consider cluster-based max-min capacity utility balancing problem in Section \ref{subsubsec:LB_A} as an example. In the downlink the optimization problem is written as \begin{align} \vspace{-0.2em} \max_{\bm{q}, \bm{b}}\min_c &\frac{U_c^{(\text{d},1)}(\bm{q}, \bm{b})}{\gamma_c}, \mbox{s.t. } \|\bm{q}\|_1\leq P^{\text{max}}\nonumber\\ \mbox{where } & U_c^{(\text{d},1)} :=\frac{q_c}{[\bm{\Psi}\bm{A}\V_{\ve{b}}^T\ma{A}_{\ve{\alpha}}^T\bm{q}+\bm{\Psi}\bm{z}^{(\text{d})}]} \label{eqn:LB_dl} \vspace{-0.2em} \end{align} The cluster-based received noise is written as $\bm{z}^{(\text{d})}:=\bm{A}\bm{\sigma}^{(\text{d})}$. In the following we present a virtual dual uplink network in terms of the feasible utility region for the downlink network in \eqref{eqn:LB_dl} via Perron-Frobenius theory, such that the solution of problem \eqref{eqn:LB_dl} can be derived by solving the uplink problem \eqref{eqn:LB_ul} with the algorithm introduced in Section \ref{sec:OPAlgor}. % \begin{proposition} Define a virtual uplink network where the link gain matrix is modified as $\bm{W}_{\bm{b}}:=\mathop{\mathrm{diag}}\{\bm{\alpha}\}\V_{\ve{b}}\mathop{\mathrm{diag}}^{-1}\{\bm{\alpha}\}$, i.e., $w_{lk}:=v_{lk}\frac{\alpha_l}{\alpha_k}$, and the received uplink noise is denoted by $\bm{\sigma}^{(\text{u})}:=[{\sigma^2_1}^{(\text{u})}, \ldots, {\sigma^2_K}^{(\text{u})}]^T$, where ${\sigma_k^2}^{(\text{u})}:=\frac{\Sigma_{\text{tot}}}{|\set{K}_{c_k}|\cdot C}$ for $k\in\set{K}$, and assume $\Sigma_{\text{tot}}:=\|\bm{\sigma}^{(\text{u})}\|_1=\|\bm{\sigma}^{(\text{d})}\|_1$ (which means, the sum noise is equally distributed in clusters, while in each cluster the noise is equally distributed in the subordinate users). The dual uplink problem of problem \eqref{eqn:LB_dl} is given by \begin{align} \vspace{-0.2em} \max_{\bm{q},\bm{b}}\min_c & \frac{U_c^{(\text{u},1)}(\bm{q}, \bm{b})}{\gamma_c}, \mbox{s.t. } \|\bm{q}\|_1\leq P^{\text{max}}\nonumber\\ \mbox{where } & U_c^{(\text{u},1) }:=\frac{q_c}{[\bm{\Psi}\bm{A}\bm{W}_{\bm{b}}\ma{A}_{\ve{\alpha}}^T\bm{q}+\bm{\Psi}\bm{z}^{(\text{u})}]} \label{eqn:LB_ul} \vspace{-0.2em} \end{align} where $\bm{z}^{(\text{u})}:=\bm{A}\bm{\sigma}^{(\text{u})}$. \label{prop:Duality} \end{proposition} \begin{proof} The proof is given in the Appendix. \end{proof} Note that the optimizer $\bm{b}^{\ast}$ for BS assignment in downlink can be equivalently found by minimizing the spectral radius $\bm{\Lambda^{(u)}(\bm{b})}$ in the uplink. Once $\bm{b}^{\ast}$ is found, the associate optimizer for uplink power ${\bm{q}^{(\text{u})}}^{\ast}$ is given as the dominant right-hand eigenvector of matrix $\bm{\Lambda}^{(\text{u})}(\bm{b}^{\ast})$, while the associate optimizer for downlink power ${\bm{q}^{(\text{d})}}^{\ast}$ is given as the dominant right-hand eigenvector of matrix $\bm{\Lambda}^{(\text{d})}(\bm{b}^{\ast})$. Proposition \ref{prop:Duality} provides an efficient approach to solve the downlink problem with two iterative steps (as the one proposed in \cite{BocheDuality06}): 1) for a fixed power allocation $\hat{\bm{q}}$, solve the uplink problem and derive the assignment $\bm{b}^{\ast}$ that associated with the spectral radius of extend coupling matrix $\bm{\Lambda}^{(\text{u})}$, and 2) for a fixed assignment $\hat{\bm{b}}$, update the power $\bm{q}^{\ast}$ as the solution of \eqref{eqn:DL_matrixEqua}. Although we are able to find a dual uplink problem for the downlink problem in \eqref{eqn:LB_dl} with our proposed utility functions \emph{under sum power constraints}, \insl{we are not able to construct a dual network with decoupled properties for the modified problem \emph{under per BS power constraints} \eqref{eqn:FP_LB}. However, numerical experiments show that our approach to the downlink based on the proposed uplink solution does improve the network performance, although the duality does not exactly hold between the downlink problem and our proposed uplink problem under the per BS power constraints.} % % \section{Numerical Results}\label{sec:Simu} We consider a real-world urban scenario based on a pixel-based mobility model of realistic collection of BS locations and pathloss model for the city of Berlin. The data was assembled within the EU project MOMENTUM and is available at \cite{MOMENTUM}. We select 15 tri-sectored BS in the downtown area. Users are uniformly distributed and are clustered based on their SINR distributions as shown in Fig. \ref{fig:Berlin} (UEs assigned to each sector are clustered into groups and are depicted in distinct colors). The SINR threshold is defined as -6.5 dB and the power constraint per BS is 46dBm. The 3GPP antenna model defined in \cite{3GPP36942} is applied. Fig. \ref{fig:convergence} illustrates the convergence of the algorithm. Our algorithm achieves the max-min utility balancing, and improves the feasibility level $C^{(u)}(P^{\text{max}})$ by each iteration step. In Fig.\ref{fig:cov_cap_mu} we show that the trade-off between coverage and capacity can be adjusted by tuning parameter $\mu$. By increasing $\mu$ we give higher priority to capacity utility (which is proportional to the ratio between total useful power and total interference power), while for better coverage utility (defined as minimum of SINRs) we can use a small value of $\mu$ instead. Fig. \ref{fig:coverage}, \ref{fig:capacity} and \ref{fig:power} illustrate the improvement of coverage and capacity performance and decreasing of the energy consumption in both uplink and downlink systems by applying the proposed algorithm, when the average number of the users per BS is chosen from the set $\{15,20,25,30,35\}$. In Fig. \ref{fig:capacity} we show that the actual average SINR is also improved, although the capacity utility is defined as a lower bound of the average SINR. Fig. \ref{fig:power} illustrate that our algorithm is more energy efficient when comparing with the fixed BS power budget scenario. Compared to the near-optimal uplink solutions, less improvements are observed for the downlink solutions as shown in Fig. \ref{fig:coverage}, \ref{fig:capacity} and \ref{fig:power}. This is because we derive the downlink solution by exploiting an uplink problem which is not exactly its dual due to the individual power constraints (as described in Section \ref{sec:Duality}). However, the sub-optimal solutions still provide significant performance improvements. % % % % % \section{Conclusions and Further Research}\label{sec:con} We present an efficient and robust algorithmic optimization framework build on the utility model for joint optimization of the SON use cases coverage and capacity optimization and load balancing. The max-min utility balancing formulation is employed to enforce the fairness across clusters. We propose a two-step optimization algorithm in the uplink based on fixed-point iteration to iteratively optimize the per base station antenna tilt and power allocation as well as the cluster-based BS assignment and power allocation. We then analyze the network duality via Perron-Frobenius theory, and propose a sub-optimal solution in the downlink by exploiting the solution in the uplink. Simulation results show significant improvements in performance of coverage, capacity and load balancing in a power-efficient way, in both uplink and downlink. In our follow-up papers we will further propose a more complex interference coupling model and the optimization framework where frequency band assignment is taken into account. We will also examine the suboptimality under more general form of power constraints. \begin{figure}[t] \centering \includegraphics[width=.5\textwidth]{BerlinReceivedSignalStrengthMap_v2} \caption{Berlin Scenario.} \label{fig:Berlin} \end{figure} \begin{figure}[ht] \centering \includegraphics[width=.5\textwidth]{convergence} \caption{Algorithm convergence.} \label{fig:convergence} \end{figure} % \begin{figure}[ht] \centering \includegraphics[width=.5\textwidth]{cov_cap_mu} \vspace{-1em} \caption{Trade-off between utilities depending on $\mu$.} \label{fig:cov_cap_mu} \vspace{-1.5em} \end{figure} \begin{figure}[ht] \centering \includegraphics[width=.5\textwidth]{coverage} \caption{Performance of proposed algorithm: coverage.} \label{fig:coverage} \end{figure} \begin{figure}[ht] \centering \includegraphics[width=.43\textwidth]{capacity} \caption{Performance of proposed algorithm: capacity.} \label{fig:capacity} \end{figure} \begin{figure}[!ht] \centering \includegraphics[width=.43\textwidth]{power} \caption{Performance of proposed algorithm: per-BS power budget.} \label{fig:power} \end{figure} % \input{appendices} \subsection*{Acknowledgements} We would like to thank Dr. Martin Schubert and Dr. Carl J. Nuzman for their expert advice. \ifCLASSOPTIONcaptionsoff \newpage \fi % \bibliographystyle{IEEEtran}
3655822d5d2e022f22ca2c790ab64bb902bfd217
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Cloning algorithms are numerical procedures aimed at simulating rare events efficiently, using a population dynamics scheme. In such algorithms, copies of the system are evolved in parallel and the ones showing the rare behavior of interest are multiplied iteratively~\cite{Anderson1975_63_4,Glasserman_1996,2001IbaYukito,GRASSBERGER200264, 4117599,CappeGuillinetal,Forwardinterfacesampling,PhysRevLett.94.018104, delmoral2005,Dean2009562,GuyaderArnaudetal,giardina_direct_2006,lecomte_numerical_2007, JulienNaturephys_2007,tailleur_simulation_2009,giardina_simulating_2011, 1751-8121-46-25-254002_2013,1751-8121-49-20-205002_2016} (See Fig.~\ref{fig:traj}). One of these algorithms proposed by Giardin\`a {\it et al.}~\cite{giardina_direct_2006,lecomte_numerical_2007, JulienNaturephys_2007,tailleur_simulation_2009,giardina_simulating_2011, 1751-8121-46-25-254002_2013,1751-8121-49-20-205002_2016} is used to evaluate numerically the cumulant generating function (a large deviation function, LDF) of additive (or ``time-extensive'') observables in Markov processes~\cite{opac-b1093895,Touchette20091}. It has been applied to many physical systems, including chaotic systems, glassy dynamics and non-equilibrium lattice gas models, and it has allowed the study of novel properties, such as the behavior of breathers in the Fermi-Pasta-Ulam-Tsingou chain~\cite{JulienNaturephys_2007}, dynamical phase transitions in kinetically constrained models~\cite{garrahanjacklecomtepitardvanduijvendijkvanwijland}, and an additivity principle for simple exclusion processes~\cite{PhysRevLett.92.180601,hurtado_large_2010}. While the method has been used widely, there have been fewer studies focusing on the analytical justification of the algorithm. Even though it is heuristically believed that the LDF estimator converges to the correct result as the number of copies $N_c$ increases, there is no proof of this convergence. Related to this lack of the proof, although we use the algorithm by assuming its validity, we do not have any clue how fast the estimator converges as $N_c \rightarrow \infty$. In order to discuss this convergence, we define two types of numerical errors. First, for a fixed finite $N_c$, averaging over a large number of realizations, the LDF estimator converges to an incorrect value, which is different from the desired large deviation result. We call this deviation from the correct value, {\it systematic errors}. Compared with these errors, we also consider the fluctuations of the estimated value. More precisely, for a fixed value of $N_c$, the results obtained in different realizations are distributed around this incorrect value. We call the errors associated to these fluctuations the {\it stochastic errors}. Although both errors are important in numerical simulations, the former one can lead this algorithm to produce wrong results. For example as seen in Ref.~\cite{NemotoBouchetLecomteJack}, the systematic error grows exponentially as a temperature decreases (or generically in the weak noise limit of diffusive dynamics). \begin{figure}[h] \centering \includegraphics[width=1\columnwidth]{FigurePop2} \caption{ \label{fig:traj} Schematic picture illustrating the principle of the population dynamics algorithm. `Clones' (or copies) of the system are prepared and they evolve following a mutation-and-selection process, maintaining the total population constant.} \end{figure} In order to study these errors, we employ a birth-death process~\cite{tagkey2007iii,opac-b1079113} description of the population dynamics algorithm as explained below: We focus on physical systems described by a Markov dynamics~\cite{giardina_direct_2006,giardina_simulating_2011,lecomte_numerical_2007} with a finite number of states $M$, and we denote by $i$ ($i=0,1,\cdots M-1$) the states of the system. This Markov process has its own stochastic dynamics, described by the transition rates $w(i\rightarrow j)$. In population dynamics algorithms, in order to study its rare trajectories, one prepares $N_c$ copies of the system, and simulate these copies according to {\it (i)} the dynamics of $w(i\rightarrow j)$ (followed independently by all copies) and {\it (ii)} `cloning' step in which the ensemble of copies is directly manipulated, \emph{i.e.}, some copies are eliminated while some are multiplied (See Table~\ref{tablecorrespondence_}). Formally, the population dynamics represents, for a \emph{single} copy of the system, a process that does not preserve probability. This fact has motivated the studies of auxiliary processes~\cite{jack_large_2010}, effective processes~\cite{1742-5468-2010-10-P10007} and driven processes~\cite{PhysRevLett.111.120601} to construct modified dynamics (and their approximations~\cite{PhysRevLett.112.090602}) that preserve probability. Different from these methods, in this article, we formulate explicitly the meta-dynamics of the copies themselves by using a stochastic birth-death process. The process preserves probability, and it allows us to study the numerical errors of the algorithm when evaluating LDF. \begin{table*} \begin{center} \caption{\label{tablecorrespondence_} Correspondence between the population dynamics and the birth-death process to describe it.} \begin{tabular}{c||c|c} \vphantom{\Big|} & Population dynamics algorithm & Birth-death process describing \\[-1mm] & & the population dynamics \\[1mm] \hline \hline \vphantom{\Big|} State of the system & $i$ & $n=(n_0,n_2,\cdots,n_{M-1})$ \\ & ($i=0,1,\cdots M-1$) & ($0\leq n_i \leq N_c$ with $\sum_{i}n_i=N_c$) \\[1mm] \hline \vphantom{\Big|} Transition rates & $w(i \rightarrow j)$ & $\sigma(n \rightarrow \tilde n)$ \\[-1mm] & Markov process on states $i$ & Markov process on states $n$ \\[1mm] \hline \vphantom{\Big|} Numerical procedure & Prepare $N_c$ clones and evolve those & Described by the dynamics \\[-1mm] for rare-event sampling & with a mutation-selection procedure & of rates $\sigma(n \rightarrow \tilde n)$ \\[1mm] \hline \end{tabular} \end{center} \end{table*} In this article, we consider the dynamics of the copies as a stochastic birth-death process whose state is denoted $n=(n_0,n_1,n_2,\cdots, n_{M-1})$, where $0\leq n_i\leq N_c$ represents the number of copies which are in state $i$ in the ensemble of copies. We explicitly introduce the transition rates describing the dynamics of $n$, which we denote by $\sigma (n\rightarrow \tilde n)$. We show that the dynamics described by these transition rates lead in general to the correct LDF estimation of the original system $w(i\rightarrow j)$ in the $N_c \rightarrow \infty$ limit. We also show that the systematic errors are of the order $\mathcal O(1/N_c)$, whereas the numerical errors are of the order $\mathcal O(1/(\tau N_c))$ (where $\tau$ is an averaging duration). This result is in clear contrast with standard Monte-Carlo methods, where the systematic errors are always 0. Based on this convergence speed, we then propose a simple interpolation technique to make the cloning algorithm more reliable. Furthermore, the formulation developed in this paper provides us the possibility to compute exactly the expressions of the convergence coefficients, as we do in Sec.~\ref{Section:Demonstrations} on a simple example. The analytical analysis presented in this paper is supplemented with a thorough numerical study in a companion paper~\cite{SecondPart}. In the companion paper, we employ an intrinsically different cloning algorithm, which is the continuous-time population dynamics algorithm, that cannot be studied by the methods presented in this paper (see Sec.~\ref{subsubsec:differenceContinuousTime}). We show in the companion paper \cite{SecondPart} that the validity of the scaling that we derive analytically here is very general. In particular, we demonstrate in practice the efficiency of the interpolation technique in the evaluation of the LDF, irrespective of the details of the population dynamics algorithm. The construction of this paper is as follows. We first define the LDF problem in the beginning of Sec.~\ref{sec:Birth-deathprocess}, and then formulate the birth-death process used to describe the algorithm in Sec.~\ref{subsec:TransitionmatrixAlgorithm}. By using this birth-death process, we demonstrate that the estimator of the algorithm converges to the correct large deviation function in Sec.~\ref{Subsec:DerivationLargeDeviationEstimator}. At the end of this section, in Sec.~\ref{subsec:systemsizeexpansion}, we discuss the convergence speed of this estimator (the systematic errors) and derive its scaling $\sim 1/N_c$. In Sec.~\ref{Sec:largedeviation_largedeviation}, we turn to stochastic errors. For discussing this, we introduce the large deviation function of the estimator, from which we derive that the convergence speed of the stochastic errors is proportional to $1/(\tau N_c)$. In the next section, Sec.~\ref{Section:Demonstrations}, we introduce a simple two-state model, to which we apply the formulations developed in the previous sections. We derive the exact expressions of the systematic errors in Sec.~\ref{subsection:systematicerrors_twostate} and of the stochastic errors in Sec.~\ref{subsection:largedeviationsInPopTwoState}. At the end of this section, in Sec.~\ref{subsec:Different large deviation estimator}, based on these exact expressions, we propose another large deviation estimator defined in the population dynamics algorithm. In the final section, Sec.~\ref{sec:discussion}, we first summarize the result obtained throughout this paper, and then in Sec.~\ref{subsec:Interporationtechnique}, we propose a simple interpolation technique based on the convergence speed of the systematic errors which allows us to devise a better practical evaluation of the LDF. Finally in Sec.~\ref{subsec:openquestions}, we discuss two open questions. \section{Birth-Death Process Describing the Population Dynamics Algorithm} \label{sec:Birth-deathprocess} As explained in the introduction (also see Table~\ref{tablecorrespondence_}), the state of the population is $n=(n_0,n_1,\cdots,n_{M-1})$, where $n_i$ represents the number of clones in the state $i$. The total population is preserved: $\sum_i n_i = N_c$. Below, we introduce the transition rates of the dynamics between the occupations $n$, $\sigma(n\rightarrow \tilde n)$ that describe corresponding large deviations of the original system, where the dynamics of the original system is given by the rates $w(i\rightarrow j)$ as detailed below. As the original system, we consider the continuous-time Markov process in a discrete-time representation. By denoting by $dt$ the time step, the transition matrix $R_{j,i}$ for time evolution of the state $i$ is described as \begin{equation} R_{j,i} = \delta_{i,j} + dt \Big [ w(i\rightarrow j) -\delta_{i,j}\sum_{k}w(i\rightarrow k) \Big ], \label{eq:ratesRij} \end{equation} where we set $w(i\rightarrow i)=0$. The probability distribution of the state $i$, $p_i(t)$, evolves in time as $p_i(t+dt) = \sum_{j} R_{i,j}p_j(t)$. In the $dt\rightarrow 0$ limit, one obtains the continuous-time master equation describing the evolution of $p_i(t)$~\cite{tagkey2007iii,opac-b1079113}. For simplicity, especially for the cloning part of the algorithm, we keep here a small finite $dt$. The reason why we use a discrete-time representation is solely for simplicity of the discussion. The main results can be derived even if we start with a continuous-time representation (see Sec.~\ref{subsubsec:dtDeltat}). For the original dynamics described by the transition matrix~\eqref{eq:ratesRij}, we consider an observable $b_i$ depending on the state $i$ and we are interested in the distribution of its time-averaged value during a time interval $\tau$, defined as \begin{equation} B(\tau ) = \frac{1}{\tau} \sum_{t=0}^{\tau / dt} dt \ b_{i(t)}. \label{eq:defBtau} \end{equation} Here $i(t)$ is a trajectory of the system generated by the Markov dynamics described by $R_{j,i}$. We note that $B(\tau)$ is a path- (or history-, or realization-) dependent quantity. Since $\tau B(\tau)$ is an additive observable, the fluctuations of $B(\tau)$ depending on the realizations are small when $\tau$ is large, but one can describe the large deviations of $B(\tau)$. Those occur with a small probability, and obey a large deviation principle. We denote by ${\rm Prob}(B)$ the distribution function of $B(\tau)$. The large deviation principle ensures that ${\rm Prob}(B)$ takes an asymptotic form ${\rm Prob}(B) \sim \exp (- \tau I(B))$ for large $\tau$, where $I(B)$ is a large deviation function (or `rate function')~\cite{Touchette20091,opac-b1093895}. If the rate function $I(B)$ is convex, the large deviation function is expressed as a Legendre transform of a cumulant generating function (CGF) $\psi(s)$ defined as \begin{equation} \psi(s) = \lim_{\tau \rightarrow \infty} \frac{1}{\tau} \log \left \langle e^{-s \tau B(\tau)} \right \rangle, \label{eq:defpsi} \end{equation} namely: $I(B) = - \inf_{s}\left [ s B + \psi(s) \right ]$. The large deviation function $I(B)$ and this generating function $\psi(s)$ are by definition difficult to evaluate numerically in Monte-Carlo simulations of the original system of transition rates $w(i\rightarrow j)$ (see, for example,~\cite{PhysRevE.92.052104}). To overcome this difficulty, population dynamics algorithms have been developed~\cite{giardina_direct_2006,lecomte_numerical_2007, JulienNaturephys_2007,tailleur_simulation_2009,giardina_simulating_2011, 1751-8121-46-25-254002_2013,1751-8121-49-20-205002_2016}. Here, we describe this population dynamics algorithm by using a birth-death process on the occupation state $n$ allowing us to study systematically the errors in the estimation of $\psi(s)$ within the population dynamics algorithm. We mention that, without loss of generality, we restrict our study to so-called `type-B' observables that do not depend on the transitions of the state~\cite{garrahan_first-order_2009}, \emph{i.e.}~which are time integrals of the state of the system, as in~\eqref{eq:defBtau}. Indeed, as explained for example in Refs.~\cite{giardina_simulating_2011} and~\cite{NemotoBouchetLecomteJack}, one can always reformulate the determination of the CGF of mixed-type observables into that of a type-B variable, by modifying the transition rates of the given system. \subsection{Transition Matrices Representing the Population Dynamics Algorithm} \label{subsec:TransitionmatrixAlgorithm} We denote the probability distribution of the occupation~$n$ at time~$t$ by $P_n(t)$. The time-evolution of this probability is decomposed into three parts. The first one is the original Monte-Carlo dynamics based on the transition rates $w(i\rightarrow j)$. The second one is the cloning procedure of the population dynamics algorithm, which favors or disfavors configurations according to a well-defined rule. The third one is a supplementary (but important) part which maintains the total number of clones to a constant $N_c$. We denote the transition matrices corresponding to these steps by $\mathcal T$, $\mathcal C$ and $\mathcal K$, respectively. By using these matrices, then, the time evolution of the distribution function is given as \begin{equation} P_{n}(t+dt) = \sum_{ \tilde n } \left ( \mathcal K \mathcal C \mathcal T \right )_{n,\tilde n} P_{\tilde n}(t). \label{eq:time_Evolution} \end{equation} We derive explicit expressions of these matrices in the following sub-sections. We also summarize the obtained results in Table~\ref{table_Matrices}. \begin{table*} \begin{center} \caption{\label{table_Matrices} Transition matrices (see Eq.~(\ref{eq:time_Evolution})) describing the birth-death process.} \begin{tabular}{c||c} \vphantom{\bigg|} & Transition matrices \\[2mm] \hline \hline \vphantom{\bigg|} Dynamics (``mutations'') & $ \mathcal T_{\tilde n, n} \equiv \delta_{\tilde n, n} + dt \sum_{i=0}^{M-1} n_i \sum_{j=0, (j\neq i)}^{M-1}w(i \rightarrow j) \left [ \delta _{\tilde n_i,n_i-1} \delta _{\tilde n_j,n_j+1} \ \delta ^{i,j}_{\tilde n,n} \ - \delta _{\tilde n,n} \right ] $ \\[2mm] \hline \vphantom{\bigg|} Cloning (``selection'') & $\mathcal C_{\tilde n, n} = \delta_{\tilde n, n} + s \ dt \sum_{i=0}^{M-1} n_i |\alpha_i| \left [ \delta _{\tilde n_i,n_i + \alpha_i/|\alpha_i|} \ \delta^{i}_{\tilde n, n} - \delta_{\tilde n, n} \right ] + \mathcal O(dt^2)$ \\[2mm] \hline \vphantom{\bigg|} Maintaining $N_c$ & $\mathcal K_{\tilde n,n} = \delta_{\sum n_i,N_c} \delta_{\tilde n, n} + \sum_{k=-1,1} \delta_{\sum_i n_i, N_c + k} \sum_{i=0}^{M-1} \delta_{\tilde n_i, n_i - k} \ \delta^{i}_{\tilde n, n} \ \frac{n_i }{N_c + k} $ \\[2mm] \hline \vphantom{\bigg|} Full process & $ (\mathcal K \mathcal C \mathcal T)_{\tilde n, n} = \delta_{\tilde n, n} + dt \sum_{i=0}^{M-1} n_i \sum_{j=0, (j\neq i)}^{M-1} \left [ w(i \rightarrow j) + s \tilde w_{n}(i \rightarrow j) \right ] \left [ \delta _{\tilde n_i,n_i-1} \delta _{\tilde n_j,n_j+1} \ \delta ^{i,j}_{\tilde n,n} \ - \delta _{\tilde n,n} \right ]$ \\[2mm] & with \ \ \ $\tilde w_n (i \rightarrow j) = \frac{n_j }{N_c} \left [ \alpha_j \delta_{j\in \Omega^{(+)}} \frac{N_c}{N_c + 1} - \alpha_i \delta_{i\in \Omega^{(-)}} \frac{N_c}{N_c - 1} \right ]$ \\[2mm] \hline \end{tabular} \end{center} \end{table*} \subsubsection{Derivation of the Original Dynamics Part, $\mathcal T$} We first consider the transition matrix $\mathcal T$, which describes the evolution of the occupation state $n$ solely due to the dynamics based on the rates $w(i\rightarrow j)$. During an infinitesimally small time step $dt$, the occupation $n=(n_0,n_1,\cdots,n_{M-1})$ changes to $\tilde n = (n_0,n_1,\cdots, n_i-1, \cdots, n_j+1, \cdots, n_{M-1})$ where $0\leq i< M$ and $0\leq j< M$ (for all $i\neq j$). Since there are $n_i$ clones in the state $i$ before the transition, the transition probability of this change is given as $n_i w(i\rightarrow j) dt$. Thus, we obtain \begin{equation} \begin{split} \mathcal T_{\tilde n, n} \equiv \delta_{\tilde n, n} + & dt \sum_{i=0}^{M-1} n_i \sum_{j=0, (j\neq i)}^{M-1}w(i \rightarrow j) \\ & \times \left [ \delta _{\tilde n_i,n_i-1} \delta _{\tilde n_j,n_j+1} \ \delta ^{i,j}_{\tilde n,n} \ - \delta _{\tilde n,n} \right ], \end{split} \end{equation} where $ \delta ^{i,j}_{\tilde n,n}$ is a Kronecker-delta for the indices except for $i,j$: $\delta ^{i,j}_{\tilde n,n} \equiv \prod_{k\neq i,j} \delta_{\tilde n_k,n_k}$. One can easily check that this matrix satisfies the conservation of the probability: $\sum_{\tilde n} \mathcal T_{\tilde n, n}=1$. It corresponds to the evolution of $N_c$ independent copies of the original system with rates $w(i\to j)$. \subsubsection{Derivation of the Cloning Part, $\mathcal C$} \label{subsubsec:CloningRatio} In the population dynamics algorithm (for example the one described in the Appendix A of Ref.~\cite{NemotoBouchetLecomteJack}), at every certain time interval $\Delta t$, one evaluates the exponential factor for all clones, which is equal to $e^{-s\int_t ^{t+\Delta t}\! dt' ~ b_{i(t')}}$ if the clone is in state $(i(t'))_{t'=t}^{t+\Delta t}$ during a time interval $t\leq t' \leq t +\Delta t$. We also call this exponential factor {\it cloning ratio}, because this factor determines whether each clone is copied or eliminated after this time interval. Although the details of how to determine this selection process can depend on the specific type of algorithms, the common idea is that each of the clones is copied or eliminated in such a way that a clone in state $i(t)$ has a number of descendant(s) proportional to the cloning factor on average after this time interval. In order to implement this idea in our birth-death process, we assume this time step $\Delta t$ to be small. For the sake of simplicity, we set this $\Delta t$ to be our smallest time interval $d t$: $\Delta t =dt$. This condition is not mandatory whenever the $\Delta t \rightarrow 0$ limit is taken at the end (see Sec.~\ref{subsubsec:dtDeltat} for the case $\Delta t > dt$). Then, noticing that the time integral $\int_t ^{t+\Delta t} dt' ~ b_{i(t')}$ is expressed as $ dt \ b_{i(t)}$ for small $dt$, we introduce the following quantity for each state $i$ ($i=0,1,2,...,M-1$): \begin{equation} \nu_i \equiv \frac{ n_i e^{-s ~ dt ~ b_i} }{\sum_{j=0}^{M-1} n_j e^{-s ~ dt ~ b_j } } N_c. \label{eq:nufac} \end{equation} Note that there is a factor $n_i$ in front of the exponential function $e^{-s ~ dt ~ b_i}$ which enumerates the number of clones that occupy the state $i$. The quantity $\nu_i$ is aimed at being the number of clones in state $i$ after the cloning process, however, since $\nu_i$ is not an integer but a real number, one needs a supplementary prescription to fix the corresponding integer number of descendants. In general, in the implementation of population dynamics, this integer is generated randomly from the factor $\nu_i$, equal either to its lower or to its upper integer part. The probability to choose either the lower or upper integer part is fixed by imposing that the number of descendants is equal to $\nu_i$ on average. For instance, if $\nu_i$ is equal to $13.2$, then $13$ is chosen with probability $0.8$, and $14$ with probability $0.2$. Generically, $\lfloor \nu_i \rfloor$ and $\lfloor \nu_i \rfloor + 1$ are chosen with probability $1+ \lfloor \nu_i \rfloor - \nu_i$ and $ \nu_i - \lfloor \nu_i \rfloor $, respectively. We note that we need to consider these two possibilities for all indices $i$. We thus arrive at the following matrix: \begin{equation} \begin{split} \mathcal C_{\tilde n, n} \equiv & \sum_{x_0 = 0}^1 \sum_{x_1 = 0}^1 \sum_{x_2 = 0}^1 \cdots \sum_{x_{M-1} = 0}^1 \prod_{i=0}^{M-1} \\ & \times \delta _{\tilde n_i, \lfloor \nu_i \rfloor + x_i} \left [ \left ( \nu_i - \lfloor \nu_i \rfloor \right ) x_i + \left ( 1+ \lfloor \nu_i \rfloor - \nu_i \right ) (1 - x_i) \right ]. \label{eq:transitionMatrixForCopyDef} \end{split} \end{equation} Now, we expand $\mathcal C$ at small $dt$ and we keep only the terms proportional to $\mathcal O(1)$ and $\mathcal O(dt)$, which do not vanish in the continuous-time limit. For this purpose, we expand $\nu_i$ as \begin{equation} \nu_i = n_i \bigg [ 1 + s \ dt \Big ( \sum_j \frac{n_jb_j}{N_c} - b_i \Big ) \bigg ] + \mathcal O(dt^2), \end{equation} where we have used $\sum_i n_i =N_c$. This expression indicates that $\lfloor \nu_i \rfloor$ is determined depending on the sign of $\sum_j n_jb_j / N_c - b_i $, where we assumed $s>0$ for simplicity without loss of generality (because when $s<0$, we can always re-define $-b$ as $b$ to make $s$ to be positive). By denoting this factor by $\alpha_i$, \emph{i.e.} \begin{equation} \alpha_i(n) \equiv \sum_j \frac{n_jb_j}{ N_c} - b_i, \label{def:eqalphai} \end{equation} we thus define the following state-space $\Omega^{(\pm)}(n)$: \begin{equation} \Omega^{(\pm)}(n) = \left \{ \ i \ \big | \ 0\leq i < M \ {\rm and} \ \pm \alpha_i(n) > 0 \right \}. \end{equation} From this definition, for sufficiently small $dt$, we obtain \begin{equation} \left \lfloor \nu_i \right \rfloor = n_i \end{equation} for $i \in \Omega^{(+)}$, and \begin{equation} \left \lfloor \nu_i \right \rfloor = n_i - 1 \end{equation} for $i \in \Omega^{(-)}$. Substituting these results into (\ref{eq:transitionMatrixForCopyDef}) and expanding in $dt$, we obtain (denoting here and thereafter $\alpha_i=\alpha_i(n))$: \begin{equation} \begin{split} \mathcal C_{\tilde n, n} = & \delta_{\tilde n, n} + s \ dt \sum_{i=0}^{M-1} n_i |\alpha_i| \big [ \delta _{\tilde n_i,n_i + \alpha_i/|\alpha_i|} \ \delta^{i}_{\tilde n, n} - \delta_{\tilde n, n} \big ] \\ & + \mathcal O(dt^2), \label{eq:Cdt0limit} \end{split} \end{equation} where $\delta^{i}_{\tilde n, n}$ is a Kronecker delta for the indices except for $i$: $\delta^{i}_{\tilde n, n} = \prod_{k\neq i}\delta_{\tilde n_k,n_k}$. One can easily check that this matrix preserves probability: $\sum_{\tilde n} \mathcal C_{\tilde n, n} = 1$. \subsubsection{Derivation of the Maintaining Part, $\mathcal K$} As directly checked, the operator $\mathcal T$ preserves the total population $\sum_i n_i$. However, the operator representing the cloning $\mathcal C$, does not. In our birth-death implementation, this property originates from the rounding process $\lfloor \nu_i \rfloor$ in the definition of $\mathcal C$: even though $\nu_i$ itself satisfies $\sum_i \nu_i = N_c$, because of the rounding process of $\nu_i$, the number of clones after multiplying by $\mathcal C$ (that is designed to be proportional to $\nu_i$ on average) can change. There are several ways to keep the number $N_c$ of copies constant without biasing the distribution of visited configurations. One of them is to choose randomly and \emph{uniformly} $\delta N_c$ clones from the ensemble, where $\delta N_c$ is equal to the number of excess (resp.~lacking) clones with respect to $N_c$, and to eliminate (resp.~multiply) them. In our birth-death description, we implement this procedure as follows. We denote by $\mathcal K$ the transition matrix maintaining the total number of clones to be the constant~$N_c$. We now use a continuous-time asymptotics $dt\rightarrow 0$. In this limit, from the expression of the transition matrix elements~(\ref{eq:Cdt0limit}), we find that at each cloning step the number of copies of the cloned configuration varies by $\pm 1$ at most. Hence, the total number of clones after multiplying by $\mathcal C$, $\sum_{i} n_i$, satisfies the following inequality \begin{equation} N_c - 1 \leq \sum_{i} n_i \leq N_c +1. \end{equation} Among the configurations $n$ that satisfy this inequality, there are three possibilities, which are $\sum_{i}n_i=N_c$ and $\sum_{i}n_i=N_c\pm 1$. If $n$ satisfies $\sum_{i}n_i=N_c$, we do not need to adjust $n$, while if $n$ satisfies $\sum_{i}n_i=N_c + 1$ (resp.~$\sum_{i}n_i=N_c - 1$), we eliminate (resp.~multiply) a clone chosen randomly and uniformly. Note that, in our formulation, we do not distinguish the clones taking the same state. This means that we can choose one of the occupations $n_i$ of a state $i$ according to a probability proportional to the number of copies $n_i$ in this state. In other words, the probability to choose the state $i$ and to copy or to eliminate a clone from this state is proportional to $ n_i / \sum_{j=0}^{M-1}n_j $. Therefore, we obtain the expression of the matrix $\mathcal K$ as \begin{equation} \begin{split} \mathcal K_{\tilde n,n} = & \delta_{\sum_i\! n_i,N_c} \delta_{\tilde n, n} \\ & + \sum_{k=-1,1} \delta_{\sum_i\!n_i, N_c + k} \sum_{i=0}^{M-1} \delta_{\tilde n_i, n_i - k} \ \delta^{i}_{\tilde n, n} \ \frac{n_i }{N_c + k} \end{split} \label{eq:Kexpression} \end{equation} for $\tilde n$ that satisfies $\sum_i \tilde n_i = N_c$, and $\mathcal K_{\tilde n,n} = 0$ otherwise. \subsubsection{Total Transition, $\mathcal K \mathcal C \mathcal T$ } We write down the matrix describing the total transition of the population dynamics (see Eq.~(\ref{eq:time_Evolution})). From the obtained expressions of $\mathcal K$, $\mathcal C$, $\mathcal T$, we calculate $\mathcal K \mathcal C \mathcal T$ \begin{equation} \begin{split} & (\mathcal K \mathcal C \mathcal T)_{\tilde n, n} = \delta_{\tilde n, n} \\ & + dt \sum_{i=0}^{M-1} n_i \sum_{j=0, (j\neq i)}^{M-1} \left [ w(i \rightarrow j) + s\, \tilde w_{n}(i \rightarrow j) \right ] \\ & \qquad \qquad \times \left [ \delta _{\tilde n_i,n_i-1} \delta _{\tilde n_j,n_j+1} \ \delta ^{i,j}_{\tilde n,n} \ - \delta _{\tilde n,n} \right ], \end{split} \label{eq:KCTexpression} \end{equation} where the population-dependent transition rate $\tilde w_n(i\rightarrow j)$ is given as \begin{equation} \tilde w_n (i \rightarrow j) = \frac{n_j }{N_c} \left [ \alpha_j \delta_{j\in \Omega^{(+)}} \frac{N_c}{N_c + 1} - \alpha_i \delta_{i\in \Omega^{(-)}} \frac{N_c}{N_c - 1} \right ]. \end{equation} The comparison of the expression (\ref{eq:KCTexpression}) with the original part $\mathcal T$ provides an insight into the obtained result. The jump ratio $w(i\rightarrow j)$ in the original dynamics is replaced by $w(i\rightarrow j) + s\, \tilde w_n(i\rightarrow j)$ in the population dynamics algorithm. We note that this transition rate depends on the population $n$, meaning that we cannot get a closed equation for this modified dynamics at the level of the states $i$ in general. We finally remark that the transition matrix $\sigma(n\rightarrow \tilde n)$ for the continuous-time limit is directly derived from (\ref{eq:KCTexpression}) as \begin{equation} \begin{split} & \sigma(n\rightarrow \tilde n) = \sum_{i=0}^{M-1} n_i \sum_{j=0, (j\neq i)}^{M-1} \left [ w(i \rightarrow j) + s \tilde w_{n}(i \rightarrow j) \right ] \\ & \qquad \qquad \qquad \times \left [ \delta _{\tilde n_i,n_i-1} \delta _{\tilde n_j,n_j+1} \ \delta ^{i,j}_{\tilde n,n} \right ]. \label{eq:sigma} \end{split} \end{equation} \subsection{Derivation of the Large Deviation Results \\ in the $N_c\rightarrow \infty$ Asymptotics} \label{Subsec:DerivationLargeDeviationEstimator} In this subsection, we study the $N_c \rightarrow \infty$ limit for the transition matrix of rates $\sigma(n\rightarrow \tilde n)$, and derive the validity of the population dynamics algorithm. \subsubsection{The Estimator of the Large Deviation Function} \label{subsubsec:estimator} One of the ideal implementations of the population dynamics algorithm is as follows: We make copies of each realization (clone) at the end of simulation, where the number of copies for each realization is equal to the exponential weight $e^{-s\tau B(\tau)}$ in Eq.~(\ref{eq:defpsi}) (so that we can discuss an ensemble with this exponential weight without multiplying the probability by it). In this implementation, the number of clones grows (or decays) exponentially proportionally as $\left \langle e^{-s\tau B(\tau)} \right \rangle$ by definition. In real implementations of the algorithm, however, since taking care of an exponentially large or small number of clones can cause numerical problems, one rather keeps the total number of clones to a constant $N_c$ at every time step, as seen in (\ref{eq:nufac}). Within this implementation, we reconstruct the exponential change of the total number of clones as follows: We compute the average of cloning ratio (see the beginning of Sec.~\ref{subsubsec:CloningRatio} for its definition) at each cloning step, and we store the product of these ratios along the cloning steps. At final time, this product gives the empirical estimation of total (unnormalized) population during the whole duration of the simulation~\cite{1751-8121-49-20-205002_2016}, {\it i.e.}~an estimator of $\left \langle e^{-s\tau B(\tau)} \right \rangle$. One thus estimates the CGF $\psi(s)$ given in Eq.~(\ref{eq:defpsi})~\cite{giardina_direct_2006,lecomte_numerical_2007, JulienNaturephys_2007,tailleur_simulation_2009,giardina_simulating_2011, 1751-8121-46-25-254002_2013,1751-8121-49-20-205002_2016} as the logarithm of this reconstructed population, divided by the total time. In our formulation, the average cloning ratio is given as $\sum _i n_i e^{- s dt b_i}/N_c$, and thus the multiplication over whole time interval reads $ \prod_{t=0}^{\tau/dt} \{ n_i(t) e^{- s dt b_i}/N_c\}$. Because we empirically assume that the CGF estimator converges to $\psi(s)$ in the $N_c, \tau \rightarrow \infty$ limit, the following equality is expected to hold in probability 1: \begin{equation} \psi(s) \stackrel{?}{=} \lim_{N_c \rightarrow \infty} \lim_{\tau \rightarrow \infty} \frac{1}{\tau} \sum_{t=0}^{\tau/dt} \ \log \sum _i \frac{n_i(t) e^{- s dt b_i}}{N_c} + O(dt). \label{eq:expected1} \end{equation} Since the dynamics of the population $n$ is described by a Markov process, ergodicity is satisfied, \emph{i.e.},~time averages can be replaced by the expected value with respect to the stationary distribution function. Applying this result to the right-hand side of (\ref{eq:expected1}), we obtain \begin{equation} \begin{split} & \lim_{\tau \rightarrow \infty} \frac{1}{\tau} \sum_{t=0}^{\tau/dt} \ \log \sum _i \frac{n_i(t) e^{- s dt b_i}}{N_c} \\ & = \frac{1}{dt} \sum_{n}P_{n}^{\rm st} \ \log \sum _i \frac{n_i e^{- s dt b_i}}{N_c} + \mathcal O(dt), \end{split} \end{equation} where $P_{n}^{\rm st}$ is the stationary distribution function of the population $n$ in the $dt \rightarrow 0$ limit, (namely, $P_{n}^{\rm st}$ is the stationary distribution of the dynamics of transition rates $\sigma(n\rightarrow \tilde n)$). By expanding this right-hand side with respect to $dt$, we rewrite the expected equality (\ref{eq:expected1}) as \begin{equation} \psi(s) \stackrel{?}{=} - s \lim_{N_c \rightarrow \infty} \sum_{n}P_{n}^{\rm st} \sum_{i}\frac{n_i b_i}{N_c} + O(dt). \label{eq:expected2} \end{equation} where we used that $\sum_i n_i=N_c$ is a conserved quantity. Below we demonstrate that this latter equality (\ref{eq:expected2}) is satisfied by analyzing the stationary distribution function $P_n^{\rm st}$. \subsubsection{The Connection between the Distribution Functions \\ of the Population and of the Original System} From the definition of the stationary distribution function $P_n^{\rm st}$, we have \begin{equation} \sum_{\tilde n} P_{\tilde n}^{\rm st} \sigma(\tilde n\rightarrow n) - \sum_{\tilde n}P_n^{\rm st} \sigma(n\rightarrow \tilde n) = 0, \label{eq:stationaryMaster} \end{equation} (which is a stationary Master equation.) In this equation, we use the explicit expression of $\sigma$ shown in (\ref{eq:sigma}). By denoting by $n^{j \rightarrow i}$ the configuration where one clone in the state $j$ moves to the state $i$: $n^{j \rightarrow i} \equiv (n_0, n_1, \cdots, n_i + 1, \cdots, n_j -1, \cdots, n_{M-1})$, and the stationary master equation (\ref{eq:stationaryMaster}) is rewritten as \begin{equation} \sum_{i,j (i\neq j)}\left [ f_{i \rightarrow j}(n^{j\rightarrow i}) - f_{i \rightarrow j}(n) \right ] = 0, \label{eq:derivation1} \end{equation} where we defined $f_{i \rightarrow j}(n)$ as \begin{equation} f_{i \rightarrow j}(n) = P_{n}^{\rm st} n_i \left [w(i \rightarrow j) + s \tilde w_n(i \rightarrow j) \right ]. \end{equation} Now we multiply expression (\ref{eq:derivation1}) by $n_k$ ($k$ is arbitrary from $k=0,1,2,\cdots, M-1$), and sum it over all configurations $n$: \begin{equation} \sum_{n}\sum_{i,j (i\neq j)} n_k \left [ f_{i \rightarrow j}(n^{j\rightarrow i}) - f_{i \rightarrow j}(n) \right ] = 0. \label{eq:derivation2} \end{equation} We can change the dummy summation variable $n$ in the first term to $n^{i\rightarrow j}$, which leads to $\sum_{n}\sum_{i,j (i\neq j)} (n^{i\rightarrow j})_k f_{i \rightarrow j}(n) $. Since the second term has almost the same expression as the first one except for the factor $n_k$, the sum in (\ref{eq:derivation2}) over the indices $(i,j)$, where none of $i$ nor $j$ is equal to $k$, becomes 0. The remaining term in (\ref{eq:derivation2}) is thus \begin{equation} \begin{split} 0 & = \sum_{n} \sum_{j(j\neq k)} \left ( (n^{k\rightarrow j})_k - n_k\right ) f_{k \rightarrow j}(n) \\ & + \sum_{n} \sum_{i(i\neq k)} \left ( (n^{i\rightarrow k})_k - n_k\right ) f_{i \rightarrow k}(n). \end{split} \end{equation} Using the definition of $n^{i \rightarrow j}$ in this equation, we arrive at \begin{equation} 0 = \sum_{n} \bigg [ \sum_{i(i\neq k)} f_{i \rightarrow k}(n) - f_{k \rightarrow i}(n) \bigg]. \label{eq:derivation3} \end{equation} This equation (\ref{eq:derivation3}) connects the stationary property of the population dynamics (described by the occupation states $n$) and the one in the original system (described by the states $i$). The easiest case where we can see this connection is when $s=0$. By defining the empirical occupation probability of the original system as $p_i \equiv \sum_{n}P_{n}^{\rm st}n_i/N_c$, Eq.~(\ref{eq:derivation3}) leads to the following (stationary) master equation for $w(i\rightarrow j)$: \begin{equation} 0 = \sum _{j}p_j w(j\rightarrow i) - \sum_{j}p_i w(i\rightarrow j) \quad\textnormal{(for $s=0$)} \label{eq:mastereqpiemp} \end{equation} This is valid for any $N_c$, meaning that, for original Monte-Carlo simulations in $s=0$, the empirical probability~$p_i$ is exactly equal to the steady-state probability, as being the unique solution of~\eqref{eq:mastereqpiemp}. It means that there are no systematic errors in the evaluation of $p_i $ (see the introduction of this paper for the definition of the term ``systematic errors''). However, in the generic case $s\neq 0$, this property is not satisfied. One thus needs to understand the $N_c \rightarrow \infty$ limit to connect the population dynamics result with the large deviation property of the original system. \subsubsection{Justification of the Convergence of the Large Deviation Estimator as Population Size becomes Large} In order to take the $N_c \rightarrow \infty$ limit, we define a scaled variable $x_i$ as $n_i/N_c$. With keeping this occupation fractions $x_i$ to be $\mathcal O(1)$, we take the $N_c \rightarrow \infty$ limit in (\ref{eq:derivation3}), which leads to \begin{equation} \begin{split} 0 = \sum_n P_n^{\rm st} & \Bigg [ \sum _{j}x_j w(j\rightarrow i) - \sum_{j}x_i w(i\rightarrow j) \\ & - s ~ x_i \left ( b_i - \sum_k x_k b_ k \right ) \Bigg ] + \mathcal O(1/N_c). \end{split} \label{eq:derivation4} \end{equation} Inspired by this expression, we define a matrix $L_{i,j}^s$ as \begin{equation} L_{i,j}^s = w(j\rightarrow i) - \delta _{i,j} \left (\sum_{k} w(i \rightarrow k) + s ~ b_i\right ), \end{equation} and a correlation function between $x_i$ and $x_j$ as \begin{equation} c_{i,j} = \sum_{n}x_ix_j P_{n}^{\rm st} - p_i p_j, \end{equation} (where we recall $p_i \equiv \sum_n x_i P_{n}^{\rm st}$). From these definitions, (\ref{eq:derivation4}) is rewritten as \begin{equation} \sum_{j}p_j L_{i,j}^{s} = - s p_i \sum_{k} p_kb_k - s \sum_{k}c_{i,k} b_k + \mathcal O \left ( \frac{1}{N_c} \right ). \label{eq:generaleigenvalue} \end{equation} Since $x_i$ is an averaged quantity (an arithmetic mean) with respect to the total number of clones ($x_i\equiv n_i/N_c$), we can safely assume that the correlation $c_{i,j}$ becomes 0 in $N_c \rightarrow \infty$ limit: \begin{equation} \lim_{N_c\rightarrow \infty}c_{i,k}=0. \label{eq:Assumption} \end{equation} (For more detailed discussion of why this is valid, see the description after Eq.~(\ref{eq:systemsizeexpansion})). Thus, by defining $p_i^{\infty}\equiv \lim_{N_c \rightarrow \infty} p_i$, we obtain \begin{equation} \sum_{j}p_j^{\infty} L_{i,j}^{s} = - s p^{\infty}_i \sum_{k} p^{\infty}_kb_k. \end{equation} From the Perron-Frobenius theory, the positive eigenvector of the matrix $L_{i,j}^{s}$ is unique and corresponds to its eigenvector of largest eigenvalue (in real part). This means that $-s \sum_{k}p_{k}^{\infty } b_k$ is the largest eigenvalue of the matrix $L_{i,j}^{s}$. Finally, by recalling that the largest eigenvalue of this matrix $L_{i,j}^{s}$ is equal to the generating function $\psi(s)$ (see Ref.~\cite{garrahan_first-order_2009} for example), we have finally justified that the CGF estimator (\ref{eq:expected2}) is valid in the large-$N_c$ limit. \subsection{Systematic Errors due to Finite $N_c$; Convergence Speed of the Large Deviation Estimator as $N_c\to\infty$} \label{subsec:systemsizeexpansion} In the introduction of this paper, we defined the systematic errors as the deviations of the large deviation estimator from the correct value due to a finite number of clones~$N_c$. From (\ref{eq:expected2}), we quantitatively define this systematic error $\epsilon_{\rm sys} $ as \begin{equation} \epsilon_{\rm sys} \equiv \left | \psi(s) + s \sum_i p_i b_i \right |. \label{eq:systematicdef} \end{equation} From a simple argument based on a system size expansion, we below show that this $\epsilon_{\rm sys}$ is of order $\mathcal O(1/N_c)$. We first show that one can perform a system size expansion (as, \emph{e.g.} in~van Kampen~\cite{tagkey2007iii}) for the population dynamics. In (\ref{eq:derivation1}), by recalling the definition of the vector $x$ as $x=n/N_c$, and by denoting $\tilde P^{\rm st}(x) = P^{\rm st}_{x N_c} $, we obtain \begin{equation} \begin{split} 0 = & \sum_{i,j (i\neq j)}\sum_{r=1}^{\infty}\frac{1}{r !} \frac{1}{N_{c}^r} \left (\frac{ \partial}{\partial x_i} - \frac{\partial}{\partial x_j} \right )^r x_i \tilde P^{\rm st}(x) \\ & \qquad \qquad \quad \times \left [ w(i\rightarrow j) + s \tilde w_n(i\rightarrow j)|_{n=xN_c} \right ]. \label{eq:systemsizeexpansion} \end{split} \end{equation} This indicates that the stochastic process governing the evolution of $x$ becomes deterministic in the $N_c \rightarrow \infty$ limit. The deterministic trajectory for $x$ is governed by a differential equation derived from the sole term $r=1$ in the expansion (\ref{eq:systemsizeexpansion}) (see \emph{e.g.}~Sec.~3.5.3 {\it Deterministic processes - Liouville's Equation} in Ref.~\cite{opac-b1079113} for the detail of how to derive this property). Thus if $x$ converges to a fixed point as $N_c$ increases, {\it which is normally observed in implementations of cloning algorithms}, the assumption $(\ref{eq:Assumption})$ is satisfied. From the expression of $\epsilon_{\rm sys}$, we see that the dependence in~$N_c$ comes solely from $p_i$, which can be calculated from the first order correction of $P^{\rm st}_n$ (at large $N_c$). The equation to determine $P^{\rm st}_n$ is the stationary master equation (\ref{eq:stationaryMaster}) or equivalently, the system-size expansion formula (\ref{eq:systemsizeexpansion}). We expand the jump ratio $w(i\rightarrow j) + s \tilde w_n(i\rightarrow j)$ in (\ref{eq:systemsizeexpansion}) with respect to $1/N_c$ as: \begin{equation} \begin{split} & w(i\rightarrow j) + s \tilde w_n(i\rightarrow j) \\ & = w(i\rightarrow j) + s \tilde w_x^\infty (i \rightarrow j) + \frac{s}{N_c} \delta w_x(i\rightarrow j) + \mathcal O(1/N_c^2), \label{eq:wexpansion} \end{split} \end{equation} where $\tilde w_x^\infty (i \rightarrow j)$ and $\delta w_x (i \rightarrow j)$ are defined as \begin{equation} \tilde w_x^\infty (i \rightarrow j) = x_j \left [\alpha_j \delta_{j\in \Omega^{(+)}} - \alpha_i \delta_{i\in \Omega^{(-)}} \right ] \end{equation} and \begin{equation} \delta w_x (i \rightarrow j) = - x_j \left [\alpha_j \delta_{j\in \Omega^{(+)}} + \alpha_i \delta_{i\in \Omega^{(-)}} \right ]. \end{equation} By substituting (\ref{eq:wexpansion}) into the system-size expansion formula (\ref{eq:systemsizeexpansion}) and performing a perturbation expansion, we find that a first-order correction of $p$ is naturally of order $\mathcal O(1/N_c)$, \emph{i.e.}~$\epsilon_{\rm sys}=\mathcal O(1/N_c)$. For a practical scheme of how to implement this perturbation on a specific example, see Sec.~\ref{subsection:systematicerrors_twostate}. In our companion paper~\cite{SecondPart}, the scaling analysis of the $1/N_c$ correction is shown to hold numerically with the continuous-time cloning algorithm (see Sec.~\ref{subsubsec:differenceContinuousTime}). We also show that the $1/N_c$ correction behavior remains in fact valid at {\it finite time}~\cite{SecondPart}, an open question that remains to be investigated analytically. \begin{table} \begin{center} \caption{\label{Table:Numericalerrors} Magnitudes of the numerical errors} \begin{tabular}{c||c} & Magnitude of errors \\ \hline \hline Systematic errors & $\mathcal O(1/N_c)$ \\ \hline Numerical errors & $\mathcal O(1/(\tau N_c))$ \\ \hline \end{tabular} \end{center} \end{table} \subsection{Remarks} Here, we discuss some remarks on the formulation presented in this section. \subsubsection{Relaxing the Condition $dt=\Delta t$} \label{subsubsec:dtDeltat} In Sec.~\ref{subsubsec:CloningRatio}, we set the discretization time of the process $dt$ to be equal to the time interval for cloning $\Delta t$, and we took the $dt=\Delta t \rightarrow 0$ limit at the end. We note that the condition $\Delta t = dt $ is not necessary if both limits $\Delta t \rightarrow 0$ and $d t \rightarrow 0$ (with $dt<\Delta t$) are taken at the end. This is practically important, because we can use the continuous-time {\it process} to perform the algorithm presented here by setting $dt=0$ first, and $\Delta t \rightarrow 0$ limit afterwards. More precisely, replacing $dt$ by $\Delta t$ in the matrix $\mathcal C$ and $\mathcal K$, we build a new matrix $\mathcal K \mathcal C (\mathcal T^{\Delta t/dt})$. Taking the $dt \rightarrow 0$ limit in this matrix while keeping $\Delta t$ non-infinitesimal (but small), this matrix represents the population dynamics algorithm of a continuous-time process with a finite cloning time interval $\Delta t$. The arguments presented in this section can then be applied in the same way, replacing $dt$ by $\Delta t$. We note that the deviation due to a non-infinitesimal $\Delta t$ should thus appear as $O(\Delta t)$ [see Eq.~(\ref{eq:expected1}) for example]. \subsubsection{A Continuous-time Algorithm Used in the companion paper} \label{subsubsec:differenceContinuousTime} The $\Delta t \rightarrow 0$ limit is the key point in the formulation developed in this section. Thanks to this limit, upon each cloning step, the total number of clones $\sum_{j=0}^{M-1}n_j$ always varies only by $\pm 1$, which makes the expression of the matrices $\mathcal C$ and $\mathcal K$ simple enough to develop the arguments presented in Secs.~\ref{Subsec:DerivationLargeDeviationEstimator} and~\ref{subsec:systemsizeexpansion}. Furthermore, during the time interval $\Delta t$ separating two cloning steps, the configuration is changing {\it at most once}. The process between cloning steps is thus simple, which allows us to represent the corresponding time-evolution matrix as $\mathcal T$ (by replacing $dt$ by $\Delta t$ as explained in Sec.~\ref{subsubsec:dtDeltat} above). Generalizing our analytical study to a cloning dynamics in which the limit $\Delta t\to 0$ is not taken is therefore a very challenging task, which is out of the scope of this paper. However, interestingly, in the companion to this paper~\cite{SecondPart} we observe numerically that our predictions for the finite-time and finite-population scalings are still valid in a different version of algorithm for which $\sum_{j=0}^{M-1}n_j$ can vary by an arbitrary amount --~supporting the hypothesis that the analytical arguments that we present here could be extended to more general algorithms. More precisely, we use a continuous-time version of the algorithm~\cite{lecomte_numerical_2007} to study numerically an observable of `type~A'~\cite{garrahan_first-order_2009}. This version of the algorithm differs from that considered in this paper, in the sense that the cloning steps are separated by {\it non-fixed} non-infinitesimal time intervals. These time intervals are distributed exponentially, in contrast to the fixed ones taken in here (where $\Delta t$ is a constant). This results in an important difference: The effective interaction between copies due to the cloning/pruning procedure is unbounded (it can \emph{a~priori} affect any proportion of the population), while in the algorithm of the present paper, this effective interaction is restricted to a maximum of one cloning/pruning event in the $\Delta t \to 0$ limit. We stress that the $dt\to 0$ limit of the cloning algorithm studied here with a fixed $\Delta t$ \emph{does not yield the continuous-time cloning algorithm}, stressing that these two versions of the population dynamics present essential differences. \section{Stochastic Errors: Large Deviations of the Population Dynamics} \label{Sec:largedeviation_largedeviation} In the previous section, we formulated the population dynamics algorithm as a birth-death process and evaluated the systematic errors (which are the deviation of the large deviation estimator from the correct value) due to a finite number of clones (Table~\ref{Table:Numericalerrors}). In this section, we focus on stochastic errors corresponding to the run-to-run fluctuations of the large deviation estimator within the algorithm, at fixed $N_c$ (see the introduction of this paper for the definition of the terms {\it stochastic errors} and {\it systematic errors}). In order to study stochastic errors, we formulate the large deviation principle of the large deviation estimator. In the population dynamics algorithm, the CGF estimator to measure is the time-average of the average cloning ratio of the population (see Sec.~\ref{subsubsec:estimator}): \begin{equation} \psi_{N_c,\tau}(s) \equiv - s \frac{1}{\tau} \int_{0}^{\tau} dt \sum_{i=0}^{M-1} \frac{n_i(t) b_i}{N_c}. \label{eq:estimetor2} \end{equation} As $\tau$ increases, this quantity converges to the expected value (which depends on $N_c$) with probability 1. However whenever we consider a finite $\tau$, dynamical fluctuations are present, and there is a probability that this estimator deviates from its expected value. Since the population dynamics in the occupation states $n$ is described by a Markov process, the probability of these deviations are themselves described by a large deviation principle~\cite{Touchette20091,opac-b1093895}: By denoting by $\rm Prob (\psi)$ the probability of $\psi_{N_c,\tau}(s)$, one has: \begin{equation} {\rm Prob} (\psi) \sim \exp \left ( - \tau I_{N_c,s}(\psi) \right ), \end{equation} where $I_{N_c,s}(\psi)$ is a large deviation ``rate function'' (of the large deviation estimator). To study these large deviations, we can apply a standard technique using a biased evolution operator for our population dynamics: For a given Markov system, to calculate large deviations of additive quantities such as~\eqref{eq:estimetor2}, one biases the time-evolution matrix with an exponential factor~\cite{opac-b1093895}. Specifically, by defining the following matrix \begin{equation} L^{h}_{\tilde n,n} = \sigma(n \rightarrow \tilde n) - \delta_{\tilde n, n} \sum_{n^{\prime}}\sigma(n \rightarrow n^{\prime}) - h s \sum_{i=0}^{M-1} \frac{n_i b_i}{N_c}. \label{eq:DefinitionLh} \end{equation} and by denoting the largest eigenvalue of this matrix $G(h,s)$ (corresponding, as a function of $h$, to a scaled cumulant generating function for the observable~\eqref{eq:estimetor2}), the large deviation function $I_{N_c,s}(\psi)$ is obtained as the Legendre transform $\sup_{h} \left [h \psi - G(h,s) \right ]$. In the companion paper~\cite{SecondPart}, we show that a quadratic approximation of the rate function~$I_{N_c,s}(\psi)$ (\emph{i.e.}~a Gaussian approximation) can be estimated directly from the cloning algorithm. We consider the scaling properties of $I_{N_c,s}$ in the large-$N_c$ limit. For this, we define a scaled variable $\tilde h \equiv h /N_c$ and a scaled function $\tilde G(\tilde h,s) \equiv G(\tilde h N_c,s)/N_c$. If this scaled function $\tilde G(\tilde h,s) \equiv G(\tilde h N_c,s)/N_c$ is well-defined in the $N_c \rightarrow \infty$ limit (which is natural as checked in the next paragraph), then we can derive that $I_{N_c,s}$ has the following scaling: \begin{equation} I_{N_c,s}(\psi) = N_c I_{s}(\psi) + o(N_c) \label{eq:scalingNcINcs} \end{equation} or equivalently, \begin{equation} {\rm Prob}(\psi) \sim e^{-\tau N_c I_{s}(\psi)}, \label{eq:largedeviation_largedeviation} \end{equation} where $I_{s}(\psi) = \max_{\tilde h} \left [ \tilde h \psi - \tilde G(\tilde h,s) \right ]$. The scaling form~\eqref{eq:scalingNcINcs} is validated numerically in Ref.~\cite{SecondPart}. From this large deviation principle, we can see that the stochastic errors of the large deviation estimator is of $\mathcal O(1/(N_c \tau))$ as shown in Table~\ref{Table:Numericalerrors}. In the largest eigenvalue problem for the transition matrix (\ref{eq:DefinitionLh}), by performing a system size expansion (see Sec.~\ref{subsec:systemsizeexpansion}), we obtain \begin{equation} \begin{split} \tilde G(\tilde h,s) = & \sum_{i,j (i\neq j)} \left (\frac{ \partial}{\partial x_i} - \frac{\partial}{\partial x_j} \right ) x_i q(x) \\ & \qquad \qquad \times \left [ w(i\rightarrow j) + s \tilde w_x^{\infty}(i\rightarrow j) \right ] \\ & - \frac{\tilde h}{s} s \sum_i x_i b_i q(x) + \mathcal O(1/N_c), \label{eq:systemsizeexpansion2} \end{split} \end{equation} where $q(x)$ is the right-eigenvector associated to the largest eigenvalue of $L_{\tilde n,n}^h$ (represented as a function of $x\equiv n/N_c$). The first order of the right-hand side is of order $\mathcal O(N_c^0)$, so that $\tilde G(\tilde h,s)$ is also of order $\mathcal O(N_c^0)$ in $N_c \rightarrow \infty$. (For an analytical example of the function $\tilde G(\tilde h, s)$, see Sec.~\ref{subsection:largedeviationsInPopTwoState}). \section{Example: \\ A Simple Two-State Model} \label{Section:Demonstrations} In this section, to illustrate the formulation that we developed in the previous sections, we consider a simple two state model. In this system, the dimension of the state $i$ is two ($M=2$) and the transition rates $w( i \rightarrow j)$ are \begin{equation} w(0 \rightarrow 1) = c, \end{equation} \begin{equation} w(1 \rightarrow 0) = d \end{equation} with positive parameters $c, d$ and $w(i \rightarrow i) = 0$. In this model, the quantity $\alpha_{i}$ defined in (\ref{def:eqalphai}) becomes \begin{equation} \alpha_{i} = \delta_{i,0} \frac{n_1}{N_c}\left (b_1 - b_0 \right ) + \delta_{i,1} \frac{n_0}{N_c} (b_0 - b_1). \end{equation} Hereafter, we assume that $b_1>b_0$ without loss of generality. From this, the space $\Omega^{(\pm)}$ is determined as $\Omega^{(+)}=\{ 0 \}$ and $\Omega^{(-)}=\{ 1 \}$, which leads to the jump ratio $\tilde w_{n}(i \rightarrow j)$ as \begin{equation} \tilde w_{n}(i \rightarrow j) =\delta_{i,1}\delta_{j,0} \frac{n_0}{N_c}(b_1 - b_0) \left [\frac{n_1}{N_c+1} + \frac{n_0}{N_c -1} \right ]. \end{equation} Finally, from the conservation of the total population: $n_0+n_1=N_c$, we find that the state of the population $n$ can be uniquely determined by specifying only the variable $n_0$. Thus the transition rate for the population dynamics is a function of $n_0$ (and $\tilde n_0$), $\sigma (n_0 \rightarrow \tilde n_0)$, which is derived as \begin{equation} \begin{split} & \sigma(n_0\rightarrow \tilde n_0) = \delta_{\tilde n_0, n_0 + 1 } \bigg [ (N_c - n_0) d \\ &+ k(n_0,N_c - n_0) \Big( \frac{n_0}{N_c -1} + \frac{N_c - n_0}{N_c + 1} \Big) \bigg ] + \delta_{\tilde n_0, n_0 - 1 } \ n_0 \,c, \end{split} \end{equation} where we have defined \begin{equation} k(n_0,n_1) = \frac{n_0 n_1}{N_c} s \left [b_1 - b_0 \right ]. \end{equation} \subsection{Systematic Errors} \label{subsection:systematicerrors_twostate} We first evaluate the systematic errors (see Sec.~\ref{subsec:systemsizeexpansion}). For this, we consider the distribution function $P_{n}^{\rm st}$. Since the system is described by a one dimensional variable $n_0$ restricted to $0\leq n_0 \leq N_c$, the transition rates $\sigma(n_0 \rightarrow \tilde n_0)$ satisfy the detailed balance condition: \begin{equation} P_{n_0}^{\rm st} \sigma(n_0 \rightarrow n_0 + 1) = P_{n_0+1}^{\rm st} \sigma(n_0 + 1 \rightarrow n_0). \end{equation} We can solve this equation exactly, but to illustrate the large-$N_c$ limit, it is in fact sufficient to study the solution in an expansion $1/N_c\ll 1$. The result is \begin{equation} P^{\rm st}_{x N_c} = C \exp \left [ - N_c I_{\rm conf}(x) + \delta I(x) + \mathcal O(1/N_c) \right ] \end{equation} (with here $x\equiv n_0 / N_c$), where, explicitly \begin{equation} \begin{split} I_{\rm conf}(x) = & x + \log(1-x) - \frac{d \log \left [d+(b_1-b_0) s x\right ]}{(b_1-b_0) s} \\ &- x \log \left [ \frac{1}{cx}(1-x)\left (d + (b_1 - b_0) s x \right ) \right ] \end{split} \end{equation} and \begin{equation} \begin{split} \delta I(x) & = - x - \frac{2 d x}{(b_1-b_0) s} + x^2 - \log x \\ & + \frac{2 d^2 \log \left [ d + (b_1 - b_0)s x\right ] }{(b_1-b_0)^2 s^2} + \frac{d \log \left[ d + (b_1-b_0) s x \right ]}{(b_1-b_0)s}. \end{split} \end{equation} We now determine the value of $x$ that minimizes $ - N_c I_s(x) + \delta I(x)$, which leads to a finite-size correction (\emph{i.e.}~the systematic errors) of the population dynamics estimator. Indeed, denoting this optimal value of $x$ by $x_{N_c}^*$, the large deviation estimator is obtained as \begin{equation} \psi_{N_c}(s) = - s \left [ x^*_{N_c} b_0 +(1-x^*_{N_c})b_1 \right ] \end{equation} (see Sec.~\ref{subsubsec:estimator}). From a straightforward calculation based on the expressions $I_{\rm conf}(x)$ and $\delta I(x)$, we obtain the expression of $x_{N_c}^*$ as \begin{equation} x^*_{N_c} = x^{*} + \frac{1}{N_c} \delta x^* + \mathcal O((1/N_c)^2), \end{equation} with \begin{equation} \begin{split} x^{*} & = \frac{-c - d + (b_1-b_0) s }{2 (b_1 - b_0) s} \\ & + \frac{ \sqrt{4 d (b_1-b_0) s + \left [-c-d + (b_1-b_0)s \right ]^2}}{2 (b_1 - b_0) s} \end{split} \label{eq:xstar} \end{equation} and \begin{equation} \begin{split} \delta x ^* = & \left (2 d + 2 (b_1 - b_0) s x^* \right )^{-1} \\ & \times \frac{2 c \left [ - d - (b_1 - b_0) s x^* \left ( 1 + x^* - 2 (x^*)^2 \right ) \right ]}{\sqrt{4 d (b_1-b_0)s + [c+d-(b_1-b_0)s ]^2} }. \end{split} \end{equation} We thus arrive at \begin{equation} \begin{split} \psi(s) = & \frac{-c - d - (b_1 + b_0) s }{2 } \\ & + \frac{ \sqrt{4 d (b_1-b_0) s + \left [-c-d + (b_1-b_0)s \right ]^2}}{2 } \label{eq:PsiInfinitetime_InfiniteCopy} \end{split} \end{equation} and \begin{equation} \begin{split} \epsilon _{\rm sys} = & \frac{1}{N_c} \frac{1}{\left | d+(b_1-b_0)sx^*\right |} \\ & \times \left | \frac{s c(b_0-b_1) \left ( d + (b_0 - b_1) s (x^* - 1) x^* (1+2x^*) \right ) }{ \sqrt{4(b_1-b_0)d s + [c+d +(b_0 - b_1)s]^2}} \right | \end{split} \end{equation} (see Eq.~(\ref{eq:systematicdef}) for the definition of the systematic error $\epsilon_{\rm sys}$.) We check easily that the expression of $\psi(s)$ is the same as the one obtained from a standard method by solving the largest eigenvalue problem of a biased time-evolution operator (see for example, Ref.~\cite{1751-8121-49-20-205002_2016}). \subsection{Stochastic Errors} \label{subsection:largedeviationsInPopTwoState} We now turn our attention to the stochastic errors. The scaled cumulant generating function $N_c \tilde G(\tilde h, s)$ is the largest eigenvalue of a matrix $L_{\tilde n,n}^{h}$ (see Eq.~(\ref{eq:DefinitionLh}) and the explanations around it). We then recall a formula to calculate this largest eigenvalue problem from the following variational principle: \begin{equation} \begin{split} &\tilde G(\tilde h,s) \\ & = \sup_{\phi>0} \sum_{n} p_{\rm st}(n_0)\phi(n_0)^2 \Bigg [ \frac{\sigma(n\rightarrow n+1)}{N_c} \left ( \frac{\phi(n_0 + 1)}{\phi(n_0)} - 1 \right ) \\ & + \frac{\sigma(n\rightarrow n-1)}{N_c } \left ( \frac{\phi(n_0 - 1)}{\phi(n_0)} - 1 \right ) - s \tilde h \frac{\sum_{i}n_i b_i}{N_c^2} \Bigg ]. \end{split} \end{equation} (See, \emph{e.g.}, Appendix G of Ref.~\cite{PhysRevE.84.061113} or Ref.~\cite{garrahan_first-order_2009} for the derivation of this variational principle). By following the usual route to solve such equations (see, \emph{e.g.}, Sec.~2.5 of Ref.~\cite{1742-5468-2014-10-P10001}), we obtain \begin{equation} \begin{split} \tilde G(\tilde h,s) = \sup_{x} & \Bigg [ - \left ( \sqrt{(1-x)(d+(b_1 - b_0)s x)} - \sqrt{c x} \right )^2 \\ & - s \tilde h \left [ x b_0 + (1-x) b_1 \right ] \Bigg ]. \end{split} \end{equation} Thus, $\tilde G(\tilde h,s)$ is well-defined, demonstrating that the large deviation principle (\ref{eq:largedeviation_largedeviation}) is satisfied. Furthermore, by expanding this variational principle with respect to $\tilde h $, we obtain \begin{equation} \tilde G(\tilde h,s) =\psi (s) \tilde h + \frac{\kappa_s}{2} \tilde h ^2 + \mathcal O(\tilde h^3), \label{eq:expansionGtilde} \end{equation} where $ \psi (s)$ is given in (\ref{eq:PsiInfinitetime_InfiniteCopy}), and the variance $\kappa_s$ is given as \begin{equation} \begin{split} \kappa_s = & c + \frac{c s(b_1 - b_0)}{\sqrt{4(b_1-b_0) s d + (c+d+(b_0 - b_1)s)^2}} \\ & - \frac{ c(c+d)^2 + c (b_0 - b_1)(c-3d)s }{c^2 + 2 c \left [ d + (b_0 - b_1) s \right ] + (d+(b_1-b_0)s)^2}. \end{split} \end{equation} We note that the expansion (\ref{eq:expansionGtilde}) is equivalent to the following expansion of the large deviation function $I_{s}(\psi)$ (see (\ref{eq:largedeviation_largedeviation})) around the expected value $\psi (s)$: \begin{equation} I_{s}(\psi) = \frac{(\psi - \psi(s))^2}{2\kappa_s} + \mathcal O ((\psi - \psi_s)^3). \end{equation} The variance of the obtained large deviation estimator is thus $\kappa_s/(N_c \tau)$. \subsection{A Different Large Deviation Estimator} \label{subsec:Different large deviation estimator} As an application of these exact expressions, we expand the systematic error $\epsilon _{\rm sys}$ and the stochastic error (variance) $\kappa_s$ with respect to $s$. A straightforward calculation leads to \begin{equation} \begin{split} \epsilon _{\rm sys} N_c = &\Bigg | \frac{2c(b_0 - b_1)}{c+d} s \Bigg | + \mathcal O(s^2) \label{eq:epsilon_expansion} \end{split} \end{equation} and \begin{equation} \kappa_s = \frac{2 (b_0 - b_1)^2 c d }{(c+d)^3} s^2 + \mathcal O(s^3). \label{eq:kappa_expansion} \end{equation} We thus find that the first-order of the error $\epsilon _{\rm sys}$ scales as $\mathcal O(s)$ at small $s$, but that the variance $\kappa_s$ is of order $\mathcal O(s^2)$. From this scaling, as we explain below, one can argue that the following large deviation estimator can be better than the standard one for small $s$: \begin{equation} \tilde \Psi(s) \equiv \frac{1}{\tau} \log \overline{ \prod_{t=0}^{\tau / dt} \sum _i \frac{n_i(t) e^{- s dt b_i}}{N_c} }, \label{Definition_Psi_s} \end{equation} where the overline represents the averaging with respect to the realizations of the algorithm. (Normally, this realization-average is taken \emph{after} calculating the logarithm, which corresponds to the estimator (\ref{eq:expected1}).) Mathematically, this average (\eqref{Definition_Psi_s}, before taking the logarithm) corresponds to a bias of the time-evolution matrix $\sigma$ as seen in (\ref{eq:DefinitionLh}) for $h=1$. This means that, in the limit $\tau \rightarrow \infty$ with a sufficiently large number of realizations, this averaged value behaves as $\tilde \Psi(s) \sim e^{\tau G(1,s)}$. By combining this result with the expansion (\ref{eq:expansionGtilde}), we thus obtain \begin{equation} \begin{split} \lim_{\tau \rightarrow \infty} \lim_{\substack{ \text{many} \\ \text{realizations}}} \tilde \Psi(s) = \psi(s) + \frac{\kappa_s}{2} N_c^{-1} + \mathcal O(N_c^{-2}) \label{eq:G(1s)} \end{split} \end{equation} (recalling $\tilde G = G/N_c$ and $\tilde h = h/N_c$). When we consider small $s$, by recalling $\epsilon_{\rm sys}N_c=\mathcal O(s)$ and $\kappa_s=\mathcal O(s^2)$, we thus find that the deviations from the correct value are smaller in the estimator $\tilde \Psi(s)$ than in the normal estimator given in (\ref{eq:expected1}), which comes as a surprise because in~\eqref{Definition_Psi_s} the average and the logarithm are inverted with respect to a natural definition of the CGF estimator. To use this estimator, we need to discuss the two following points. First, since the scaled cumulant generating function $G(1,s)$ has small fluctuations, one needs a very large number of realizations in order to attain the equality (\ref{eq:G(1s)}). The difficulty of this measurement is the same level as the one of direct observations of a large deviation function, see for example Ref.~\cite{PhysRevE.92.052104}. However, we stress that this point may not be fatal in this estimator, because we do not need to attain completely this equality, \emph{i.e.}~our aim is the zero-th order coefficient, $\psi(s)$, in (\ref{eq:G(1s)}). Second, we have not proved yet the scaling properties with respect to $s$, which are $\epsilon_{\rm sys}N_c=\mathcal O(s)$ and $\kappa_s=\mathcal O(s^2)$, in a general set-up aside from this simple two state model. We show in practice in Ref.~\cite{SecondPart} that for small values of~$s$, the estimator~\eqref{Definition_Psi_s} is affected by smaller systematic errors, in the numerical study of the creation-annihilation process studied in this section. We will focus on the generality of our observations on these points in a future study. \section{Discussion} \label{sec:discussion} In this paper, we formulated a birth-death process that describes population dynamics algorithms and aims at evaluating numerically large deviation functions. We derived that this birth-death process leads generically to the correct large deviation results in the large limit of the number of clones $N_c \rightarrow \infty$. From this formulation, we also derived that the deviation of large deviation estimator from the desired value (which we called systematic errors) is small and proportional with $\mathcal O(N_c^{-1})$. Below, based on this observation, we propose a simple interpolation technique to improve the numerical estimation of large deviation functions in practical uses of the algorithm. \subsection{An Interpolation Technique using the $\mathcal O(1/N_c)$ Scaling of the Systematic Error} \label{subsec:Interporationtechnique} Imagine that we now apply the population dynamics algorithm to a given system. We need to carefully consider the asymptotic limit of the two large parameters $\tau$ and $N_c$ in the convergence of the large deviation estimator (\ref{eq:estimetor2}). Indeed, what one needs to do in this simulation is, (\emph{i}) take the large-$\tau$ limit for a fixed $N_c$ and estimating the $\tau \rightarrow \infty$ value of the estimator for this fixed $N_c$, and then (\emph{ii}) estimate this large-$\tau$ value for several (and increasing) $N_c$, and finally estimate large-$\tau$-$N_c$ limit value. This is different from standard Monte-Carlo simulations, where one needs to consider only the large-$\tau$ limit, thanks to ergodicity. Any method that can make the LDF estimation easier thus will be appreciated. Based on our observations, we know that the second part [(\emph{ii}) above] converges with an error proportional to $1/N_c$. Also, from the large deviation estimator (\ref{eq:estimetor2}), one can easily see that the convergence speed with respect to $\tau$ for a fixed $N_c$ is proportional to $1/\tau$ (\emph{i.e.}~the first part [(\emph{i}) above] converges proportionally to $1/\tau$). By using these $1/\tau$- and $1/N_c$-scalings for (\emph{i}) and (\emph{ii}), one can interpolate the large-$\tau$ and large-$N_c$ asymptotic value of the LDF estimator from the measured values for finite $\tau$ and $N_c$. We introduce this numerical method in practice in the companion paper~\cite{SecondPart}. We demonstrate numerically that the interpolation technique is very efficient in practice, by a direct comparison of the resulting estimation of the CGF to its analytical value, which is also available in the studied system. We also stress that it is developed for a different cloning algorithm by using a continuous-time population dynamics~\cite{lecomte_numerical_2007} (see Sec.~\ref{subsubsec:differenceContinuousTime} for the description of the conceptual difference). From these results, we conjecture that the validity of the large-$\tau$ and large-$N_c$ scalings is very general and independent of the details of the algorithm. \bigskip \subsection{Open Questions} \label{subsec:openquestions} We mention two open questions. The first question is about the precise estimate of the error due to a non-infinitesimal time interval $\Delta t$ between cloning steps: As explained in Secs.~\ref{subsubsec:dtDeltat} and \ref{subsubsec:differenceContinuousTime}, taking the $\Delta t \rightarrow 0$ limit is important in our analysis, in order to make the estimator converge to the correct LDF. The error due to non-infinitesimal $\Delta t$ is at most of order $\Delta t$ as seen from Eq.~(\ref{eq:expected1}) (see also Sec.~\ref{subsubsec:dtDeltat}). From a practical point of view, taking this limit can, however, be problematic, since it requires infinitely many cloning procedures per unit time (as $\Delta t\to 0$). Interestingly, most of existing algorithms do not take such a limit (see for instance the original version of the algorithm~\cite{giardina_direct_2006}). Empirically, one thus expects that the error goes to zero as $N_c\to\infty$ while keeping $\Delta t$ finite. Within the method developed in this paper, the analytical estimation of this error is challenging (see Sec.~\ref{subsubsec:differenceContinuousTime}) and remains an open problem, but for example, one can approach to this issue numerically at least. The second question is about possible extensions of the formulation developed in this paper. In our algorithm, we perform a cloning procedure for a fixed time interval, which means that our formulation cannot cover the case of algorithms where $\Delta t$ itself is statistically distributed, as in continuous-time cloning algorithms~\cite{lecomte_numerical_2007}. Moreover, our formulation is limited to Markov systems, although population dynamics algorithms are applied to chaotic deterministic dynamics~\cite{JulienNaturephys_2007,1751-8121-46-25-254002_2013} or to non-Markovian evolutions~\cite{Non-MarkovPD}. Once one removes the Markov condition in the dynamics, developing analytical approaches becomes more challenging. However, as the physics of those systems are important scientifically and industrially~\cite{PhysRevLett.116.150002}, the understanding of such dynamics cannot be avoided for the further development of population algorithms. \begin{acknowledgments} T.~N.~gratefully acknowledges the support of Fondation Sciences Math\'ematiques de Paris -- EOTP NEMOT15RPO, PEPS LABS and LAABS Inphyniti CNRS project. E.~G.~thanks Khashayar Pakdaman for his support and discussions. Special thanks go to the Ecuadorian Government and the Secretar\'ia Nacional de Educaci\'on Superior, Ciencia, Tecnolog\'ia e Innovaci\'on, SENESCYT, for support. V.~L.~acknowledges support by the National Science Foundation under Grant No.~NSF PHY11-25915 during a stay at KITP, UCSB and support by the ANR-15-CE40-0020-03 Grant LSD. V.~L.~acknowledges partial support by the ERC Starting Grant No. 680275 MALIG. T.~N.~and V.~L.~are grateful to B.~Derrida and S.~Shiri for discussions. We are grateful to S. Shiri for discussions about the first paragraph of Sec.~\ref{subsec:openquestions} (open questions). \end{acknowledgments}
a00517b16ddf030f7acc99bf33b2cb34a44f31f4
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{0pt}{12pt plus 4pt minus 2pt}{2pt plus 2pt minus 2pt} \usepackage{titling} \posttitle{\par\end{center}} \predate{} \postdate{} \setlength{\droptitle}{-40pt} \renewcommand{\abstractname}{\vspace{-\baselineskip}} \DeclareFontFamily{U}{mathb}{\hyphenchar\font45} \DeclareFontShape{U}{mathb}{m}{n}{ <5> <6> <7> <8> <9> <10> gen * mathb <10.95> mathb10 <12> <14.4> <17.28> <20.74> <24.88> mathb12 }{} \DeclareSymbolFont{mathb}{U}{mathb}{m}{n} \DeclareFontFamily{U}{matha}{\hyphenchar\font45} \DeclareFontShape{U}{matha}{m}{n}{ <5> <6> <7> <8> <9> <10> gen * matha <10.95> matha10 <12> <14.4> <17.28> <20.74> <24.88> matha12 }{} \DeclareSymbolFont{matha}{U}{matha}{m}{n} \DeclareMathSymbol{\oasterisk}{3}{matha}{"66} \DeclareMathSymbol{\boxasterisk}{3}{mathb}{"66} \usepackage{graphicx} \newcommand{\ket}[1]{\left|#1\right\rangle} \newcommand{\cket}[1]{\left|\widetilde{#1}\right\rangle} \newcommand{\bra}[1]{\left\langle #1\right|} \newcommand{\cbra}[1]{\left\langle \widetilde{#1}\right|} \newcommand{\ketbra}[1]{\ket{#1}\bra{#1}} \newcommand{\cketbra}[1]{\cket{#1}\cbra{#1}} \newcommand{\bracket}[2]{\langle #1|#2\rangle} \newcommand{{\bar{\bf{x}}}}{{\bar{\bf{x}}}} \newcommand{{\hat{\bf{x}}}}{{\hat{\bf{x}}}} \newcommand{{\bar{\bf{z}}}}{{\bar{\bf{z}}}} \newcommand{{\hat{\bf{z}}}}{{\hat{\bf{z}}}} \newcommand{P}{P} \newcommand{F}{F} \newcommand{Q}{Q} \newcommand{{\text{Tr}}}{{\text{Tr}}} \newcommand{\mathbbm 1}{\mathbbm 1} \newcommand{\pi}{\pi} \newcommand{H_K}{H_K} \newcommand{\hat H_K}{\hat H_K} \newcommand{D_K}{D_K} \newcommand{\mathrm{acos}\,}{\mathrm{acos}\,} \newtheorem{theorem}{Theorem} \newtheorem{corollary}{Corollary} \newtheorem{prop}{Proposition} \newtheorem{lemma}{Lemma} \begin{comment} \usepackage{tikz} \usetikzlibrary{backgrounds,fit,arrows,decorations.pathreplacing,positioning} \usetikzlibrary{external} \tikzexternalize \end{comment} \usepackage{xfrac} \begin{document} \author{ {\normalsize Joseph M.~Renes}\\ \emph{\normalsize Institute for Theoretical Physics, ETH Zurich, 8093 Z\"urich, Switzerland} } \title{\large {\bf Belief propagation decoding of quantum channels by passing quantum messages}} \date{\vspace{-\baselineskip}} \maketitle \begin{abstract} The belief propagation algorithm is a powerful tool in a wide range of disciplines from statistical physics to machine learning to computational biology, and is ubiquitous in decoding classical error-correcting codes. The algorithm works by passing messages between nodes of the factor graph associated with the code and enables efficient decoding of the channel, in some cases even up to the Shannon capacity. Here we construct the first belief propagation algorithm which passes \emph{quantum messages} on the factor graph and is capable of decoding the classical-quantum channel with pure state outputs. This gives explicit decoding circuits whose number of gates is quadratic in the code length. We also show that this decoder can be modified to work with polar codes for the pure state channel and as part of a decoder for transmitting quantum information over the amplitude damping channel. These represent the first explicit capacity-achieving decoders for non-Pauli channels. \end{abstract} \section{Introduction} Graphical models are at the heart of the current revolution in machine learning and computational statistics. They provide simple representations of the correlations among large numbers of random variables and enable efficient algorithms for feature discovery and analysis. Among the most well-known of these algorithms is belief propagation (BP), whose origin can be traced to the Bethe-Peierls approximation in statistical physics~\cite{mezard_information_2009}. BP can be used to marginalize the joint distribution of several random variables, often efficiently. For instance, in the setting of reliable communication over noisy channels via error correction, BP is used to find the most likely input for a given set of observed outputs. Indeed, in modern coding theory BP is simply indispensible~\cite{richardson_modern_2008}. The joint distribution of channel inputs and outputs can be represented by a factor graph, and BP works by passing messages between the nodes of this graph (an instance of more general message-passing algorithms). This leads to efficient decoding algorithms for high rate codes, several of which are employed in current wireless communication standards. Moreover, it was recently shown that belief propagation decoding of a certain class of low-density parity-check (LDPC) codes can achieve the Shannon capacity~\cite{kudekar_spatially_2013}. Factor graphs have been adapted to the quantum-mechanical setting from several different perspectives~\cite{tucci_quantum_1999,leifer_quantum_2008,loeliger_factor-graph_2012,loeliger_factor_2015}. Applied to quantum communication, BP and other message passing methods have been constructed for syndrome decoding of a variety of stabilizer codes subjected to Pauli noise channels~\cite{ollivier_description_2003-1,mackay_sparse-graph_2004,poulin_optimal_2006,poulin_iterative_2008,leifer_quantum_2008,poulin_quantum_2009,duclos-cianci_fast_2010,ferris_tensor_2014}. Despite their use in decoding quantum codes, these message passing algorithms are classical. Indeed, decoding any stabilizer code used for a Pauli channel or the erasure channel is essentially a classical task due to the Gottesman-Knill theorem~\cite{gottesman_heisenberg_1998}. However, stabilizer decoding is not optimal for non-Pauli channels such as the amplitude damping channel, for either the entanglement fidelity achievable by fixed-size codes or the largest achievable rates for codes with increasing blocklength. Therefore it would be of interest to extend BP decoding to more general channels. As much also holds in the setting of quantum polar codes, where the classical decoding method (ultimately a variant of BP) can only be employed without loss of rate for Pauli channels or the erasure channel~\cite{renes_efficient_2012,wilde_towards_2013,renes_polar_2014}. Note that the quantum decoding problem is different than the one solved by the classical algorithm for ``quantum belief propagation'' in~\cite{leifer_quantum_2008}.\footnote{The algorithm of \cite{hastings_quantum_2007} is also a classical algorithm.} There, one is interested in computing marginals of quantum states which have a structure given by a factor graph. For classical decoding, computing such marginals is indeed sufficient, as we will describe in more detail below. But even for bitwise decoding of a classical-quantum (CQ) channel having classical input and quantum output, it is not enough to know the relevant marginal state; we need a way to perform the optimal (Helstrom) measurement~\cite{helstrom_quantum_1976} or some suitable approximation. Put differently, a quantum BP decoder is a quantum algorithm, and we may expect that it will need to pass quantum messages. In this paper we construct a quantum BP decoding algorithm for the pure state channel, a binary input CQ channel whose outputs are pure states. The algorithm for estimating a single input bit works by passing single qubits as well as classical information along the factor graph, while sequential estimation of all input bits requires passing many qubits. For codes whose factor graphs are trees, as well as for polar codes, we show how the BP decoder leads to explicit circuits for the optimal measurement that have quadratic size in the code length. To the best our knowledge, this is the first instance of a quantum algorithm for belief propagation. The pure state channel arises, for instance, in binary phase-shift keying (BPSK) modulation of a pure loss Bosonic quantum channel, whose channel outputs are coherent states~\cite{guha_polar_2012}. Thus, our result gives an explicit construction of a successive cancellation decoder for the capacity-achieving polar code described in \cite{guha_polar_2012}, and addresses the issue of decoding CQ polar codes discussed in \cite{wilde_towards_2013}. Moreover, the pure state channel also arises as part of the quantum polar decoder for the amplitude damping channel~\cite{renes_efficient_2012,renes_polar_2014}, and therefore our result gives an explicit decoder for polar codes over this channel. The remainder of the paper is structured as follows. In the next section give a very brief overview of factor graphs and their use in classical decoding, and then rewrite the BP rules in a manner that lead to the quantum algorithm. Section~\ref{sec:qBP} gives the quantum BP decoding algorithm and applications to polar codes are given in Section~\ref{sec:polar}. We finish with several open questions for future research raised by our result. \section{Belief propagation decoding on factor graphs} \label{sec:classicalBP} Let us first examine BP on factor graphs directly in the coding context; for a more general treatment see~\cite{mackay_information_2002,richardson_modern_2008}. Consider the problem of reliable communication over a memoryless channel $W$ using a linear code $C$. Fix $C$ to be an $n$-bit code, i.e.\ a linear subspace of $\mathbb Z_2^n$, and suppose that the channel $W$ maps inputs in $\mathcal X=\mathbb Z_2$ to some alphabet $\mathcal Y$ according to the transition probabilities $P_{Y|X=x}=W(y|x)$. Now suppose a codeword $x_1^n=(x_1,x_2,\dots,x_n)\in C$ is picked at random and its consituent bits are each subjected to $W$, producing the output $y_1^n$. The goal of decoding is to invert this process and determine the input codeword from the channel output. This is a task of statistical inference, whose nominal solution is to output the $x^n_1$ which maximizes the conditional probability of inputs given outputs, $P_{X^n|Y^n}$. Since we assume the inputs are uniformly chosen from $C$, we can directly work with the joint distribution $P_{X^nY^n}$ of inputs and outputs. In general, though, this task is known to be computationally intractable. A simpler approach is to decode bitwise and find the most likely value of $x_k$ given $y_1^n$, for each $k$. Then we are interested in the marginal distribution $P_{X_kY^n}$, and we need only determine which of the two values of $x_k$ maximize $P_{X_kY^n}(x_k,y_1^n)$. Exact marginalization is also generally computationally intractable since the size of the joint distribution grows exponentially in the number of variables. \begin{comment} Since we assume the inputs are uniformly chosen from $C$, we can directly work with the joint distribution $P_{X^nY^n}$ of inputs and outputs, which is given by \begin{align} \label{eq:jointprob} P_{X^nY^n}(x_1^n,y_1^n)=\frac1{|C|}\mathbbm 1[{x^n_1\in C}]\prod_{j=1}^n W(y_j|x_j), \end{align} where $|C|$ is the number of encoded bits, i.e.\ the dimension of $C$ and $\mathbbm 1$ is the indicator function. The optimal decoder output is then $\text{argmax}_{x_1^n} P_{X^nY^n}(x_1^n,y_1^n)$. In general, this task is known to be computationally intractable. Alternately, we may decode bitwise and find the most likely value of $x_k$ given $y_1^n$, for each $k$. Then we are interested in the marginal distribution $P_{X_kY^n}$, and we need only determine which of the two values of $x_k$ maximize $P_{X_kY^n}(x_k,y_1^n)$. Nevertheless, exact marginalization is also generally computationally intractable since the size of the joint distribution grows exponentially in the number of variables. Then we are interested in the effective channel from any given bit, say $x_1$ to the entire output $y_1^n$, which is defined by the marginal distribution $P_{X_1Y^n}$. To find the most likely input $x_1$ given the observed $y_1^n$ we simply need to compute $P_{X_1Y^n}(x_1,y_1^n)$ and determine which value of $x_1$ maximizes this value. Exact marginalization is however generally intractable since the size of the joint distribution grows exponentially in the number of variables. In general, however, these tasks are both computationally intractable. Consider an arbitrary linear $n$-bit code $C$, i.e.\ a linear subspace of $\mathbb Z_2^n$, and suppose that a codeword is chosen uniformly at random and the constituent bits are each subjected a classical channel $W$ mapping inputs in $\mathcal X=\mathbb Z_2$ to some alphabet $\mathcal Y$ according to the transition probabilities $P_{Y|X=x}=W(y|x)$. , and consider the joint probability distribution $P_{X^nY^n}$ of input $X^n$ and output $Y^n$, supposing that the codewords are chosen uniformly at random. It is simply \begin{align} \label{eq:jointprob} P_{X^nY^n}(x_1^n,y_1^n)=\frac1{|C|}\mathbbm 1[{x^n_1\in C}]\prod_{j=1}^n W(y_j|x_j), \end{align} where $|C|$ is the number of encoded bits, i.e.\ the dimension of $C$, $\mathbbm 1$ is the indicator function, and $x_1^n$ denotes $x_1,x_2,\dots,x_n$. The effective channel from any given bit, say $x_1$ to the entire output $y_1^n$ is defined by the marginal distribution $P_{X_1Y^n}$. To find the most likely input $x_1$ given the observed $y_1^n$ we simply need to compute $P_{X_1Y^n}(x_1,y_1^n)$ and determine which value of $x_1$ maximizes this value. Exact marginalization is however generally intractable since the size of the joint distribution grows exponentially in the number of variables. \end{comment} However, for linear codes the joint distribution can be factorized, which often greatly simplifies the marginalization task. The joint distribution $P_{X^nY^n}$ can be written \begin{align} \label{eq:jointprob} P_{X^nY^n}(x_1^n,y_1^n)=\frac1{|C|}\mathbbm 1[{x^n_1\in C}]\prod_{j=1}^n W(y_j|x_j). \end{align} Since the channel is memoryless, the channel contribution to \eqref{eq:jointprob} is already in factorized form. Meanwhile, code membership is enforced by a sequence of parity-check constraints associated with the code, which also leads to factorization. In the three-bit repetition code, for instance, there are two parity constraints, $x_1+x_2=0$ and $x_2+x_3=0$ (or $x_1+x_3=0$), and therefore $\mathbbm 1[{x_1^3\in C}]=\mathbbm 1[x_1+x_2=0]\,\mathbbm 1[x_2+x_3=0]$. We can represent the joint distribution of any linear code (up to normalization) by a factor graph; Figure~\ref{fig:fgrepcode} shows the factor graph of a code involving two parity checks on four bits. For an arbitrary factorizeable function, the factor graph contains one (round) variable node for each variable and one (square) factor node for each factor, and factor nodes are connected to all their constituent variable nodes. This convention is violated in the figure by not including $y_j$ variable nodes; instead they are treated as part of the channel factors since their values are fixed and in any case each is connected to only one factor node. \begin{figure} \centering \includegraphics{factor-final-figure0.pdf} \caption{\label{fig:fgrepcode} Factor graph for the joint probability distribution of a four-bit code with two parity checks $x_1+x_3=0$ and $x_1+x_2+x_4=0$.} \end{figure} For factor graphs which are trees, meaning only one path connects any two nodes as in Figure~\ref{fig:fgrepcode}, the belief propagation algorithm can compute the marginal distributions exactly. In the present context of coding, it directly finds the most likely input value. Supposing we are interested in determining $x_1$, treat variable node $x_1$ as the root of the tree. BP then proceeds by passing messages between nodes, starting from the leaves (here, channel outputs) and working inward, combining all relevant information as it goes. Simplifying the general BP rules (see \cite{richardson_modern_2008}) to the decoding problem, the initial messages from the channel factors to the variable nodes can be taken as the log-likelihood ratios $\ell=\log [W(y_j|0)/W(y_j|1)]$ of the channel given the output $y_j$ (here we suppress the dependence of $\ell$ on the channel output $y_j$). At variable nodes the messages simply add, so that the outgoing $\ell$ is the sum of incoming $\ell_k$. At check nodes the rule is more complicated: $\tanh\frac\ell 2 =\prod_k \tanh \frac{\ell_k}2$. After all messages have arrived at the root, the algorithm produces the log-likelihood ratio for $x_1$ given all the channel outputs, and the decoder simply outputs 0 if the ratio is positive or 1 if negative. By adopting a modified update rule it is in fact possible to compute all the marginals at once with only a modest overhead. Instead of only proceeding inward from the leaves, we send messages in both directions along each edge, starting by sending channel log-likelihoods in from the leaves. Each node sends messages on each edge once it has received messages on all its other edges. For graphs that contain loops, the algorithm is not guaranteed to converge, but one can nevertheless hope that the result is a good approximation and that the decoder outputs the correct value. This is borne out in practice for turbo codes and LDPC codes. There is an intuitive way of understanding the BP decoding algorithm which is the basis of our quantum generalization. At every step the message can be interpreted as the log-likelihood ratio of the effective channel from that node to its descendants. This is sensible as the likelihood ratio is a sufficient statistic for estimating the (binary) input from the channel output. The rules for combining messages can then be interpreted as rules for combining channels, and the algorithm can be seen as successively simplifying the channel from the root to the leaves by utilizing the structure of the factor graph. At variable nodes, adding the log-likelihood ratios for two channels $W$ and $W'$ amounts to considering the convolution channel $W\oasterisk W'$ with transition probabilities given by \begin{align} \label{eq:variablenode} [W\oasterisk W'](y,y'|x)=W(y|x)W'(y'|x). \end{align} That is, the effective channel associated with a variable node is simply the convolution $W_1\oasterisk\cdots \oasterisk W_k$ of its descendants. The form of the effective channel at check nodes is not as immediate, but it is not too difficult to verify that the appropriate channel convolution $W\boxasterisk W'$ has transition probabilities \begin{equation} \begin{aligned} \label{eq:checknode} [&W\boxasterisk W'](y,y'|x)=\tfrac12\left(W(y|x)W'(y'|0)+W(y|x+1)W'(y'|1)\right). \end{aligned} \end{equation} These two channel convolutions are also the fundamental building blocks of polar codes~\cite{arikan_channel_2009}, at least when the input channels are symmetric. The check node convolution is the ``worse'' channel in the channel splitting or channel synthesis step (cf.\ \cite[Eq.\ 19]{arikan_channel_2009}); this holds regardless of the symmetry of the channel. On the other hand, the ``better'' combination of $W$ and $W'$ is defined by (cf.\ \cite[Eq.\ 20]{arikan_channel_2009}) $W''(y,y',x|x')=\tfrac12W(y|x+x')W'(y'|x')$. Compared to \eqref{eq:variablenode}, the input $x$ is uniformly random and not always zero, but it is given at the channel output. When $W$ is symmetric in the sense that $W(y|x+u)=W(\pi_u(y)|x)$ for a suitable permutation $\pi$ of the output alphabet depending on $u$, we can reversibly transform $W''$ into $W\oasterisk W'$ and vice versa. \section{Belief propagation decoding of quantum outputs} \label{sec:qBP The form of the check and variable convolutions also applies to channels with quantum output.\footnote{This was first applied in the setting of polar codes in~\cite{wilde_polar_2013}.} We need only replace the probability distributions over the output alphabet by quantum states. Abusing notation, let us denote by $W(x)$ the quantum state of the output of $W$ given input $x$. This includes the previous case by considering commuting $W(x)$. The the variable and check node convolutions are now just \begin{align} [W\oasterisk W'](x)&=W(x)\otimes W'(x),\qquad \text{and}\\ [W\boxasterisk W'](x) &= \tfrac12(W(x)\otimes W'(0)+W(x{+}1)\otimes W'(1)). \end{align} To properly generalize the BP decoding algorithm we need a ``sufficient statistic'' for the quantum channels at the various nodes. For binary-input pure state channels, it turns out that a combination of classical bits and just one qubit suffices. The channel outputs can always be represented by a qubit, so suppose that $W$ outputs $\ket{\pm \theta}$, where $\ket{\theta}=\cos\frac\theta 2\ket{0}+\sin\frac\theta 2\ket 1$. Note that the overlap of the two states is $\cos\theta$ and the Helstrom measurement for these two states is measurement of the $\sigma_x$ operator. The convolution $W\oasterisk W'$ outputs either $\ket{ \theta}\otimes \ket{ \theta'}$ or $\ket{- \theta}\otimes \ket{- \theta'}$, which are again two pure states, with an overlap angle $\theta^\oasterisk$ given by $\cos\theta^\oasterisk=\cos\theta\cos\theta'$. The following unitary transformation compresses the states to the first qubit, leaving the second in the state $\ket{0}$: \begin{align} U_\oasterisk(\theta,\theta')= \begin{pmatrix} a_+ & 0 & 0 & a_-\\ a_- & 0 & 0 & -a_+\\ 0 & b_+ & b_- & 0\\ 0 & b_- & -b_+ & 0 \end{pmatrix}, \end{align} with $a_\pm\sqrt{1+\cos\theta\cos\theta'}=\tfrac1{\sqrt2}(\cos(\frac{\theta-\theta'}2)\pm\cos(\frac{\theta+\theta'}2))$ and $b_\pm\sqrt{1-\cos\theta\cos\theta'}=\tfrac1{\sqrt2}(\sin(\frac{\theta+\theta'}2)\mp\sin(\frac{\theta-\theta'}2))$. To combine more than two channels, we just perform the pairwise convolution sequentially. Thus, the $\oasterisk$ convolution of pure state channels can itself be represented as a pure state channel. \begin{comment} The form of the check and variable convolutions also applies to channels with quantum output.\footnote{This was first applied in the setting of polar codes in~\cite{wilde_polar_2013}.} We need only replace the probability distributions over the output alphabet by quantum states. Abusing notation, let us denote by $W(x)$ the quantum state of the output of $W$ given input $x$. This includes the previous case by considering commuting $W(x)$. The the variable and check node convolutions are now just \begin{align} [W\oasterisk W'](x)&=W(x)\otimes W'(x),\qquad \text{and}\\ [W\boxasterisk W'](x) &= \tfrac12(W(x)\otimes W'(0)+W(x{+}1)\otimes W'(1)). \end{align} To properly generalize the BP decoding algorithm we need a ``sufficient statistic'' for the quantum channels at the various nodes. For pure state channels we can get away with a combination of classical bits and just one qubit. We may as well regard the two pure outputs of the bare channels themselves as states of a qubit. For concreteness, suppose that $W$ outputs $\ket{\pm \theta}$ and $W$ $\ket{\pm \theta'}$, where $\ket{\theta}=\cos\frac\theta 2\ket{0}+\sin\frac\theta 2\ket 1$. Note that the overlap of the two states is given by $\cos\theta$ and the Helstrom measurement is just projection onto $\ket{\pm \frac\pi 2}$. The convolution $W\oasterisk W'$ outputs either $\ket{ \theta}\otimes \ket{ \theta'}$ or $\ket{- \theta}\otimes \ket{- \theta'}$, which are again two pure states with an overlap angle $\theta^\oasterisk$ given by $\cos\theta^\oasterisk=\cos\theta\cos\theta'$. All the information about the input can be exactly stored in just one qubit, and therefore the algorithm for combining two channels at a variable node is to unitarily compress the output into one qubit and pass the qubit to the parent node. This is accomplished by the unitary transformation \begin{align} U_\oasterisk(\theta,\theta')=\begin{pmatrix} \frac{\sqrt2 \cos\frac{\theta}2\cos\frac{\theta'}2}{\sqrt{1+\cos\theta\cos\theta'}} & 0 & 0 & \frac{\sqrt2 \sin\frac{\theta}2\sin\frac{\theta'}2}{\sqrt{1+\cos\theta\cos\theta'}}\\ \frac{\sqrt2 \sin\frac{\theta}2\sin\frac{\theta'}2}{\sqrt{1+\cos\theta\cos\theta'}} & 0 & 0 & -\frac{\sqrt2 \cos\frac{\theta}2\cos\frac{\theta'}2}{\sqrt{1+\cos\theta\cos\theta'}}\\ 0 & \frac{\sqrt2 \cos\frac{\theta}2\sin\frac{\theta'}2}{\sqrt{1-\cos\theta\cos\theta'}} & \frac{\sqrt2 \sin\frac{\theta}2\cos\frac{\theta'}2}{\sqrt{1-\cos\theta\cos\theta'}} & 0\\ 0 & \frac{\sqrt2 \sin\frac{\theta}2\cos\frac{\theta'}2}{\sqrt{1-\cos\theta\cos\theta'}} & -\frac{\sqrt2 \cos\frac{\theta}2\sin\frac{\theta'}2}{\sqrt{1-\cos\theta\cos\theta'}} & 0 \end{pmatrix}, \end{align} which leaves the second qubit in the state $\ket{0}$. To combine more than two channels, we just perform the pairwise convolution sequentially. \end{comment} The $\boxasterisk$ convolution is more complicated because the outputs are no longer pure. However, applying the unitary $U_\boxasterisk=\textsc{cnot}_{1\to 2}$ results in a CQ state of the form $\sum_{j\in\{0,1\}} p_j\ketbra{\pm \theta^\boxasterisk_j}\otimes \ketbra j$. We are free to measure the second qubit, and conditional state of the first qubit is again one of two pure states, though now the overlap $\cos\theta^\boxasterisk_j$ depends on the measurement outcome $j$. In particular, $p_0=\tfrac12(1+\cos\theta\cos\theta')$, $p_1=1-p_0$, and the two overlaps are given by \begin{align} \label{eq:box1} \cos\theta^\boxasterisk_0 &=\frac{\cos\theta+\cos\theta'}{1+\cos\theta\cos\theta'},\\ \cos\theta^\boxasterisk_1 &=\frac{\cos\theta-\cos\theta'}{1-\cos\theta\cos\theta'}. \label{eq:box2} \end{align} For outcome $j=0$ the angle between the states has decreased, while for outcome $j=1$ the angle has increased. Therefore, the $\boxasterisk$ convolution of pure state channels can be represented by two pure state channels, corresponding to the two measurement outcomes. As before, several channels can be combined sequentially. The quantum decoding algorithm now proceeds as in classical BP, taking the quantum outputs of the channels and combining them at variable and check nodes. At a variable node the algorithm combines the outputs using $U_\oasterisk$ and forwards the output to its parent node. At check nodes the algorithm applies $U_\boxasterisk$, measures the second qubit, and forwards both the qubit and the measurement result to its parent node. The classical messages are required to inform parent variable nodes how to choose the angles in subsequent $U_\oasterisk$ unitaries. Ultimately this procedure results in one qubit at the root node such that measurement of $\sigma_x$ corresponds to the optimal Helstrom measurement for the associated bit. This then is sufficient to estimate one input bit. For example, return to the code depicted in Figure~\ref{fig:fgrepcode} for a pure state channel with overlap $\theta$, and suppose we are interested in decoding the first bit. Starting at the leaves, the outputs of all but the first channel can be immediately passed to their corresponding variable nodes, since these variable nodes do not have any other outward branches. (Formally this follows from the convolution rules by considering convolution with a trivial channel, having $\theta=0$.) The output of the first channel, meanwhile, must wait to be combined according to the $\oasterisk$ convolution with several other qubit messages. Next, since 2 and 4 are connected by a check node, we combine qubits 2 and 4 into one qubit (2) and one classical bit (4) by applying $U_\boxasterisk$ and measuring the 4th qubit. As qubits 1 and 3 are connected by a variable node, we can simultaneously combine these with $U_\oasterisk(\theta,\theta)$. Finally, we combine qubits 1 and 2 by applying $U_\oasterisk(\theta^\oasterisk,\theta^\boxasterisk_j)$, where $\cos\theta^\oasterisk=\cos^2\theta$ and $\cos\theta_{0}^\boxasterisk=\frac{2\cos\theta}{(1+\cos^2\theta)}$, $\theta_1^\boxasterisk=\frac\pi 2$, depending on the value $j$ of the earlier measurement. A quantum circuit implementing these steps is shown in Figure~\ref{fig:circuit}. \begin{figure} \centering \[\Qcircuit @C=1.5em @R=.7em { & \qw & \multigate{1}{U_\oasterisk} & \qswap & \qw{|}\\ & \qswap & \ghost{U_\oasterisk} & \qswap \qwx & \multigate{1}{U_\oasterisk} & \gate{H}&\meter & \cw\\ & \qswap \qwx & \multigate{1}{U_\boxasterisk} & \qw & \ghost{U_\oasterisk} & \qw{|}\\ & \qw & \ghost{U_\boxasterisk} & \qw & \meter \cwx & } \] \caption{\label{fig:circuit} Circuit decoding the first bit of the code depicted in Figure~\ref{fig:fgrepcode}. The first $\oasterisk$ convolution is $U_\oasterisk(\theta,\theta)$, the second $U_\oasterisk(\theta^\oasterisk,\theta^\boxasterisk_j)$ for $\cos\theta^\oasterisk=\cos^2\theta$ and $\cos\theta_{0}^\boxasterisk=\frac{2\cos\theta}{(1+\cos^2\theta)}$, $\theta_1^\boxasterisk=\frac\pi 2$, depending on the value $j$ of the measurement outcome in the bottom wire. The symbol $\dashv$ denotes that the qubit is discarded. The final Hadamard gate and measurement implement the Helstrom measurement. } \end{figure} One drawback is that the above procedure implements the Helstrom measurement destructively, since once we estimate the first bit we no longer have the original channel output in order to estimate the second bit. And we cannot run the algorithm backwards to reproduce the channel output as we have made measurements at every check node. To implement the Helstrom measurement as nondestructively as possible, we can leave the CQ output states unmeasured and instead use the classical subsystems to coherently control the variable node unitaries $U_\oasterisk$. In this way the steps in the algorithm can be reversed, save the final measurement. For example, in Figure~\ref{fig:circuit} all output qubits are kept and the classical measurement and subsequent conditioning of the second $U_\oasterisk$ gate is performed by a coherent conditional gate involving three qubits. Denoting the unitary action of the algorithm for the $j$th bit by $V_j$, the Helstrom measurement can be implemented by the projective measurement with projectors $\Pi_{j,k}=V_j^* |\tilde k\rangle\langle \tilde k|_j V_j$, where $|\tilde k\rangle\langle \tilde k|_j$ denotes the $k$th $\sigma_x$ basis projector on the $j$th qubit. Each $V_j$ is composed of $O(n)$ gates, yielding an overall circuit size of $O(n^2)$ to decode all bits. Supposing that the code is designed such that the $j$th input bit can be estimated with error no larger than $\epsilon_j$, Gao's non-commutative union bound~\cite{gao_quantum_2015} implies that the error in sequentially estimating all bits is no worse than $4\sum_j \epsilon_j$. \section{Applications to polar codes} \subsection{Polar codes for the pure state channel} \label{sec:polar} Polar codes for the pure state channel may also be decoded with this algorithm. Indeed, the successive cancellation decoding algorithm proposed by Ar\i{}kan in~\cite{arikan_channel_2009} proceeds precisely by combining channels using the $\oasterisk$ and $\boxasterisk$ rules, and was adapted to the case of classical-quantum channels in~\cite{wilde_polar_2013}. The difference is that successive cancellation does not use the factor graph of the code, but a graph related to a fixed reversible encoding circuit. Importantly, the graph associated to each input of the encoding circuit is a tree. In fact, each such graph has logarithmic depth from all channel factors to each variable, and every node has degree three. Unlike the BP decoder, however, the successive cancellation decoder used by polar codes takes previously decoded bits into account. But these bits can be handled by the BP decoder since the pure state channel is symmetric in the manner described at the end of \S\ref{sec:classicalBP}. There, the value of the previous bits is incorporated into the better channel by appropriately permuting the output symbols, which is equivalent to flipping the input value. Similarly, for the pure state channel, applying $\sigma_z$ to the output is equivalent to flipping the input. Therefore, the quantum BP decoding algorithm gives a successive cancellation decoder for polar codes over the pure loss Bosonic channel using the BPSK constellation~\cite{guha_polar_2012}. \subsection{Quantum polar codes for amplitude damping} The idea behind the quantum polar coding scheme of \cite{renes_efficient_2012,renes_polar_2014} is to decompose the problem of transmitting quantum information over a channel $\mathcal N_{A\to B}$ into transmitting classical information about two conjugate observables, ``amplitude'' and ``phase'', consider polar codes for each subproblem, and then combine the coding schemes using CSS codes at the encoder and coherent sequential decoding of amplitude and phase at the decoder. This decoding strategy is depicted in \cite[Fig.\ 3]{renes_efficient_2012} for Pauli channels and \cite[Fig.\ 1]{renes_uncertainty_2016-1} for the general case. As detailed in \cite{renes_polar_2014}, the two classical transmission tasks are to transmit ``amplitude'' information over the CQ channel given by $z\to \rho_z=\mathcal{N}(\ketbra z)$ and ``phase'' information over the CQ channel given by $x\to \varphi_x=(Z^x\otimes \mathbbm 1) (\mathcal I\otimes \mathcal N)[\Phi](Z^x\otimes \mathbbm 1)$. Here $\ket{z}$ is an arbitrary basis, and we choose that of $\sigma_z$ for convenience, while $\ket{\Phi}_{A'A}=\sum_z\sqrt{p_z}\ket{z}\ket{z}$ is a bipartite pure state in this same basis with coefficients of our choosing. (See \cite{renes_polar_2014} for the precise relation to the conjugate observables $\sigma_x$ and $\sigma_z$.) Let us now show how to build a decoder for the amplitude damping channel $\mathcal N_\gamma$ with damping parameter $\gamma\in [0,1]$. First note that the amplitude outputs all commute due to the form of $\mathcal N_\gamma$; the amplitude channel is effectively a classical Z channel in which the input 0 is always transmitted perfectly, but the input 1 may decay to 0 with probability $\gamma$. Therefore we can use the classical polar encoder and decoder for this channel~\cite{honda_polar_2013}. Since the Z channel is not symmetric, the optimal input distribution in the capacity formula is not the uniform distribution, but one with probabilities $p$ and $1-p$. Now suppose that the bipartite pure state in the phase channel is the state $\ket{\Phi}=\sqrt{p}\ket{00}+\sqrt{1-p}\ket{11}$. Abusing notation slightly and denoting the channel outputs $\varphi_\pm$, it is not difficult to verify that for $U=\textsc{cnot}_{A'\to B}$, \begin{align} U\varphi_\pm U^*&=(1-\gamma(1-p))\ketbra{\pm\theta_0}\otimes \ketbra{0}+\gamma(1-p)\ketbra 1\otimes \ketbra{1},\qquad \text{with}\\ \cos\theta_0&=\frac{1-2p-\gamma(1-p)}{1-\gamma(1-p)}. \end{align} Each of these states is a CQ state with the first qubit pure and the second qubit classical, just as in a $\boxasterisk$ output. Given the second qubit, the first is either in the pure state $\ket{\pm \theta_0}$ corresponding to the channel input $\pm$, or the state $\ket{1}$ independently of the input; the latter is equivalent to $\ket{\theta_1=0}$. Hence the decoder can begin just as at a $\boxasterisk$ step, measuring the second qubit to determine the angle associated to the first qubit. The rate achievable by the quantum polar code construction is simply $R=\max_{p\in [0,1]} \left(1-H(Z|B)_\psi-H(X|BA')_\xi\right)$, where $\psi_{ZB}=p\ketbra 0\otimes \rho_0+(1-p)\ketbra 1\otimes \rho_1$ and $\xi_{XBA'}=\tfrac12\sum_{x\in \{0,1\}}\ketbra x\otimes \varphi_x$. A cumbersome but straightforward calculation confirms that $R$ equals the capacity of the channel, $C(\mathcal N_\gamma)=\max_{p\in[0,1]} \left(h_2((1-\gamma)p)-h_2(\gamma p)\right)$, for $h_2$ is the binary entropy~\cite[Prop.\ 24.7.2]{wilde_quantum_2017}. Moreover, since the amplitude damping channel is degradable, the arguments in \cite{renes_efficient_2012} ensure that no entanglement-assistance is required to meet the CSS constraint when constructing the quantum polar code. \section{Discussion} We have presented a belief propagation algorithm for bitwise decoding of CQ channels which operates by passing quantum messages on tree factor graphs, and shown several applications to polar codes. This invites the study of quantum message passing algorithms, and not just in the context of decoding. More generally we may look for BP and related algorithms for any task of statistical inference where the input data comes in the form of many quantum bits, for instance in quantum metrology. This work also raises many interesting questions. Most immediately in the context of decoding is whether the complexity of the algorithm can be reduced for structured factor graphs. Classical polar codes, for instance, have decoding complexity $O(n\log n)$. Can this also be achieved for the pure state channel? Similarly, can one find a quantum version of the max-product or Viterbi algorithm for determining the most likely $x_1^n$ given the channel outputs? More generally, it would be very interesting to understand how to run the algorithm on a factor graph with loops, or how it can be modified to handle some set of non-pure output states. In the former case it may be useful to explore the characterization of loopy BP as a variational problem~\cite{wainwright_graphical_2007,mezard_information_2009}. Perhaps in the latter case one can make use of the work on quantum sufficiency (see e.g.\ \cite{jencova_sufficiency_2006,buscemi_comparison_2012} and references therein) to find a suitable set of quantum messages for a given decoding problem. Another interesting question with potentially far-reaching consequences is the relation of the BP algorithm to tensor network methods. The problem of marginalization in the commutative setting is explicitly treated as tensor network contraction in~\cite{ferris_tensor_2014}, and the particulars of the quantum BP decoder bear a similarity with the data gathering approach using tensor network states in \cite{blume-kohout_quantum_2013}. Can the methods of approximating quantum states by tensor networks be used to create efficient approximate decoders? \\[-2mm] \section*{Acknowledgments} It is a pleasure to acknowledge helpful conversations with R\"udiger Urbanke, Marco Mondelli, and David Sutter. Thanks also to Narayanan Rengaswamy for pointing out an error in $U_\boxasterisk$ in a previous version of this paper. This work was supported by the Swiss National Science Foundation (SNSF) via the National Centre of Competence in Research ``QSIT'', and by the European Commission via the project ``RAQUEL''. \printbibliography[heading=bibintoc,title=References] \end{document}
a47b728159bd2fcc5ebef8effcf778f11c3a74ad
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{More details for section \ref{sec:2wst}} \label{app:2way} \begin{figure}[h] \tikzstyle{trans}=[-latex, rounded corners] \begin{center} \scalebox{0.7}{ \begin{tikzpicture}[->,>=stealth',shorten >=1pt,auto, semithick,scale=.9] \tikzstyle{every state}=[fill=gold] \node[state] at (0,4) (A) {$p$} ; \node[state] at (3,4) (B) {$q$} ; \node[state,initial,initial where=left,initial text={}, accepting] at (-3,4) (C) {$t$} ; \path(A) edge[loop above] node {$\alpha\left|\begin{array}{lll} \alpha, -1 \end{array}\right.$} (A); \path(A) edge node[above] {$\# \left| \epsilon, +1 \right.$} node[below] {$\vdash \left| \epsilon, +1 \right.$} (B); \path(B) edge[loop above] node {$\alpha\left|\begin{array}{llllllll} \alpha, +1 \end{array}\right.$} (B); \path(B) edge [bend left=40] node {$\# \left| \epsilon, +1 \right.$} (C); \path(C) edge[loop above] node {$\neg \texttt{reach}_{\#}, \alpha\left|\begin{array}{llllllll} \alpha, +1 \end{array}\right.$} (C); \path(C) edge[loop below] node {$\texttt{reach}_{\#}, \alpha\left|\begin{array}{llllllll} \epsilon, +1 \end{array}\right.$} (C); \path(C) edge node {$\# \left| \epsilon, -1 \right.$} (A); \end{tikzpicture} } \end{center} \vspace{-1em} \caption{Transformation $f_1$ given as two-way transducers with look-ahead. Here symbol $\alpha$ stands for both symbols $a$ and $b$, and the predicate $\texttt{reach}_{\#}$ is the lookahead that checks whether string contains a $\#$ in future. $\{t\}$ is the only Muller set. \label{fig:2wst}} \vspace{-1em} \end{figure} First we give some examples of transition monoids for the \ensuremath{\mathsf{2WST}}{} in Figure \ref{fig:2wst}. Consider the string $ab\#$. The transition monoid is obtained by using all 4 behaviours as shown below. Note that on reading $ab \#$, on state $t$, when we reach symbol $\#$, the look-ahead $\neg reach_{\#}$ evaluates to true. \[ M^{\ell r}_{ab\#}=\begin{blockarray}{cccc} &\matindex{t} & \matindex{q} & \matindex{p} \\ \begin{block}{c(ccc)} \matindex{t} & (0) & \bot & \bot\\ \matindex{q} & (0) &\bot & \bot \\ \matindex{p} & \bot & \bot & \bot \\ \end{block} \end{blockarray},~ M^{r \ell}_{ab\#}=\begin{blockarray}{cccc} &\matindex{t} & \matindex{q} & \matindex{p} \\ \begin{block}{c(ccc)} \matindex{t} & \bot & \bot & (0)\\ \matindex{q} & \bot &\bot & \bot \\ \matindex{p} & \bot & \bot & \bot \\ \end{block} \end{blockarray} \] \[ M^{\ell \ell}_{ab\#}=\begin{blockarray}{cccc} &\matindex{t} & \matindex{q} & \matindex{p} \\ \begin{block}{c(ccc)} \matindex{t} & \bot & \bot & (0)\\ \matindex{q} & \bot &\bot & \bot \\ \matindex{p} & \bot & \bot & (0) \\ \end{block} \end{blockarray},~ M^{r r}_{ab\#}=\begin{blockarray}{cccc} &\matindex{t} & \matindex{q} & \matindex{p} \\ \begin{block}{c(ccc)} \matindex{t} & (0) & \bot & \bot\\ \matindex{q} & (0) &\bot & \bot \\ \matindex{p} & \bot & (0) & \bot \\ \end{block} \end{blockarray} \] If we consider the string $ab\#ab\#$, then we can compute for instance, $M^{\ell r}_{ab\#ab\#}$ using $M^{\ell r}_{ab\#}, M^{\ell \ell}_{ab\#}$ and $M^{rr}_{ab\#}$. It can be checked that we obtain $M^{\ell r}_{ab\#ab\#}$ same as $M^{\ell r}_{ab \#}$. The transition monoid of the two-way automaton is obtained by using all the four matrices $M^{xy}_s$ for a string $s$. In particular, given a string $s$, we consider the matrix \[ M_s=\left ( {\begin{array}{cc} M^{\ell \ell}(s) & M^{\ell r}(s)\\ M^{r \ell}(s) & M^{ r r}(s)\\ \end{array}} \right ) \] as the transition matrix of $s$ in the two-way automaton. The identity element is \small{$\left ( {\begin{array}{cc} $\textbf{1}$ & $\textbf{1}$\\ $\textbf{1}$ & $\textbf{1}$\\ \end{array}} \right )$} where $\textbf{1}$ is the $n \times n$ matrix whose diagonal entries are $(\emptyset, \emptyset, \ldots, \emptyset)$ and non-diagonal entries are all $\bot$'s. The matrix corresponding to the empty string $\epsilon$ is $\left ( {\begin{array}{cc} $\textbf{1}$ & $\textbf{1}$\\ $\textbf{1}$ & $\textbf{1}$\\ \end{array}} \right )$. Given a word $w \in \Sigma^*$, we can find a decomposition of $w$ into $w_1$ and $w_2$ such that we can write all behaviours of $w$ in terms of behaviours of $w_1$ and $w_2$, denoted by $M_{w_1 w_2}$. We enumerate the possibilities in each kind of traversal, for a successful decomposition. \begin{enumerate} \item For $w=w_1w_2$, and a left-left traversal, we can have $M^{\ell \ell}(w)=M^{\ell \ell}(w_1)$ or \\ $M^{\ell \ell}(w)=M^{\ell r}(w_1) \times (M^{\ell \ell}(w_2) \times M^{r r}(w_1))^* \times M^{\ell \ell}(w_2) \times M^{r \ell}(w_1)$. We denote these cases as $LL_1, LL_2$ respectively. \item For $w=w_1w_2$ and a left-right traversal, we have \\ $M^{\ell r}(w)=M^{\ell r}(w_1) \times (M^{\ell \ell}(w_2) \times M^{r r}(w_1))^* \times M^{\ell r}(w_2)$. We denote this case as $LR$. \item For $w=w_1w_2$ and a right-left traversal, we have \\ $M^{r \ell}(w)= M^{r \ell}(w_2) \times (M^{r r}(w_1) \times M^{\ell \ell}(w_2))^* \times M^{r \ell}(w_1)$. We denote this case as $RL$. \item For $w=w_1w_2$ and a right-right traversal, we have\\ $M^{r r}(w)=M^{r r}(w_2)$ or \\ $M^{r r}(w)=M^{r \ell}(w_2) \times (M^{r r}(w_1) \times M^{\ell \ell}(w_2))^* \times M^{r r}(w_1) \times M^{\ell r}(w_2)$. We denote these cases as $RR_1, RR_2$ respectively. \end{enumerate} With these, for a correct decomposition of $w$ as $w_1w_2$, $M_{w_1w_2}$ is one of the four matrices as given below for $i, j \in \{1,2\}$. \[M_{w_1w_2}=\left ( {\begin{array}{cc} LL_i & LR\\ RL & RR_j\end{array}} \right ) \] The multiplication of matrices $M^{xy}(w)$ using the $\times$ operator is as defined for DMA using the multiplication $\odot$ and addition $\oplus$ of elements in $(\{0,1\} \cup 2^Q)^n \cup \bot$. We define a new operation $\otimes$, which takes $M_{w_1}, M_{w_2}$, and for a ``correct'' decomposition of $w$ using $w_1,w_2$, denoted $w_1 \otimes w_2$, we obtain $M_{w_1} \otimes M_{w_2}=M_{w_1 \otimes w_2}$. It can be seen that with the $\otimes$ operation, the transition matrix of a string decomposed as $w_1.w_2$ correctly follows from the left-right behaviours of the strings $w_1, w_2$. Let $\mathcal{T}(\mathcal{A})$ be the transition monoid of the two-way automaton $\mathcal{A}$. Let $|Q|=n$, the number of states of $\mathcal{A}$, and let there be $m$ muller sets. For ease of notation, we do not include the states of the look-behind automaton and set of states of the look-ahead automaton in the transition monoid. Thus, $\mathcal{T}(\mathcal{A})$ contains matrices of the above form, with identity and the binary operation as defined above. It can be seen that recursively we can find decompositions of $w$ as $w_1$ followed by $w_2w_3$, and $w_1w_2$ followed by $w_3$ such that $M_{w}=M_{w_1} \otimes M_{w_2w_3}= M_{w_1w_2} \otimes M_{w_3}$. \begin{lemma} \label{app:lem1-conv} Let $\mathcal{A}$ be a two-way (aperiodic) Muller automaton. The mapping $h$ which maps any string $s$ to its transition monoid $\mathcal{T}(\mathcal{A})$ is a morphism that recognizes $L(\mathcal{A})$. The language $L(\mathcal{A})$ is then aperiodic. \end{lemma} \begin{proof} The proof of Lemma \ref{app:lem1-conv} is similar to Lemma \ref{lem:ap-rec}. As seen in the case of Lemma \ref{lem:ap-rec}, it can then be proved that $\mathcal{T}(\mathcal{A})$ is a monoid. To define a morphism from $\Sigma^*$ to $\mathcal{T}(\mathcal{A})$, we first define a mapping $h: (\Sigma^*, \cdot, \epsilon) \rightarrow (\Sigma^* \times \Sigma^*, R, (\epsilon, \epsilon))$ as $h(s) = (s_1, s_2)$ iff $s = s_1 \otimes s_2$ is a ``correct breakup''; that is, $M_s = M_{s_1} \otimes M_{s_2}$. The operator $R$ is defined as $(a_1, a_2) R (b_1, b_2) = (a_1 a_2 ,b_1 b_2)$, and $h(\epsilon) = (\epsilon, \epsilon).$ It is easy to see that $h$ is a morphism, and $(\Sigma^* \times \Sigma^*, R, (\epsilon, \epsilon))$ is a monoid. Next, we define another map $g$ which works on the range of $h$. $g: (\Sigma^* \times \Sigma^*, R, (\epsilon, \epsilon)) \rightarrow (T_\mathcal{A}, \otimes, \textbf{1})$ as $g((s_1, s_2))=M_{s_1 s_2}$, $g[(s_1,s_2) R (s_3,s_4)] = M_{s_1s_2} \otimes M_{s_3s_4}$, which by the correct breakup condition, will equal $M_{s_1s_2s_3s_4}$. Also, $g((\epsilon,\epsilon))=\textbf{1}$. Again, $(T_\mathcal{A}, \otimes, \textbf{1})$ is a monoid and the composition of $g$ and $h$ is a morphism from $(\Sigma^*, \cdot, \epsilon)$ to $\mathcal{T}(\mathcal{A})$. Further, one can show using a similar argument to Lemma \ref{lem:ap-rec} that $L(\mathcal{A})$ is recognized by the morphism $g\circ h$. It remains to show that $L(\mathcal{A})$ is aperiodic. By definition, we know that the languages $L^{xy}_{pq}$ are aperiodic. This means that there is some $m$ such that $w^m \in L^{xy}_{pq}$ iff $w^{m+1} \in L^{xy}_{pq}$ for all strings $w$. This implies that $M^{xy}(w^m)=M^{xy}(w^{m+1})$ for all strings $w$. This would in turn imply (since we have the result for all four quadrants) that we obtain for the matrices of $\mathcal{T}(\mathcal{A})$, $M^m_{w}=M_{w^m}=M_{w^{m+1}}=M^{m+1}_w$, where $M^m$ stands for the iterated product using $\otimes$ of matrix $M$ $m$ times. This implies that there is an $m$ such that for all matrices $M$ in $\mathcal{T}(\mathcal{A})$, $M^m=M^{m+1}$. Hence, $\mathcal{T}(\mathcal{A})$ is aperiodic. Since $h$ is a morphism to an aperiodic monoid recognizing $L(\mathcal{A})$, we obtain that $L(\mathcal{A})$ is aperiodic. \end{proof} \section{Proofs from Section \ref{sec:2wst-sst} : $\ensuremath{\mathsf{2WST}}_{\ensuremath{sf}} \subset \mathsf{SST}_{\ensuremath{sf}}$} \label{app-2wst-sst} In this section, we show that $\ensuremath{\mathsf{2WST}}_{\ensuremath{sf}} \subset \mathsf{SST}_{\ensuremath{sf}}$. Let the \ensuremath{\mathsf{2WST}}$_{\ensuremath{sf}}$ be $(T, A, B)$ with $T = (\Sigma, \Gamma, Q, q_0, \delta, F)$. We assume wlg that $L(A_p) \cap L(A_{p'})=\emptyset$ for all $p, p' \in Q_A$ and $L(A_r) \cap L(A_{r'})=\emptyset$ for all $r, r' \in Q_B$. Also, we assume that $F=2^Q$, and check particular accepting conditions in the look-ahead automaton $A$. The \sst$_{sf}$ we construct is $\mathcal{T}=((\Sigma, \Gamma, Q', q_0', \delta', X, \rho, F'), A, B)$ where $Q'=Q \times [Q \rightharpoonup Q]$ where $\rightharpoonup$ represents a partial function. $q'_0=(q_0, I)$ where $I$ is the identity. The set of variables $\mathcal{X}=\{X_q \mid q \in Q\} \cup \{O\}$, and $F': 2^{Q'} \rightarrow X^*$ is defined for all $P' \in F'$ as $F'(P')=O$. $\delta': Q' \times Q_B \times \Sigma \times Q_A \rightarrow Q'$ is defined as follows: $\delta'((q,f),r,a,p)=(f'(q),f')$ where $f'$ is defined as follows: \[ f'(q)=\left \{ \begin{array}{lll} & t & \mbox{if}~ \delta(q,r,a,p)=(t, \gamma, 1),\\ & f'(t) & \mbox{if}~ \delta(q,r,a,p)=(t, \gamma, 0),\\ & f'(f(t)) & \mbox{if}~ \delta(q,r,a,p)=(t, \gamma, -1) \end{array}\right. \] Thanks to the determinism of the \ensuremath{\mathsf{2WST}}$_{sf}$ and the restriction that the entire input be read, $f'$ is a well-defined, partial function. In each cell $i$, the state of the \sst$_{sf}$ will coincide with the state the \ensuremath{\mathsf{2WST}}$_{sf}$ is in, when reading cell $i$ for the first time. The variable update $\rho$ on each transition is defined as follows: for $\delta'((q,f),a)=(f'(q),f')$, the update $\rho((q,f),a)$ is defined as follows: \[ \rho((q,f),a)(Y)= \left \{ \begin{array}{ll} O\rho(X_q) & \mbox{if}~Y=O,\\ \gamma & \mbox{if}~Y=X_q~\mbox{and}~\delta(q,r,a,p)=(t, \gamma,1),\\ \gamma\rho(X_t) & \mbox{if}~Y=X_q~\mbox{and}~\delta(q,r,a,p)=(t, \gamma,0),\\ \gamma X_t \rho(X_{f(t)}) & \mbox{if}~Y=X_q~\mbox{and}~\delta(q,r,a,p)=(t, \gamma,-1)~\mbox{and}~f(t)~\mbox{is defined},\\ \epsilon & \mbox{otherwise}. \end{array}\right. \] At the moment, the \sst$_{sf}$ is not 1-bounded. One place where 1-boundedness is violated happens because it is possible to have $f(s_1)=f(s_2)=t$ for $s_1 \neq s_2$. In particular, assume $\delta(s_1,r,a,p)=\delta(s_2,r,a,p)=(t,\gamma,-1)$, and let $f(t)=q$. Then we have $X_{s_2}, X_{s_1}:=\gamma X_t \rho(X_q)$. This would mean that we visit the same cell in the same state $t$ twice, which would lead to non-determinism or not fully reading the input by the \ensuremath{\mathsf{2WST}}$_{sf}$. Thus, this cannot happen in an accepting run of the \ensuremath{\mathsf{2WST}}$_{sf}$. Hence, if we have $\delta'((q,f),r,a,p)=(f'(q), f')$ where $f'$ is a partial function which is not 1-1, we can safely replace it with a subset $f''$ of $f'$ which is 1-1. Since $f'(s_1)=f'(s_2)$ for $s_1 \neq s_2$ does not contribute to an accepting run, we can define $f''$ on those states which preserve the 1-1 property. The second place where 1-boundedness is violated is when the contents of $\rho(X_q)$ flows into $O$ as well as $X_q$. This can be fixed by composition with another \sst{} where on each transition, all variables other than $X_q$ are unchanged ($X:=X$ for $X \neq X_q$) and resetting $X_q$ ($X_q:=\epsilon$). Note that this does not affect the ouput since $O$ is the only variable that is responsible for the output, and $\rho(X_q)$ has faithfully been reflected into it. With these two fixes, we regain the 1-boundedness property. \subsection{Connecting runs of the \ensuremath{\mathsf{2WST}}$_{sf}$ and \sst$_{sf}$} \begin{lemma} Let $a_1a_2 \dots \in \Sigma^{\omega}$, and let $r=(q_0, i_0=0) \stackrel{r_1,s(i_0),p_1|z_1}{\rightarrow}(q_1,i_1)\stackrel{r_2,s(i_1),p_2|z_2}{\rightarrow}(q_2,i_2)\dots$ be an accepting run of the \ensuremath{\mathsf{2WST}}$_{sf}$. Let $j_n=min\{m \mid i_m=n\}$ be the index when cell $n$ is read for the first time. Let $r'=(q'_0, f_0) \stackrel{r'_1,a_1,p'_1}{\rightarrow} (q'_1, f_1) \stackrel{r'_2,a_2,p'_2}{\rightarrow}(q'_2, f_2) \dots$ be the corresponding run in the constructed \sst$_{sf}$. Then we have \begin{itemize} \item $q'_i=q_{j_i}, a_i=s(j_i), r'_i=r_{j_i}, p'_i=p_{j_i}$, \item the output $\sigma'(O)$ till position $i$ of the input is $z_1z_2 \dots z_{j_i}$ \item In short, if the $\ensuremath{\mathsf{2WST}}_{sf}$ reads $a_i$ with $r_i \in Q_B, p_i \in Q_A$, and state $q$, then the \ensuremath{\mathsf{2WST}}$_{sf}$ will be in state $f_i(q)$ when it next reads $a_{i+1}$, and produces $\sigma_{r', i}(X_q)$ as output. \end{itemize} \end{lemma} \begin{proof} The proof is by induction on the length of the prefixes of the run $r'$. Consider a prefix of length 1. In this case, we are at the first position of the input, in initial state $q'_0=q_0$, since the \ensuremath{\mathsf{2WST}}$_{sf}$ moves right. We also have the states $r'_1=r_1, p'_1=p_1$, and the output obtained as $O:=O \rho(X_{q_0})$, and $\rho(X_{q_0})=z_1$. Thus, for a prefix of length 1, the states coincide, and the output in $O$ coincides with the output produced so far. Assume that upto a prefix of length $k$, we obtain the conditions of the lemma, and consider a prefix of length $k+1$. \\ Let $r=(q_0, i_0=0) \stackrel{r_1,s(i_0),p_1|z_1}{\rightarrow}(q_1,i_1)\dots (q_{k-1},i_{k-1})\stackrel{r_{k},s(i_{k-1}),p_{k}|z_{k}}{\rightarrow}(q_k,i_k)\stackrel{r_{k+1},s(i_k),p_{k+1}|z_{k+1}}{\rightarrow}(q_{k+1},i_{k+1})\dots $ be a prefix of length $k+1$. The corresponding run $r'$ in the \sst$_{sf}$ is $r'=(q'_0, f_0) \stackrel{r'_1,a_1,p'_1}{\rightarrow} (q'_1, f_1) \dots (q'_{k-1}, f_{k-1}) \stackrel{r'_k,a_k,p'_k}{\rightarrow}(q'_k, f_k) \stackrel{r'_{k+1},a_{k+1},p'_{k+1}}{\rightarrow}(q'_{k+1},f_{k+1})\dots $, where by inductive hypothesis, we know that $q'_k=q_{j_k}, a_k=s_{j_k}, r'_k=r_{j_k}, p'_k=p_{j_k}$, and the output $\sigma'(O)$ till position $k$ of the input is $z_1z_2 \dots z_{j_k}$. $q'_k=q_{j_k}$ is the state the \ensuremath{\mathsf{2WST}}$_{sf}$ is in, when it moves to the right of cell $k$ containing $a_k$, for the first time, and reads $a_{k+1}$. By construction of the \sst$_{sf}$, $\delta'((q_{j_k}, f), r_{j_k}, p_{j_k})=(f'(q_{j_k}), f')$, where $f'$ is defined based on the transition in the \ensuremath{\mathsf{2WST}}$_{sf}$ from state $q_{j_k}$ on reading $a_{k+1}$. \begin{enumerate} \item If $\delta(q_{j_k},a,r_{j_k}, p_{j_k})=(t, \gamma,1)$, then by construction of the \sst$_{sf}$, we obtain $f'(q_{j_k})=t$ and $\rho((q_{j_k},f),a)(O)=O\rho(X_{q_{j_k}})=O\gamma$. By inductive hypothesis, the contents of $O$ agrees with the output till $k$ steps. By catenating $\gamma$, in this case also, we obtain the same situation. \item If $\delta(q_{j_k},a,r_{j_k}, p_{j_k})=(t, \gamma,0)$, then by construction of the \sst$_{sf}$, we obtain $f'(q_{j_k})=f'(t)$ and $\rho((q_{j_k},f),a)(O)=O\gamma\rho(X_t)$. Inducting on the number of steps it takes for the \ensuremath{\mathsf{2WST}}$_{sf}$ to move to the right of cell $k+1$, we can show that the condition in the lemma holds. \begin{itemize} \item If the \ensuremath{\mathsf{2WST}}$_{sf}$ moves to the right of cell $k+1$ in the next step (one step), in state $u=q'_{k+1}$, then we obtain $f'(t)=u$, where $\delta(t,a,r,p)=(u,\gamma',1)$, in which case we obtain $\rho((q_{j_k},f),a)(O)=O\gamma\gamma'$ ($O \rho(X_t)$, where the current contents of $O$ is the old contents of $O$ followed by $\gamma$ and $\rho(X_t)=\gamma'$) which indeed is the output obtained in the \ensuremath{\mathsf{2WST}}$_{sf}$ when it moves to the right of cell $k+1$. \item Assume as inductive hypothesis, that the result holds (that is, the output $O$ agrees with the output in the \ensuremath{\mathsf{2WST}}$_{sf}$ so far, and that the state of the $\sst_{sf}$ agrees with the state of the \ensuremath{\mathsf{2WST}}$_{sf}$) when it moves to the right of cell $k$ in $\leq m$ steps. Consider now the case when it moves to the right of cell $k+1$ in $m+1$ steps. At the end of $m$ steps, the \ensuremath{\mathsf{2WST}}$_{sf}$ is on cell $k+1$, in some state $q$, and in the constructed \sst$_{sf}$, by hypothesis, the contents of $O$ correctly reflect the output so far. By assumption, we have $\delta(q, a_{k+1}, r,p)=(q'', \gamma'',1)$ for some state $q''$. Then we obtain, by the inductive hypothesis applied to cell $k$ at the end of $m$ steps, and the construction of the \sst$_{sf}$, the \sst$_{sf}$ to be in state $(f'(q),f')$, where $f'(q)=q''$, and $\rho((q'',f),a_{k+1})(O)=O\gamma''$. Since $O$ is up-to-date with the output with respect to the \ensuremath{\mathsf{2WST}}$_{sf}$, catenating $\gamma''$ indeed keeps it up-to-date, and the state of the \sst$_{sf}$ $q'_{k+1}$ is indeed the state the \ensuremath{\mathsf{2WST}}$_{sf}$ is in, when it moves to the right of cell $k+1$ for the first time. \end{itemize} \item If $\delta(q_{j_k},a,r_{j_k}, p_{j_k})=(t, \gamma,-1)$, then by construction of the \sst$_{sf}$, we obtain $f'(q_{j_k})=f'(f(t))$ and $\rho((q_{j_k},f),a)(O)=O\gamma X_t \rho(X_{f(t)})$. The \ensuremath{\mathsf{2WST}}$_{sf}$ is in cell $k$ at this point of time. Inducting on the number of steps it takes for the \ensuremath{\mathsf{2WST}}$_{sf}$ to move to the right of cell $k+1$ as above, we can show that the condition in the lemma holds. \end{enumerate} \end{proof} \subsection{Aperiodicity of the \sst$_{sf}$} \label{app:aper-sst} Having constructed the \sst$_{sf}$, the next task is to argue that the \sst$_{sf}$ is aperiodic. \begin{lemma} The constructed \sst$_{sf}$ is aperiodic. \end{lemma} \begin{proof} We start with an aperiodic \ensuremath{\mathsf{2WST}}$_{sf}$ $\mathcal{A}$ and obtain the \sst$_{sf}$ $\mathcal{T}$. Since $\mathcal{A}$ is aperiodic, there exists an $m \in \mathbb N$ such that $u^m \in L_{pq}^{xy}$ iff $u^{m+1} \in L_{pq}^{xy}$ for all $p, q \in Q$ and $x, y \in \{\ell, r\}$. To show the aperiodicity of the \sst$_{sf}$, we have to show that the transition monoid which captures state and variable flow is aperiodic. Lets consider a string $u \in \Sigma^*$ and a run in the \sst$_{sf}$ $(q,f) \stackrel{u}{\rightarrow} (q',f')$. Based on the transitions of the \sst$_{sf}$, we make a basic observation on variable assignments as follows. We show a correspondence between the flow of states for each kind of traversal in the \ensuremath{\mathsf{2WST}}$_{sf}$, and the respective state and variable flow in the constructed \sst$_{sf}$. Using this information and the aperiodicity of the \ensuremath{\mathsf{2WST}}$_{sf}$, we prove the aperiodicity of the \sst$_{sf}$ by showing the state, variable flow to be identical for strings $w^m, w^{m+1}$, for some chosen $m \in \mathbb{N}$, and any $w$. We know already that if $u=a$, and if $\delta(q,r,a,p)=(\gamma, q', 1)$ in the \ensuremath{\mathsf{2WST}}$_{sf}$ $\mathcal{A}$, then in the \sst$_{sf}$ we have the variable update $\rho(X_q)=\gamma$. If $u=a$, and if $\delta(q,r,a,p)=(\gamma, q', -1)$ in the \ensuremath{\mathsf{2WST}}$_{sf}$ $\mathcal{A}$, then in the \sst{} we have the variable update $\rho(X_q)=\gamma X_{q'} \rho(X_{f(q')})$. We now generalize the above for any $u \in \Sigma^*$ as follows: \begin{itemize} \item[(a)] If the run of $u$ in the \ensuremath{\mathsf{2WST}}$_{sf}$ $\mathcal{A}$ starts at the rightmost letter of $u$ in state $q$, and leaves it at the right with output $\gamma$, then we have $\rho(X_q)=\gamma$. \item[(b)] Assume the run of $u$ in the \ensuremath{\mathsf{2WST}}$_{sf}$ $\mathcal{A}$ starts at the rightmost letter of $u$ in state $q$, and leaves it at the left in state $t_1$. Since the whole input is read, eventually the \ensuremath{\mathsf{2WST}}$_{sf}$ will leave $u$ at the right in state $f'(q)$. Then we have $\rho(X_q)=\gamma_0 X_{t_1} \gamma_1 X_{t_2} \dots \gamma_{n-1} X_{t_n} \gamma_n$ such that (i) $u \in L_{q,t_1}^{r\ell}$ with output $\gamma_0$, (ii) $u \in L_{f(t_1),t_2}^{\ell\ell}$ with output $\gamma_1$. In general, $u \in L_{f(t_i),t_{i+1}}^{\ell\ell}$ with output $\gamma_i$, and (iii) since $\mathcal{A}$ completely reads the input, at some point, $u \in L_{f(t_n),f'(q)}^{\ell,r}$ with output $\gamma_n$. Also, in state $t_i$, we move one position left from some cell $h$, the leftmost position of $u$, and in state $f(t_i)$ we move to the right of cell $h$ for the first time, and the output generated in the interim is obtained from $\rho(X_{t_i})$. \end{itemize} We will show the above (a), (b) by inducting on $u$. For the base case $u=a$, we already have the proof, by the construction of the updates in the \sst$_{sf}$. Let $p$ be any state. Consider the case when in the \ensuremath{\mathsf{2WST}}$_{sf}$ we start at the right of $ua$ in state $p$ and leave it on the right. \begin{itemize} \item If we start at state $p$ on $a$, and $\delta(p,a)=(\gamma, q, -1)$, then we are at the right of $u$ in state $q$. We will be back on $a$ in state $f(q)$. By inductive hypothesis on $u$, we have $\rho(X_q)$ as some string over $\Gamma^*$, since we will leave $u$ at the right starting in state $q$ in the right. Again, by inductive hypothesis for $a$, we have $\rho(X_p)=\gamma X_q \rho(X_{f(q)})$; by the inductive hypothesis on $u$, when we leave $u$ on the right starting in state $q$ on the rightmost symbol of $u$, we obtain $\rho(X_q)$ as a constant. We are now on $a$ in state $f(q)$. If we leave $a$ to the right starting at $f(q)$, then we have by inductive hypothesis, $\rho(X_{f(q)})$ is a constant, and then we obtain $\rho(X_p)$ as a constant, while considering moving to the right of $ua$ starting at $a$ in state $p$. This agrees with (a) above. \item If we start at state $p$ on $a$, and $\delta(p,a)=(\gamma, q, 0)$. By construction of the \sst$_{sf}$, we have $\rho(X_p)=\gamma \rho(X_q)$. As long as we stay in the same position, we continue obtaining the same situation. If we leave $a$ to the right eventually, then by inductive hypothesis, we obtain $\rho(X_p)$ as a string over $\Gamma^*$. If from some state $t$, while on $a$, we move one position to the left, (on the rightmost symbol of $u$), then by inductive hypothesis on $u$, we obtain $\rho(X_t)$ as a constant; finally, since we move to the right of $a$ starting in state $f(t)$, we obtain $\rho(X_{f(t)})$ as a string over $\Gamma^*$, and hence, $\rho(X_p)$ as well. Thus, the inductive hypothesis works since $f(p)=f(q)$. \item If we start at state $p$ on $a$, and $\delta(p,a)=(\gamma, q, 1)$, then we straightaway have $\rho(X_p)$ as a constant, and are done. \end{itemize} Now consider the case when we start at the right of $ua$ and leave it at the left. Let us start in state $p$ on $a$. \begin{itemize} \item If we start at state $p$ on $a$, and $\delta(p,a)=(\gamma, q, -1)$, then we are at the right of $u$ in state $q$. By inductive hypothesis for $a$ we also have $\rho(X_p)=\gamma X_q \rho(X_{f'(q)})$. If we leave $ua$ at the left in state $r$, then starting in the right of $u$ in state $q$, we leave it at the left in $r$. By inductive hypothesis on $u$, we have $\rho(X_q)=\gamma_0 X_{r} \dots X_{r'}\gamma'$, such that $u \in L_{f(r'),f'(q)}^{\ell r}$. Leaving $a$ on the right from state $f'(q)$ gives by inductive hypothesis $\rho(X_{f'(q)})$ as a constant string $\beta$. On $ua$ starting in state $q$ on the right, the content of $X_p$ is then $\gamma \gamma_0 X_r \dots X_{r'} \gamma' \rho(\rho(X_{f'(q)}))$ =$\gamma \gamma_0 X_r \dots X_{r'} \gamma' \beta$, which agrees with (b), since $r, \dots, r'$ are the states in which we leave $u$ at the left. \item If we start at state $p$ on $a$, and $\delta(p,a)=(\gamma, q, 0)$. Then the inductive hypothesis works since $f(p)=f(q)$. \item Lastly, the case $\delta(p,a)=(\gamma, q, 1)$ does not apply since we are leaving $ua$ at the left starting on $a$ in state $p$. \end{itemize} The aperiodicity of the \sst$_{sf}$ is now proved as follows. Since the \ensuremath{\mathsf{2WST}}$_{sf}$ is aperiodic, we know that there is an $m \in \mathbb{N}$ such that $u^m \in L_{pq}^{xy}$ iff $u^{m+1} \in L_{pq}^{xy}$. Moreover, we also know that the matrices of $u^m, u^{m+1}$ are identical in the \ensuremath{\mathsf{2WST}}$_{sf}$, which tells us that the sequence of states seen in the left, right traversals of $u^m, u^{m+1}$ are identical. If this is the case, then in the above argument, we obtain the variable substitutions and state, variable flow for $u^m$ and $u^{m+1}$ to be identical, since the state and variable flow of the \sst$_{sf}$ only depends on the state flow of the \ensuremath{\mathsf{2WST}}$_{sf}$. Then we obtain the transition monoids of $u^m, u^{m+1}$ to be the same in the \sst$_{sf}$. The states of the look-behind and look-ahead automata in the \sst$_{sf}$ also follow the same sequence of states of the look-behind and look-ahead automata in the \ensuremath{\mathsf{2WST}}$_{sf}$. Since the look-behind and look-ahead automata are aperiodic, we obtain $M_{u^m}=M_{u^{m+1}}$ in the \sst$_{sf}$. Hence the \sst$_{sf}$ is aperiodic. \end{proof} \section{Example of an Aperiodic Monoid Recognizing a Language} \label{app:example-aperiodic} The language $L=(ab)^{\omega}$ is aperiodic, since it is recognized by the morphism $h: \{a,b\}^* \rightarrow M$ where $M$ is an aperiodic monoid. $M=(\{1,2,3\}, ., 1)$ with 2.3=1, 3.2=3, 2.2=3, 3.3=3 and $x.1=1.x=x$ for all $x \in M$. Define $h(\epsilon)=1, h(a)=2, h(b)=3$. Then $M$ is aperiodic, as $x^n=x^{n+1}$ for all $x \in M$ and $n \geq 3$. It is clear that $h$ recognizes $L$. \section{First-Order Logic: Examples} \label{app:fo} We define the following useful FO-shorthands. \begin{itemize} \item $x \succ y \rmdef \neg (x \preceq y)$ and $x \prec y \rmdef (x \preceq y) \wedge \neg (x = y)$, \item $first(x)=\neg \exists y(y \prec x)$ \item \texttt{is\_string}{} is defined as \[ \forall x(~\texttt{u\_succ}(x) \wedge \texttt{u\_pred}(x))\wedge \exists y[\texttt{first}(y) \wedge \forall z[\texttt{first}(z) \rightarrow z=y]] \wedge \forall x \exists y [x \prec y \wedge x \neq y] \] where $\texttt{u\_succ}(x)=\{\exists y[x \prec y \wedge \neg \exists z[x \prec z \wedge z \prec y]] \wedge \exists y'[x \prec y' \wedge \neg \exists z[x \prec z \wedge z \prec y']] \rightarrow (y=y')\} \wedge \exists y[x \prec y \wedge \neg \exists z[x \prec z \wedge z \prec y]]$ and \\ $\texttt{u\_pred}(x)=\{\exists y[y \prec x \wedge \neg \exists z[y \prec z \wedge z \prec x]] \wedge \exists y'[y' \prec x \wedge \neg \exists z[y' \prec z \wedge z \prec x]] \rightarrow (y=y')\} \wedge \exists y[y \prec x \wedge \neg \exists z[y \prec z \wedge z \prec x]]$ characterize that the position $x$ has unique predecessor and successor. \end{itemize} It is easy to see that a structure satisfying $\texttt{is\_string}$ characterizes a string. \section{Proofs from Section \ref{ap-omega}} \subsection{Transition Monoid of Muller Automata} \label{app:basic-muller} We start with an example for a transition monoid. The Muller automaton given in figure \ref{fig:muller-ex1} has two Muller acceptance sets $\{q\}, \{r\}$. Consider the strings $ab$ and $bb$. The transition monoids are \[ M_{ab}=\begin{blockarray}{cccc} & \matindex{$q$} & \matindex{$r$} & \matindex{$t$} \\ \begin{block}{c(ccc)} \matindex{$q$} & \bot & (0,0) & \bot \\ \matindex{$r$} & \bot & \bot & (0,0) \\ \matindex{$t$} & (0,0) & \bot & \bot \\ \end{block} \end{blockarray} ~~M_{bb}=\begin{blockarray}{cccc} & \matindex{$q$} & \matindex{$r$} & \matindex{$t$} \\ \begin{block}{c(ccc)} \matindex{$q$} & (1,0) & \bot & \bot \\ \matindex{$r$} & \bot & (0,0) & \bot \\ \matindex{$t$} & \bot & \bot & (0,0) \\ \end{block} \end{blockarray} \] \input{muller-ex} \begin{lemma} $(M_T,\times,\textbf{1})$ is a monoid, where $\times$ is defined as matrix multiplication and the identity element $\textbf{1}$ is the matrix with diagonal elements $(\emptyset,\emptyset,\dots,\emptyset)$ and all non-diagonal elements being $\bot$. \end{lemma} \begin{proof} Consider any matrix $M_s$ where $s \in \Sigma^*$. Let there be $m$ states $\{p_1, \dots, p_m\}$ in the Muller automaton. Consider a row corresponding to some $p_i$. Only one entry can be different from $\bot$. Let this entry be $[p_i][p_j] = (\kappa_1, \dots, \kappa_n)$, where each $\kappa_h \in \{0,1,2^Q\}$, $1 \leq h \leq n$. Consider $M_s \times \textbf{1}$. The $[p_i][p_j]$ entry of the product are obtained from the $p_i$th row of $M_s$ and the $p_j$th column of $\textbf{1}$. The $p_j$th column of $\textbf{1}$ has exactly one entry $[p_j][p_j] = (\emptyset, \dots, \emptyset)$, while all other elements are $\bot$. Then the $[p_i][p_j]$ entry for the product matrix $M_s \times \textbf{1}$ is of the form $\bot \oplus \dots \oplus \bot \oplus (\kappa_1, \dots, \kappa_n) \odot (\emptyset, \dots, \emptyset) \oplus \bot \oplus \dots \oplus \bot$. Clearly, this is equal to $(\kappa_1, \dots, \kappa_n)$, since $\kappa \odot \emptyset=\kappa \cup \emptyset=\kappa$ and $\bot \oplus \kappa=\kappa$. Similarly, it can be shown that the $[p_i][p_j]$th entry of $M_s$ is preserved in $\textbf{1} \times M_s$ as well. Associativity of matrix multiplication follows easily. \end{proof} We exploit the following lemma, proved in \cite{dg08SIWT}, in our proofs. \begin{lemma} \label{lem-gd} Let $h: \Sigma^* \rightarrow M$ be a morphism to a finite monoid $M$ and let $w=u_0u_1\dots$ be an infinite word with $u_i \in \Sigma^+$ for $i \geq 0$. Then there exist $s, e \in M$ and an increasing sequence $0< p_1 < p_2 < \dots$ such that \begin{enumerate} \item $se=s$ and $e^2=e$ \item $h(u_0\dots u_{p_1-1})=s$ and $h(u_{p_i} \dots u_{p_j})=e$ for all $0 < i<j$. \end{enumerate} \end{lemma} Using Lemma \ref{lem-gd}, we prove the following lemma, which is used in the proof of Lemma \ref{lem-conv}. \begin{lemma} \label{lem:ap-rec} Let $\mathcal{A}$ be a DMA. The mapping $h$ which maps any string $s$ to its transition matrix $M_s$, is a morphism from $(\Sigma^*, ., \epsilon)$ to $(M_T, \times, \textbf{1})$. Hence, $h$ recognizes $L(\mathcal{A})$. \end{lemma} \begin{proof} It is easy to see that $h$ is a morphism : $h(s_1s_2)$ is by definition, $M_{s_1s_2}$. This is clearly equal to $M_{s_1}M_{s_2}=h(s_1)h(s_2)$. Let $w \in L(\mathcal{A})$ and let $w = w_1 w_2 w_3 w_4 \dots$ be a factorization of $w$, with $w_i \in \Sigma^+$ for all $i$. Consider another string $w' \in \Sigma^{\omega}$ with a factorization $w' = w'_1 w'_2 w'_3 w'_4 \dots$, such that $h(w_i) = h(w'_i)$ forall $i$. We have to show that $w' \in L(\mathcal{A})$. Since $w \in L(\mathcal{A})$ we know that after some position $i$, the run $r$ of $w$ will contain only and all states from some muller set $F_k$. Assume that the factorization of $w$ is such that, after the factor $w_i$ of $w$, the run visits only states from $F_k$. We can write $w$ as $u v_1 v_2 v_3 \dots$ such that $u = w_1 w_2 \dots w_i$, $v_1 = w_{i+1} w_{i+2} \dots w_{i+j_1}$, $v_2 = w_{i+j_1+1} w_{i+j_1+2} \dots w_{i+j_1+j_2}$ and so on, such that each of $v_1, v_2, \dots $ witness all states of $F_k$ in the run. Since $h(w_i) = h(w'_i)$ forall $i$, we know that $M_{w_i}=M_{w'_i}$ for all $i$. So we can factorize $w'$ as $u'v'_1v'_2 \dots$ such that $h(u)=h(u'), h(v_i)=h(v'_i)$ for all $i$. Hence, $w'$ also has a run that also witnesses the muller set $F_k$ from some point onwards. Hence, $w' \in L(\mathcal{A})$. Thus, $h$ recognizes $L(\mathcal{A})$ since $w \in L(\mathcal{A}) \rightarrow [w]_h \subseteq L(\mathcal{A})$. \end{proof} \subsection{Aperiodicty of DMA $\mathcal{A}$ $\equiv$ Aperiodicity of $L(\mathcal{A})$} \label{app:aper-equiv} Note that a result similar to Theorem \ref{muller-ap} (whose proof is below) has been proved for B\"uchi automata in \cite{dg08SIWT}, where aperiodicity of B\"uchi automata was defined using transition monoids. \begin{theorem} \label{muller-ap} A language $L \subseteq \Sigma^{\omega}$ is aperiodic iff there exists an aperiodic Muller automaton $\mathcal{A}$ such that $L = L(\mathcal{A})$. \end{theorem} \begin{proof} We obtain the proof of Theorem \ref{muller-ap} by proving the following two lemmas. \begin{lemma} \label{lem-forward} Let $L \subseteq \Sigma^{\omega}$ be an aperiodic language. Then there is an aperiodic Muller automaton accepting $L$. \end{lemma} \begin{proof} Let $L \subseteq \Sigma^{\omega}$ be an aperiodic language. Then by definition, $L$ is recognized by a morphism $h: \Sigma^* \rightarrow M$ where $M$ is a finite aperiodic monoid. We first show that we can construct a counter-free Muller automaton $\mathcal{A}$ such that $L=L(\mathcal{A})$. A counter-free automaton is one having the property that $u^m \in L_{pp} \rightarrow u \in L_{pp}$ for all $u \in \Sigma^*$, states $p$ and $m \geq 1$. It can be shown \cite{dg08SIWT} that the aperiodic language $L$ can be written as the finite union of languages $UV^{\omega}$ where $U, V$ are aperiodic languages of finite words. Moreover, for any $u_0u_1u_2 \dots \in \Sigma^{\omega}$, there is an increasing sequence $0<p_1<p_2 \dots$ of natural numbers such that for the morphism $h: \Sigma^* \rightarrow M$ recognizing $L$, we have $h(u_0\dots u_{p_1})=s \in M$ and $h(u_{p_i}\dots u_{p_j})=e \in M$ for all $0<i<j$. The $e \in M$ is an idempotent element in $M$. Then we have $h^{-1}(e)=V$ and $h^{-1}(s)=U$. Since $V$ is an aperiodic language $\subseteq \Sigma^*$, there is a minimal DFA $\mathcal{D}$ that accepts $V$. The initial state of this DFA is $[\epsilon]$, and the states are of the form $[x], x \in \Sigma^*$, such that $xw \in L$ iff there is a run from $[x]$ on $w$ to an accepting state. Accepting states have the form $[w]$ with $w \in V$ and the transition function is $\delta([x],a)=[xa]$. For all $x,y,v \in \Sigma^*$, ($xv \in V$ iff $yv \in V$) $\Rightarrow [x]=[y]$. We first show that $\mathcal{D}$ is counter-free. Since $V$ is aperiodic, there is a morphism $g: V \rightarrow M_V$ recognizing $V$, for an aperiodic monoid $M_V$. As $M_V$ is aperiodic, there exists some $m \geq 1$ such that for all $x \in M_V$, $x^m=x^{m+1}$. If $[u]=[uv^m]$ for some $u, v \in \Sigma^*$, then $g(u)=g(uv^m)=g(uv^{m+1})=g(uv^mv)=g(uv)$. If $[u] \neq [uv]$, then we can find some string $w$ such that $uw \in V$ but $uvw \notin V$ or viceversa. This contradicts the hypothesis that $V$ is recognized by a morphism $g: \Sigma^* \rightarrow M_V$, since $uw \in V$ and $g(uw)=g(uvw)$ implies either both $uw, uvw$ belong to $V$ or neither. Thus, $[u]=[uv^m]$ implies $[u]=[uv]$ for all $u, v \in \Sigma^*$. That is, whenever $\delta^*([u],v^m)=[uv^m]=[u]$, we have $\delta^*([u],v)=[uv]=[u]$ as well, which shows that the minimal DFA $\mathcal{D}$ for $V$ is counter-free. If we intepret $\mathcal{D}$ as a Buchi automaton, and consider $\alpha \in L(\mathcal{D})$, then $\alpha$ has infinitely many prefixes $v_1 < v_2 < \dots$ such that each $v_i \in V$. We have to show that $\alpha \in V^{\omega}$; that is we have to show that $\alpha=\alpha_1\alpha_2 \dots$ with $\alpha_i \in V$ for all $i$. We know $v_2=v_1v'_1$, $v_3=v_2v'_2 \dots$ with $v_1, v_2, \dots \in V$. Then $\alpha=v_1v'_1v'_2v'_3 \dots$. If $v'_i \in V$ for all $i$, we are done, since in that case, $\mathcal{D}$ will be a counter-free Buchi automaton for $V^{\omega}$. To obtain the infinitely many prefixes $v_1<v_2<v_3<\dots$ such that $v_{i+1}=v_iv'_i$ with $v_i, v'_i \in V$, we consider the language $W=V.Prefree(V)$ where $Prefree(V)$ is the set of all strings in $V$ which do not contain a proper prefix also lying in $V$. Since $V$ is aperiodic, $W$ is also aperiodic. Let $\mathcal{E}$ be the minimal counter free DFA accepting $W$. We intrepret $\mathcal{E}$ as a Buchi automaton, as we did for the case of $\mathcal{D}$, and show that $\mathcal{E}$ is a counter-free Buchi automaton accepting $V^{\omega}$. Clearly, $L(\mathcal{E})$ is the set of strings which has infinitely many prefixes from $W$. If $w \in V^{\omega}$, then $w \in L(\mathcal{E})$ since $w$ has infinitely many prefixes from $W$. Conversely, let $w \in \Sigma^{\omega}$ be such that infinitely many prefixes $w_1<w_2<w_3< \dots$ of $w$ are in $W$. Then we have to show that $w \in V^{\omega}$. Let $w_i=x_iy_i$ where $x_i \in V, y_i \in Prefree(V)$ for each $i$. Then we have $$x_1 < x_1y_1 < x_2 < x_2y_2 < x_3 < \dots$$ Note that if $w_1\neq w_2$ that is, $x_1y_1 \neq x_2y_2$ and $x_1=x_2$, then $y_1$ is a prefix of $y_2$. Since $y_1, y_2 \in Prefree(V)$, this is not possible. Thus, for any two $w_i \neq w_j$, we have $x_i < x_j$. Let $x_{i+1}=x_iy_iy'_i$ for some $y'_i$. Recall that, using the morphism $h: \Sigma^* \rightarrow M$ recognizing $L$, we have $h^{-1}(e)=V$ for some idempotent $e \in M$. $h(x_{i+1})= h(x_iy_iy'_i)=e.e.h(y'_i)=e.h(y'_i)=h(y_iy'_i)$. Hence we get $w=x_1y_1y'_1y_2y'_2y_3y'_3\dots=x_1x_2x_3\dots \in V^{\omega}$. Thus, $\mathcal{E}$ is a counter-free Buchi automaton accepting $V^{\omega}$. Let $\mathcal{E}'$ be the counter-free Muller automaton obtained from $\mathcal{E}$. Since $U \subseteq \Sigma^*$ is aperiodic, we can construct as for $V$, the minimal counter-free DFA $\mathcal{U}$ for $U$. The concatenation $\mathcal{U} \mathcal{E}'$ is then counter-free. The finite union of such automata are also counter-free. Finally we show that counter-free automata are aperiodic. Let $x^{n} \in L_{pq}$ for any two states $p,q$, for a large $n$. We can decompose $x^n$ as $x^{k+l+m}$ such that $x^k \in L_{ps}, x^l \in L_{ss}$ and $x^m \in L_{sq}$, with $l \geq 2$. Then we have $x \in L_{ss}$ by the counter-freeness. Then we obtain $x^{n-1}\in L_{pq}$. Similarly, we can show that $x^{n-1} \in L_{pq} \Rightarrow x^n \in L_{pq}$. This shows that we have an aperiodic Muller automata accepting the finite union $UV^{\omega}$ that represents $L$. Thus, starting from the assumption that $L$ is an aperiodic language, we have obtained an aperiodic Muller automaton that accepts $L$. \end{proof} \begin{lemma} \label{lem-conv} Let $\mathcal{A}$ be Muller automaton whose transition monoid is aperiodic. Then $L(\mathcal{A})$ is aperiodic. \end{lemma} \begin{proof} From Lemma \ref{lem:ap-rec}, we know that we can construct a morphism mapping $(\Sigma^*, ., \epsilon)$ to the transition monoid of $\mathcal{A}$ which recognizes $L(\mathcal{A})$. Hence, $L(\mathcal{A})$ is aperiodic. \end{proof} \end{proof} \section{Example of FOT} \label{app:fot-example} \begin{figure}[h] \tikzstyle{trans}=[-latex, rounded corners] \begin{center} \scalebox{0.9}{ \begin{tikzpicture}[->,>=stealth',shorten >=1pt,auto, semithick,scale=.8] \tikzstyle{every state}=[fill=golden] \node[loc] at (-2, 0) (B) {$a$} ; \node[loc] at (0,0) (B0) {$b$} ; \node[loc] at (2,0) (B1) {$b$} ; \node[loc] at (4,0) (B2) {$b$} ; \node[loc] at (6,0) (B3) {$\#$} ; \node[loc] at (8,0) (B4) {$b$} ; \node[loc] at (10,0) (B5) {$a$} ; \node[loc] at (12,0) (B6) {$\#$} ; \node[loc,dashed] at (13.5,0) (B7) {$\{a,b\}^\omega$} ; \draw[trans] (B) -- (B0); \draw[trans] (B0) -- (B1); \draw[trans] (B1) -- (B2); \draw[trans] (B2) -- (B3); \draw[trans] (B3) -- (B4); \draw[trans] (B4) -- (B5); \draw[trans] (B5) -- (B6); \draw[trans] (B6) -- (B7); \node[loc] at (-2, -1) (C) {$a$} ; \node[loc] at (0,-1) (C0) {$b$} ; \node[loc] at (2,-1) (C1) {$b$} ; \node[loc] at (4,-1) (C2) {$b$} ; \node[loc] at (8,-1) (C4) {$b$} ; \node[loc] at (10,-1) (C5) {$a$} ; \node at (13.5,-1) (C7) {$~~~~~~$} ; \draw[trans] (C) -- (C0); \draw[trans] (C0) -- (C1); \draw[trans] (C1) -- (C2); \draw[trans] (C4) -- (C5); \node[loc] at (-2, -2) (D) {$a$} ; \node[loc] at (0,-2) (D0) {$b$} ; \node[loc] at (2,-2) (D1) {$b$} ; \node[loc] at (4,-2) (D2) {$b$} ; \node[loc] at (8,-2) (D4) {$b$} ; \node[loc] at (10,-2) (D5) {$a$} ; \node at (13.5,-2) (D7) {$~~~~~~$} ; \draw[trans] (D) -- (C); \draw[trans] (D2) -- (D1); \draw[trans] (D1) -- (D0); \draw[trans] (D0) -- (D); \draw[trans] (D5) -- (D4); \node at (-2,-3) (E) {$~~$} ; \node[loc] at (6,-3) (E3) {$\#$} ; \node[loc] at (12,-3) (E6) {$\#$} ; \node[loc,dashed] at (13.5,-3) (E7) {$\{a,b\}^\omega$} ; \draw[->,rounded corners] (C2) -- (6, -1) -- (E3); \draw[trans] (D4) -- (C4); \draw[->, rounded corners] (E3) -- (10, -3) -- (D5); \draw[->, rounded corners] (C5) -- (12, -1) -- (E6); \draw[trans] (E6) -- (E7); \begin{pgfonlayer}{background} \node at (-2.2, 0) [label=left:\texttt{input} :] {}; \node [background, fit=(C) (C7), label=left:\texttt{copy 1}:] {}; \node [background, fit=(D) (D7), label=left:\texttt{copy 2}:] {}; \node [background, fit=(E) (E7), label=left:\texttt{copy 3}:] {}; \end{pgfonlayer} \end{tikzpicture} } \end{center} \vspace{-1em} \caption{Transformation $f_1$ given as FO-definable transformation for the string $abbb\#ba\#\{a,b\}^\omega$. \label{fig:app-fo-example}} \vspace{-1em} \end{figure} We give the full list of $\phi^{c,d}$ here. Let $btw(x,y,z)=(y \prec z \prec x) \vee (x \prec z \prec y)$ be a shorthand that says that $z$ lies between $x, y$. Let $btw(x,y,\gamma)=\exists z(L_{\gamma}(z) \wedge btw(x,y,z))$ for $\gamma \in \Gamma$ and let $reach_{\#}(x)=\exists y(x \prec y \wedge L_{\#}(y))$ be a shorthand which says there is a $\#$ that is ahead of $x$. \begin{enumerate} \item $\phi_{dom}=\texttt{is\_string}_{\#}$, \item $\phi^1_{\gamma}(x) = \phi^2_{\gamma}(x) = L_{\gamma}(x) \wedge \neg L_{\#}(x) \wedge \texttt{reach}_{\#}(x)$, since we only keep the non $\#$ symbols that can ``reach'' a $\#$ in the input string in the first two copies. \item $\phi^3_{\gamma}(x) = L_{\#}(x) \vee (\neg L_{\#}(x) \wedge \neg \texttt{reach}_{\#}(x))$, since we only keep the $\#$'s, and the infinite suffix from where there are no $\#$'s. \end{enumerate} The transitive closure of the output successor relation is given by : \begin{enumerate} \item $\phi^{1,1}_{\prec}(x,y)=(x \prec y)=\phi^{3,3}_{\prec}(x,y)$, \item $\phi^{2,2}_{\prec}(x,y)= [\neg btw(x,y,\#) \rightarrow (y \prec x) ]$ $\wedge [btw(x,y,\#) \rightarrow (x \prec y)]$ since we are reversing the arrows within a $\#$-free block from which a $\#$ is reachable. \item $\phi^{1,3}_{\prec}(x,y)=L_{\#}(y) \wedge (x \prec y)=\phi^{2,3}_{\prec}(x,y) \item $\phi^{1,2}_{\prec}(x,y)=x \prec y \wedge btw(x,y,\#)$ since each position in a $\#$-free block is related to each position in a $\#$-free block that comes later. \item $\phi^{3,2}_{\prec}(x,y)$ and $\phi^{3,1}_{\prec}(x,y)$ are given by $L_{\#}(x) \wedge (x \prec y)$ \item $\phi^{2,1}_{\prec}(x,y)$ expresses that each position $x$ in the $i$-th $\#$-free block is related to position $y$ appearing in the $j$-th $\#$-free block, $j > i$. Within a \#-free block, the arrows are reversed. This translates simply to \\ $[x \prec y \wedge btw(x,y,\#)] \vee \{\neg btw(x,y,\#) \wedge [(y \prec x) \vee y=x]\}$. \end{enumerate} \input{app-2way.tex} \input{app-sst.tex} \input{app-fot-2wst.tex} \section{Proofs from Section \ref{sec:sst-fot}} \label{app:varflow} \begin{proposition}(FO-definability of variable flow) \label{prop:foflow} Let $T$ be an aperiodic,1-bounded \sst{} $T$ with set of variables $\mathcal{X}$. For all variables $X,Y\in \mathcal{X}$, there exists an FO-formula $\phi_{X\rightsquigarrow Y}(x,y)$ with two free variables such that, for all strings $s\in dom(T)$ and any two positions $i\leq j\in dom(s)$, $s\models \phi_{X\rightsquigarrow Y}(i,j)$ iff $(q_i,X)\rightsquigarrow^{s[i{+}1{:}j]}_1 (q_j,Y)$, where $q_0\dots q_n \dots$ is the accepting run of $T$ on $s$. \end{proposition} Let $X\in\mathcal{X}$, $s\in\mathrm{dom}(T)$, $i\in \mathrm{dom}(s)$. Let $P \in 2^Q$ be a Muller accepting set with output $F(P)=X_1 \dots X_n$. Let $r = q_0\dots q_{n}\dots$ be an accepting run of $T$ on $s$. Then there is a position $j$ such that $\forall k >j$, only states in $P$ are visited. On all transitions beyond $j$, $X_1, \dots, X_{n-1}$ remain unchanged, while $X_n$ is updated as $X_nu$ for some $u \in (\mathcal{X} \cup \Gamma)^*$. We say that the pair $(X,i)$ is \emph{useful} if the content of variable $X$ before reading $s[i]$ will be part of the output after reading the whole string $s$. Formally, for an accepting run $r = q_0\dots q_{j}q_{j+1}\dots$, such that only states from the Muller set $P$ are seen after position $j$, we say that $(X,i)$ is useful for $s$ if $(q_{i-1}, X)\rightsquigarrow^{s[i{:}j] }_1(q,X_k)$ for some variable $X_k \in F(P)$, $k \leq n$, or if $(q_{i-1}, X)\rightsquigarrow^{s[i{:}\ell] }_1(q,X_n)$, with $q \in P$ and $\ell >j$. Thanks to Proposition \ref{prop:foflow}, this property is FO-definable. \begin{proposition}\label{prop:contribution} For all $X\in\mathcal{X}$, there exists an FO-formula $\text{useful}_X(i)$ s.t. for all strings $s\in dom(T)$ and all positions $i\in \mathrm{dom}(s)$, $s\models \text{useful}_X(i)$ iff $(X,i)$ is useful for string $s$. \end{proposition} Proofs of propositions \ref{prop:foflow} and \ref{prop:contribution} are below. \subsection{Proof of Proposition \ref{prop:foflow}} First, we show that reachable states in accepting runs of aperiodic \sst{} are FO-definable: \begin{proposition} \label{prop:fostates} Let $T$ be an aperiodic \sst{} $T$. For all states $q$, there exists an FO-formula $\phi_q(x)$ such that for all strings $s\in \Sigma^{\omega}$, for all positions $i$, $s\models \phi_q(i)$ iff $s\in dom(T)$ and the state of the (unique) accepting run of $T$ before reading the $i$-th symbol of $s$ is $q$. \end{proposition} \begin{proof} Let $A$ be the underlying (deterministic) aperiodic, Muller automaton of the SST $T$. Since $T$ is aperiodic, so is $A$. For all states $q$, let $L_q$ be the set of strings $s \in \Sigma^+$ such that there exists a run of $T$ on string $s$ that ends in state $q$. Clearly, $L_q$ can be defined by some aperiodic finite state automaton $A_q$ obtained by setting the set of final states of $A$ to $\{q\}$. Here, $A_q$ is interpreted as a DFA. Therefore $L_q$ is definable by some FO-formula $\psi^L_q$. Let $L_{q,p}$ be the set of strings that have a run in $T$ from state $q$ to state $p$. This also is obtained from $T$ by considering $q$ as the start state and $p$ as the accepting state. Let $\psi^L_{q,p}$ be the FO formula which captures this. For $P \in 2^Q$, let $R_P$ be the set of strings $s \in \Sigma^{\omega}$ such that there exists a run of $T$ on $s$ from some state $p \in P$, which stays in the set of states $P$, and all states of $P$ are witnessed infinitely often. Clearly, $u\in dom(T)$ iff there exists $p\in P$ with $v_1\in L_q, v_2 \in L_{q,p}$ and $v_3\in R_P$ such that $u=v_1v_2v_3$. To capture $v_3$, we have the FO-formula $\psi^{Rec}_P$ $\bigvee_{p \in P} L_p(x) \wedge \forall y(y \geq x \rightarrow \bigvee_{q \in P} L_q(y)) \wedge \forall y \exists k_1 \dots \exists k_{|P|}(k_i \geq y \wedge \bigwedge_{i \neq j}(k_i \neq k_j) \wedge \bigwedge_{q_i \in P} L_{q_i}(k_i))$. Then, $\phi_q(x)$ is defined as $$ \phi_q(x) = [\psi^L_q]_{\prec x} \wedge [\psi^L_{q,p}]_{x \prec y} \wedge [L_p(y) \wedge \psi^{Rec}_P]_{y\preceq} $$ where $[\psi^L_q]_{\prec x}$ is the formula $\psi^L_q$ in which all quantifications of any variable $z$ is guarded by $z\prec x$, $[\psi^L_{q,p}]_{x \prec y}$ is the formula where all variables lie in between $x,y$, and finally, $[\psi^{Rec}_P]_{y\preceq}$ is the formula $\psi^{Rec}_P$ in which all quantifications of any variable $z$ is guarded by $y\preceq z$. Therefore, $s\models \phi_q(i)$ iff $s[1{:}i)\in L_q$, $s[i{:}j)\in L_{q,p}$ and $s[j{:}\infty]\in R_P$. % \end{proof} Now we start the proof of Proposition \ref{prop:foflow}. \begin{proof} For all states $p,q\in Q$, let $L_{(p,X)\rightsquigarrow (q,Y)}$ be the language of strings $u$ such that $(p,X)\rightsquigarrow_1^{u} (q,Y)$. We show that $L_{(p,X)\rightsquigarrow (q,Y)}$ is an aperiodic language. It is indeed definable by an aperiodic non-deterministic automaton $A$ that keeps track of flow information when reading $u$. It is constructed from $T$ as follows. Its state set $Q'$ are pairs $(r,Z)\in 2^{Q\times \mathcal{X}}$. Its initial state is $\{(p,X)\}$ and final states are all states $P$ such that $(q,Y)\in P$. There exists a transition $P\xrightarrow{a} P'$ in $A$ iff for all $(p_2,X_2)\in P'$, there exists $(p_1,X_1)\in P$ and a transition $p_1\xrightarrow{a|\rho} p_2$ in $T$ such that $\rho(X_2)$ contains an occurrence of $X_1$. Note that by definition of $A$, there exists a run from a state $P$ to a state $P'$ on some $s\in\Sigma^*$ iff for all $(p_2,X_2)\in P'$, there exists $(p_1,X_1)\in P$ such that $(p_1,X_1)\rightsquigarrow^{s}_1 (p_2,X_2)$ (Remark $\star$). Clearly, $L(A) = L_{(p,X)\rightsquigarrow (q,Y)}$. It remains to show that $A$ is aperiodic, i.e. its transition monoid $M_A$ is aperiodic. Since $T$ is aperiodic, there exists $m\geq 0$ such that for all matrices $M\in M_T$, $M^m = M^{m+1}$. For $s\in\Sigma^*$, let $\Phi_A(s) \in M_A$ (resp. $\Phi_T(s)$) the square matrix of dimension $|Q'|$ (resp. $|Q|$) associated with $s$ in $M_A$ (resp. in $M_T$). We show that $\Phi_A(s^m) = \Phi_A(s^{m+1})$, i.e. $(P,P')\in \Phi_A(s^m)$ iff $(P,P')\in\Phi_A(s^{m+1})$, for all $P,P'\in Q'$. First, suppose that $(P,P')\in \Phi_A(s^m)$, and let $(p_2,X_2)\in P'$. By definition of $A$, there exists $(p_1,X_1)\in P$ such that $(p_1,X_1)\rightsquigarrow^{s^m}_1 (p_2,X_2)$, and by aperiodicity of $T$, it implies that $(p_1,X_1)\rightsquigarrow^{s^{m+1}}_1 (p_2,X_2)$. Since it is true for all $(p_2,X_2)\in P'$, it implies by Remark $(\star)$ that there exists a run of $A$ from $P$ to $P'$ on $s^{m+1}$, i.e. $(P,P')\in \Phi_A(s^{m+1})$. The converse is proved similarly. We have just proved that $L_{(p,X)\rightsquigarrow (q,Y)}$ is aperiodic. Therefore it is definable by some FO-formula $\phi_{(p,X)\rightsquigarrow (q,Y)}$. To capture the variable flow in an accepting run, we also need to have some conditions on the states $p$ and $q$. $\phi_{X\rightsquigarrow Y}(x,y)$ is defined by $$ \phi_{X\rightsquigarrow Y}(x,y)\equiv x\preceq y \wedge \bigvee_{q,p\in Q}\{ [\phi_{(q,X)\rightsquigarrow (p,Y)}]^{x\preceq \cdot\preceq y} \wedge [\psi^L_q]_{\prec x} \wedge [\psi^L_{q,p}]_{x \prec y} \wedge \exists z.\{[\psi^L_{p,r}]_{y \prec z} \wedge [L_r(z) \wedge \psi^{Rec}_P]_{z\preceq}\} \}$$ where $[\psi^L_q]_{\prec x}, [\psi^L_{q,p}]_{x \prec y}, [\psi^R_p]_{y\preceq}$ were defined in Proposition \ref{prop:fostates} and $[\phi_{(p,X)\rightsquigarrow (q,Y)}]^{x\preceq \cdot\preceq y}$ is obtained from $\phi_{(p,X)\rightsquigarrow (q,Y)}$ by guarding all the quantifications of any variable by $x\preceq z'\preceq y$. The Muller set $P$ starts from some position $z$ ahead of $y$, in some state $r \in P$. In the case when $p \in P$, consider $r=p$ in the above formula. \end{proof} \subsection{Proof of Proposition \ref{prop:contribution}} \label{app:useful} \begin{proof} The formula $\text{useful}_X(x)$ is defined by $$ \begin{array}{llllllll} \text{useful}_X(x) & = & \exists y\cdot[ \bigvee_{P \in 2^Q}\psi_P^{Rec}(y) \wedge \bigvee_{p \in P}\psi^L_p(y) \wedge \bigvee_{\{X_1, \dots, X_n \mid F(P)=X_1 \dots X_n\}} \Phi_{X\rightsquigarrow X_i}(x,y)] \end{array} $$ where $\psi_P^{Rec}(y)$ defines a position $y$ from where the Muller set $P$ is visited continously. $\psi_P^{Rec}(y)$ is defined in proposition \ref{prop:fostates} and $\Phi_{X\rightsquigarrow Y}(x,y)$ in proposition \ref{prop:foflow}. \end{proof} \subsection{Definition of \sst-output graphs} \label{app:sst-output} \input{outputgraph-app.tex} Figure \ref{fig:outputgraph} gives an example of \sst-output structure. We show only the variable updates. Dashed arrows represent variable updates for useless variables, and therefore does not belong to the \sst-output structure. Initially the variable content of $Z$ is equal to $\epsilon$. It is represented by the $\epsilon$-edge from $(Z^{in}, 0)$ to $(Z^{out}, 0)$ in the first column. Then, variable $Z$ is updated to $Zc$. Therefore, the new content of $Z$ starts with $\epsilon$ (represented by the $\epsilon$-edge from $(Z^{in},1)$ to $(Z^{in},0)$, which is concatenated with the previous content of $Z$, and then concatenated with $c$ (it is represented by the $c$-edge from $(Z^{out},0)$ to $(Z^{out},1)$). Note that the invariant is satisfied. The content of variable $X$ at position 5 is given by the label of the path from $(X^{in},5)$ to $(X^{out},5)$, which is $c$. Also note that some edges are labelled by strings with several letters, but there are finitely many possible such strings. In particular, we denote by $O_T$ the set of all strings that appear in right-hand side of variable updates. Let $T = (Q, q_0, \Sigma, \Gamma, \mathcal{X}, \delta, \rho, Q_f)$ be an \sst{}. Let $u\in (\Gamma\cup X)^*$ and $s\in \Gamma^*$. The string $s$ is said to \emph{occur} in $u$ if $s$ is a factor of $u$. In particular, $\epsilon$ occurs in $u$ for all $u$. Let $O_T$ be the set of constant strings occurring in variable updates, i.e. $O_T = \{ s\in\Gamma^*\ |\ \exists t\in \delta,\ s\text{ occurs in } \rho(t)\}$. Note that $O_T$ is finite since $\delta$ is finite. Let $w \in dom(T)$. The \emph{\sst-output graph} of $w$ by $T$, denoted by $G_T(w)$, is defined as an infinite directed graph whose edges are labelled by elements of $O_T$. Formally, it is the graph $G_T(w) = (V,(E_\gamma)_{\gamma\in O_T})$ where $V = \{0,1,\dots,\}\times \mathcal{X}\times \{in,out\}$ is the set of vertices, $E := \bigcup_{\gamma\in O_T} E_\gamma \subseteq V\times V$ is the set of labelled edges defined as follows. Vertices $(i,X,d)\in V$ are denoted by $(X^d,i)$. Let $r = q_0\dots q_n \dots $ be an accepting run of $T$ on $w$. The set $E$ is defined as the smallest set such that for all $X\in \mathcal{X}$, \begin{enumerate} \item $((X^{in},0),(X^{out},0))\in E_\epsilon$ if $(X,0)$ is useful, \item for all $i$ and $X\in X$, if $(X,i)$ is useful and if $\rho(q_{i},w[i+1],q_{i+1})(X) = \gamma$, then $((X^{in},i+1),(X^{out},i+1))\in E_\gamma$, \item for all $i$ and $X\in X$, if $(X,i)$ is useful and if $\rho(q_{i},w[i+1],q_{i+1})(X) = \gamma_1X_1\dots \gamma_kX_{k}\gamma_{k+1}$ (with $k>1$), then \begin{itemize} \item $((X^{in},i+1), (X_1^{in},i))\in E_{\gamma_1}$ \item $((X_k^{out},i), (X^{out},i+1))\in E_{\gamma_{k+1}}$ \item for all $1\leq j< k$, $((X_j^{out},i), (X_{j+1}^{in},i))\in E_{\gamma_{j+1}}$ \end{itemize} \end{enumerate} Note that since the transition monoid of $T$ is $1$-bounded, it is never the case that two copies of any variable (say $X$) flows into a variable (say $Y$), therefore this graph is well-defined and there are \textbf{no} multiple edges between two nodes. We next show that the transformation that maps an $\omega$-string $s$ into its output structure is FO-definable, whenever the SST is 1-bounded and aperiodic. Using the fact that variable flow is FO-definable, we show that for any two variables $X,Y$, we can capture in FO, a path from $(X^d, i)$ to $(Y^e, j)$ for $d, e\in \{in, out\}$ in $G_T(s)$ and all positions $i, j$. We are in state $q_i$ at position $i$. For example, \begin{enumerate} \item There is a path from $(Z^{d},1)$ to $(Y^{out},5)$ for $d \in \{in,out\}$. This is because $Z$ at position 1 flows into $Z$ at position 4 (path $(Z^{in},4)$ to $(Z^{in},1)$, edge from and $(Z^{in},1)$ to $(Z^{out},1)$, path from $(Z^{out},1)$ to $(Z^{out},4)$) this value of $Z$ is used in updating $Y$ at position 5 as $Y:=bZc$. (edge from $(Z^{out},4)$ to $(Y^{out},5)$). \item There is a path from $(Y^{in},5)$ to $(Z^{d},2)$. This is because $Z$ at position 2 flows into $Y$ at position 5 by the update $Y:=YbZc$. (path from $(Z^{in},4)$ to $(Z^{in},2)$; path from $(Z^{in},2)$ to $(Z^{out},2)$ and path from $(Z^{out},2)$ to $(Z^{out},4)$; lastly, edge from $(Z^{out},4)$ to $(Y^{out},5)$. Also, note the edge from $(Y^{in},5)$ to $(Y^{in},4)$, and the path from $(Y^{in},4)$ to $(Y^{out},4)$, edge from $(Y^{out},4)$ to $(Z^{in},4)$). \item There is a path from $(X^{in},3)$ to $(Z^{in},1)$ in Figure \ref{fig:outputgraph}. However, $Z$ at position 1 does not flow into $X$ at position 3. Note that this is because of the update $X:=XY$ at position 6, and $Z$ at position 1 flows into $Y$ at position 5, (note the path from $(Z^{out},1)$ to $(Z^{out},4)$, and the edge from $(Z^{out},4)$ to $(Y^{out},5)$) and $X$ at position 3 flows into $X$ at position 5 (note the path from $(X^{out},3)$ to $(X^{out},5)$) and $X$ and $Y$ are catenated in order at position 5 (the edge from $(X^{out},5)$ to $(Y^{in},5)$) to define $X$ at position 6 (edge from $(Y^{out},5)$ to $(X^{out},6)$). \end{enumerate} Thus, the SST output graphs have a nice property, which connects a path from $(X^d,i)$ to $(Y^{d'},j)$ based on the variable flow, and the catenation of variables in updates. Formally, let $T$ be an \textbf{aperiodic,1-bounded} \sst{} $T$. Let $s\in dom(T)$, $G_T(s)$ its \sst-output structure and $r=q_0\dots q_n\dots$ the accepting run of $T$ on $s$. For all variables $X,Y\in \mathcal{X}$, all positions $i,j\in \mathrm{dom}(s)\cup\{0\}$, all $d,d'\in\{in,out\}$, there exists a path from node $(X^{d},i)$ to node $(Y^{d'},j)$ in $G_T(s)$ iff $(X,i)$ and $(Y,j)$ are both useful and one of the following conditions hold: either \begin{enumerate} \item $(q_i,X)\rightsquigarrow^{s[i{+}1{:}j]}_1 (q_j,Y)$ and $d' = out$, or \item $(q_j,Y)\rightsquigarrow^{s[j{+}1{:}i]}_1(q_i,X)$ and $d = in$, or \item there exists $k\geq max(i,j)$ and two variables $X',Y'$ such $(q_i,X)\rightsquigarrow^{s[i{{+}1:}k]}_1 (q_k,X')$, $(q_j,Y)\rightsquigarrow^{s[j{+}1{:}k]}_1 (q_k,Y')$ and $X'$ and $Y'$ are concatenated in this order\footnote{by concatenated we mean that there exists a variable update whose rhs is of the form $\dots X'\dots Y'\dots$} by $r$ when reading $s[k+1]$. \end{enumerate} \subsection{Proof of Lemma \ref{lem:fopath}} \label{app:fopath} Let $s\in dom(T)$, $G_T(s)$ its \sst-output structure and $r=q_0\dots q_n\dots$ the accepting run of $T$ on $s$. For all variables $X,Y\in \mathcal{X}$, all positions $i,j\in \mathrm{dom}(s)\cup\{0\}$, all $d,d'\in\{in,out\}$, there exists a path from node $(X^{d},i)$ to node $(Y^{d'},j)$ in $G_T(s)$ iff $(X,i)$ and $(Y,j)$ are both useful and one of the following conditions hold: either \begin{enumerate} \item $(q_i,X)\rightsquigarrow^{s[i{+}1{:}j]}_1 (q_j,Y)$ and $d' = out$, or \item $(q_j,Y)\rightsquigarrow^{s[j{+}1{:}i]}_1(q_i,X)$ and $d = in$, or \item there exists $k\geq max(i,j)$ and two variables $X',Y'$ such $(q_i,X)\rightsquigarrow^{s[i{{+}1:}k]}_1 (q_k,X')$, $(q_j,Y)\rightsquigarrow^{s[j{+}1{:}k]}_1 (q_k,Y')$ and $X'$ and $Y'$ are concatenated in this order\footnote{by concatenated we mean that there exists a variable update whose rhs is of the form $\dots X'\dots Y'\dots$} by $r$ when reading $s[k+1]$. \end{enumerate} For all variables $X,Y\in \mathcal{X}$, we denote by $Cat_{X,Y}$ the set of pairs $(p,q,a)\in Q^2\times \Sigma$ such that there exists a transition from $p$ to $q$ on $a$ whose variable update concatenates $X$ and $Y$ (in this order). Define a formula for condition $(3)$: $$ \Psi_3^{X,Y}(x,y) \ \equiv \ \exists z\cdot x\preceq z \wedge y\preceq z \wedge \bigvee_{X',Y'\in\mathcal{X}, (p,q,a)\in Cat_{X',Y'}}[ \\ L_a(z)\wedge \phi_{X\rightsquigarrow X'}(x,z) \wedge \phi_{Y\rightsquigarrow Y'}(y,z) \wedge \phi_p(z)\wedge \phi_q(z+1)] $$ Then, formula $\text{path}_{X,Y,d,d'}(x,y)$ is defined by $$ \begin{array}{llllllllll} \text{path}_{X,Y,in,in}(x,y) & \equiv & \phi_{Y\rightsquigarrow X}(y,x) \vee \Psi_3^{X,Y} \\ \text{path}_{X,Y,in,out}(x,y) & \equiv & \phi_{Y\rightsquigarrow X}(y,x) \vee \phi_{X\rightsquigarrow Y}(x,y) \vee \Psi_3^{X,Y} \\ \text{path}_{X,Y,out,in}(x,y) & \equiv & \text{false} \\ \text{path}_{X,Y,out,out}(x,y) & \equiv & \phi_{X\rightsquigarrow Y}(x,y) \vee \Psi_3^{X,Y} \\ \end{array} $$ \begin{enumerate} \item $\text{path}_{X,Y,in,in}(x,y)$ : Recall that a path from $(X^{in},x)$ to $(Y^{in},y)$ always passes through $(X^{out},y)$. Hence, $y \leq x$, (the $X^{in}$ arrows move on the left) and it must be that there is an edge from $(X^{out},y)$ to $(Y^{in},y)$, which happens when $X$ and $Y$ are concatenated in order. This is handled by $\Psi_3^{X,Y}$. The other possibility is when $Y$ occurs in the right side of $X$ at $x$; in this case, we have an edge from $(X^{in},x)$ to some $(Z^{in},x)$ ($Z$ could be $Y$), leading into a path to $(Y^{in},y)$, $y \leq x$. Clearly, here $Y$ flows into $X$ between $y$ and $x$. \item $\text{path}_{X,Y,in,out}(x,y)$ : One possibility is that there is a path from $(X^{in},x)$ to $(X^{in},z)$, with $x > z$, and $(X^{in},z)$ has an edge to $(X^{out},z)$. An edge from $(X^{out},z)$ to $(Y^{in}, z)$ happens if $X, Y$ are concatenated in order at $z$. A path from $(Y^{in}, z)$ to $(Y^{in}, y)$ and then $(Y^{out}, y)$, $z \geq y$ can happen. This is handled by $\Psi_3^{X,Y}$. A second case is similar to 1, where we have a path from $(X^{in},x)$ to $(Y^{in},y)$, where $Y$ flows into $X$ between $y$ and $x$, and then there is a path from $(Y^{in},y)$ to $(Y^{out},y)$. A third case is when $X$ occurs in the right of $Y$ at $y$; in this case, we have an edge from $(X^{out},y)$ to $(Y^{out},y)$. Also, there is a path from $(X^{in},y)$ to $(X^{out},y)$ which goes through $(X^{in},x)$, $x < y$. Clearly, $X$ flows into $Y$ between $x$ and $y$. \item Note that $\text{path}_{X,Y,out,in}(x,y)$ is false, since there is no path or edge from $X^{out}$ to $Y^{in}$ capturing flow; the edge from $X^{out}$ to $Y^{in}$ occurs only when a catenation of $X, Y$ happens in order. \item $\text{path}_{X,Y,out,out}(x,y)$ : One case is when $X,Y$ are catenated in order on the rightside of some variable $Z$. Then there is a path from $(X^{out},x)$ to $(Y^{out},y)$ (through $(Y^{in},y)$), as handled by $\Psi_3^{X,Y}$. The other case is when $X$ occurs on the right of $Y$ at $y$. Then there is an edge from $(X^{out},x)$ to $(Y^{out},y)$ ($x=y-1$) (may be through some $(Z^{in},x)$). \end{enumerate} \input{app-sst-fot.tex} \input{app-2wst-sst.tex} \input{app-sst-la.tex} \section{Proofs from Section \ref{sec:fot-2wst}} \label{app:fot-2wst} \begin{lemma} \label{cor-lem} The proof of Lemma \ref{app:lem1-conv} works when $\mathcal{A}$ is a two-way, aperiodic Muller automaton with star-free look-around. That is, $L(\mathcal{A})$ is aperiodic. \end{lemma} \begin{proof} When we allow star-free look-around, we also have aperiodic Muller look-ahead automaton $A$ and aperiodic look-behind automata $B$ along with $\mathcal{A}$. Consider a string in $w \in \Sigma^{\omega}$ with a factorization $w=w_1w_2w_3 \dots$. Whenever we consider $M_{w_i}$ for some $w_i$ in $\mathcal{T}(\mathcal{A})$ (recall this is the monoid for $\mathcal{A}$ without the look-around), we also consider the transition matrix of $w_1 \dots w_i$ with respect to $B$, and the matrices $M_{w_{j}}, j > i$ with respect to $A$. A string $w'=w'_1w'_2 \dots$ is then equivalent to $w$ if the respective matrices $M_{w_i}, M_{w'_i}$ match in $\mathcal{A}$, and so do the others (prefixes upto $w_i, w'_i$ for look-behind $B$ and $M_{w_j}, M_{w'_j}$ for look-ahead $A$). We know that the transition monoids of $A, B$ are aperiodic. That is, there exists $m_A, m_B \in \mathbb{N}$ such that for all strings $y$, and all pairs of states $p,q$ in ($A$ or $B$), $y^{m_x} \in L_{pq}$ iff $y^{m_x+1} \in L_{pq}$ for $x \in \{A,B\}$. In the presence of look-around, we keep track of the transition monoids in $A, \mathcal{A}$ and $B$ at the same time. It can be seen that for two infinite strings $w,w'$ as above, the map $h$ from $\Sigma^*$ to the transition monoids with respect to $B, \mathcal{A}, A$ will be a morphism : this is seen by considering the respective morphisms $h_1, h_2, h_3$ where $h_1$ is the morphism from $\Sigma^*$ to the transition monoid of $B$ (a DFA), $h_2$ is the morphism from $\Sigma^*$ to the transition monoid of $\mathcal{A}$ (the \ensuremath{\mathsf{2WST}}{}), and $h_3$ is the morphism from $\Sigma^*$ to the transition monoid of $A$ (a DMA). Clearly, equivalent strings $w, w'$ (equivalent with respect to $h$) are either both accepted or rejected by the $\ensuremath{\mathsf{2WST}}_{sf}$. The aperiodicity of the combined monoid follows from the aperiodicity of the respective monoids of $B, \mathcal{A}$ and $A$. Thus, $L(\mathcal{A})$ is recognized by a morphism to an aperiodic monoid, and hence is an aperiodic language. \end{proof} \subsection{Lemma \ref{fo-behav}} \label{app:fo-behav} \begin{proof} Lets look at the underlying two-way Muller automaton of the \ensuremath{\mathsf{2WST}}$_{sf}$ $\mathcal{A}$. Since $\mathcal{A}$ is aperiodic, so is the underlying automaton. Let $s=s[1 \dots x'-1]s[x' \dots y']s[y' \dots]$ be a decomposition of $s$. Let $x,y$ be any two positions in $s$. Depending on $x,y$, we have 4 cases. If $x=y$ then there is a substring $s_1$ of $s$ such that $s_1 \in L^{rr}_{qq'}$ or $s_1 \in L^{ll}_{qq'}$. Similarly, if $x<y$, then there is a substring $s_1$ of $s$ such that $s_1 \in L^{lr}_{qq'}$. Likewise if $x>y$, then there is a substring $s_2$ of $s$ such that $s_2 \in L^{rl}_{qq'}$. If we can characterize $L_{qq'}^{xy}$ in FO for all cases, we are done, since $\psi_{q,q'}(x,y)$ will then be the disjunction of all these formulae. Using lemma \ref{cor-lem}, the underlying input language $L(\mathcal{A})\subseteq \Sigma^{\omega}$ can be shown to be aperiodic, by constructing the morphism $h: \Sigma^* \rightarrow M$ recognizing $L(\mathcal{A})$, where $M$ is the aperiodic monoid described in lemma\ref{cor-lem}. It is known \cite{dg08SIWT} that every aperiodic language $\subseteq \Sigma^{\omega}$ is FO-definable. We now show that $h$ recognizes $L_{qq'}^{xy}$. Consider $w \in L_{qq'}^{xy}$. For the morphism $h: \Sigma^* \rightarrow M$ as above, let $w' \in \Sigma^+$ be such that $h(w)=h(w')$. Then it is easy to see that $w' \in L_{qq'}^{xy}$. Then $h$ is a morphism to a finite aperiodic monoid that recognizes $L_{qq'}^{xy}$. Hence, $L_{qq'}^{xy}$ is aperiodic and hence FO-definable. For each case, $x <y, x > y, x=y$, let $\exists x \exists y \psi_{q,q'}(x,y)$ be the FO formula that captures $L_{qq'}^{xy}$. For a particular assignment of positions $x, y$, it can be seen that all words $u$ which have a run starting at position $x$ in state $q$, to state $q'$ in position $y$ will satisfy $\psi_{q,q'}(x,y)$. \end{proof} \subsection{FOT $\subseteq$ Aperiodic \ensuremath{\mathsf{2WST}}$_{sf}$} \label{app:fo-wst} \begin{definition} A \ensuremath{\mathsf{2WST}}{} with FO instructions ($\ensuremath{\mathsf{2WST}}{}_{fo}$) is a tuple $\mathcal{A}=(\Sigma, \Gamma, Q, q_0, \delta, F)$ such that $\Sigma, \Gamma, Q, q_0$ and $F$ are as defined in section \ref{sec:2wst} and $\delta: Q \times \Sigma \times \phi_1 \rightarrow Q \times \Gamma \times \phi_2$ is the transition function where $\phi_1$ is a set of FO formulae over $\Sigma$ with one free variable defining the guard of the transition, and $\phi_2$ is a set of FO formulae over $\Sigma$ with two free variables defining the jump of the input head. Given a state $q$ and position $x$ of the input string $s$, the transition $\delta(q,\phi_1)=(q',b,\phi_2)$ is enabled if $s \models \phi_1(x)$, and as a result $b$ is written on the output, and the input head moves to position $y$ such that $s \models \phi_2(x,y)$. Note that the jump is deterministic, at each state $q$ and each position $x$, the formula $\phi_2(x,y)$ is such that there is a unique position $y$ to which the reading head will jump. A $\ensuremath{\mathsf{2WST}}{}_{fo}$ is aperiodic if the underlying input language accepted is aperiodic. \end{definition} As in the case of \ensuremath{\mathsf{2WST}}{}, we assume that the entire input is read by the $\ensuremath{\mathsf{2WST}}{}_{fo}$, failing which the output is not defined. \begin{lemma}(\fot{} $\subseteq$ \ensuremath{\mathsf{2WST}}$_{fo}$) \label{foi-2wst} Any $\omega$-transformation captured by an \fot{} is also captured by a \ensuremath{\mathsf{2WST}}$_{fo}$. \end{lemma} \begin{proof} Let $T=(\Sigma, \Gamma, \phi_{dom}, C, \phi_{pos}, \phi_{\prec})$ be a \fot{}. We define the \ensuremath{\mathsf{2WST}}$_{fo}$ $\mathcal{A}=(\Sigma, \Gamma, Q, q_0, \delta, F)$ such that $\inter{T}=\inter{\mathcal{A}}$. The states $Q$ of $\mathcal{A}$ correspond to the copies in $T$. So, $Q=C$. Given a state $q$ and a position $x$ of the input string, the transition that checks a guard at $x$, and decides the jump to position $y$, after writing a symbol $b$ on the output is obtained from the formulae $is\_string$, $\phi_{b}^q(x)$ and $\phi^{q,q'}(x,y)$. $is\_string$ describes the input string, $\phi_{b}^q(x)$ is an FO formula that captures the position $x$ in copy $q$ (and also asserts that the output is $b \in \Gamma$), $\phi^{q,q'}(x,y)$ is an FO formula that enables the $\prec$ relation between positions $x,y$ of copies $q,q'$ respectively. In $\mathcal{A}$, this amounts to evaluating the guard $\phi_b^q(x)$ at position $x$ in state $q$, outputting $b$, and jumping to position $y$ of the input in state $q'$ if the input string satisfies $\phi^{q,q'}(x,y)$. Thus, we write the transition as $\delta(q, a, \phi_b^q(x))=(q',b,\phi^{q,q'}(x,y))$. The initial state $q_0$ of $\mathcal{A}$ is the copy $c$ which has its first position $y$ such that $x \nprec y$ for all positions $y \neq x$. Since the \fot{} is string-to-string, we will have such a unique copy. Thus, $q_0=d$ where $d$ is a copy satisfying the formula $\exists y[first^d(y)]$. The set $F$ of Muller states of the constructed \ensuremath{\mathsf{2WST}}$_{fo}$ is all possible subsets of $Q$, since we have captured the transitions between copies (now states) correctly. Now we have to show that $\mathcal{A}$ is aperiodic. This amounts to showing that there exists some integer $n$ such that for all pairs of states $p,q,$ $v^n \in L_{pq}^{xy}$ iff $v^{n+1} \in L_{pq}^{xy}$ for all strings $v \in \Sigma^*$, and $x,y \in \{l,r\}$. Recall that the states of the automaton correspond to the copies of the \fot{}, and $v^n \in L_{pq}^{xy}$ means that $v^n \models \phi^{p,q}(x,y)$. Thus we have to show that $v^n \models \phi^{p,q}(x,y)$ iff $v^{n+1} \models \phi^{p,q}(x,y)$ for all strings $v \in \Sigma^*$. Note that since the domain of an \fot{} is aperiodic, for any strings $u,v,w$, there exists an integer $n$ such that $uv^nw$ is in the domain of $T$ iff $uv^{n+1}w$ is. In particular, for any pair of positions $x,y$ in $v,w$ respectively, and copies $p, q$, the formula $\varphi^{p,q}(x,y)$ which asserts the existence of a path from position $x$ of copy $p$ to position $y$ of copy $q$ is such that $uv^nw \models \varphi^{p,q}(x,y)$ iff $uv^{n+1}w \models \varphi^{p,q}(x,y)$ (note that this is true since these formulae evaluate in the same way for all strings in the domain, and the domain is aperiodic). This means that $uv^nw \in L_{pq}^{xy}$ iff $uv^{n+1}w \in L_{pq}^{xy}$ for all $u,v,w$. In particular, for $u=w=\epsilon$, we obtain $v^n \in L_{pq}^{xy}$ iff $v^{n+1} \in L_{pq}^{xy}$ for all strings $v \in \Sigma^*$, showing that $\mathcal{A}$ is aperiodic. \end{proof} Before we show \ensuremath{\mathsf{2WST}}$_{fo} \subseteq \ensuremath{\mathsf{2WST}}_{sf}$, we need the next two lemmas. \begin{lemma} \label{lem-int-1} Let $\Delta, \Delta'$ be disjoint subsets of $\Sigma$, and let $L \subseteq \Sigma^{\omega}$ be an aperiodic language such that each string in $L$ contains exactly one occurrence of a symbol from $\Delta$ and one occurrence of a symbol from $\Delta'$. Then $L$ can be written as the finite union of disjoint languages $R_{\ell}.a.R_{m}.b.R_{r}$ where $(a,b) \in (\Delta \times \Delta') \cup (\Delta' \times \Delta)$, $R_{\ell}, R_m \subseteq (\Sigma -\Delta -\Delta')^*$, and $R_r \subseteq (\Sigma -\Delta -\Delta')^{\omega}$. Moreover, $R_{\ell}, R_m$ and $R_r$ are aperiodic. \end{lemma} \begin{proof} Let $\mathcal{A}$ be a deterministic, aperiodic Muller automaton accepting $L$. From our assumption, it follows that each path in $\mathcal{A}$ from the initial state $q_0$ passes through exactly one transition labeled with a symbol from $\Delta$ and one transition labeled with a symbol from $\Delta'$. Let $(q,a,q') \in Q \times \Delta \times Q, (p, b, p') \in Q \times \Delta' \times Q$ be two such transitions. Let $R_{\ell}$ consist of all finite strings from $q_0$ to $q$, and let $R_m$ consist of all finite strings from $q'$ to $p$, and let $R_r$ be the set of all strings from $p'$ which continously witnesses some Muller set from some point onwards. Since the underlying automaton is aperiodic, it is easy to see that the restricted automata for $R_{\ell}, R_m, R_r$ are also aperiodic. Hence, $R_{\ell}, R_m$ are aperiodic languages $\subseteq \Sigma^*$ while $R_r$ is an aperiodic $\omega$-language. This breakup using $R_{\ell}, R_m$ and $R_r$ accounts for strings where the symbol from $\Delta$ occurs first, and the symbol from $\Delta'$ occurs later. Symmetrically, we can define $R'_{\ell}, R'_m$ and $R'_r$ to be the breakup for strings of $L$ where a symbol of $\Delta'$ is seen first, followed by a symbol of $\Delta$. Clearly, $L$ is the finite union of languages $R_{\ell}.a.R_{m}.b.R_{r}$ and $R'_{\ell}.a.R'_{m}.b.R'_{r}$, where $R_{\ell}, R'_{\ell}, R_{m}, R'_{m}, R_{r}$ and $R'_{r}$ are all aperiodic. \end{proof} \begin{lemma} \label{lem-int-2} Let $\Delta \subseteq \Sigma$, and let $L \subseteq \Sigma^{\omega}$ be an aperiodic language such that each string in $L$ contains exactly one occurrence of a symbol from $\Delta$. Then $L$ can be written as the finite union of disjoint languages $R_{\ell}.a.R_{r}$ where $a \in \Delta$, $R_{\ell} \subseteq (\Sigma -\Delta)^*$, and $R_r \subseteq (\Sigma -\Delta)^{\omega}$. Moreover, $R_{\ell}$ and $R_r$ are aperiodic. \end{lemma} The proof of Lemma \ref{lem-int-2} is similar to that of Lemma \ref{lem-int-1}. \begin{lemma}($\ensuremath{\mathsf{2WST}}_{fo} \subseteq \ensuremath{\mathsf{2WST}}_{sf}$) \label{2wstfo-2wst} An $\omega$-transformation captured by an aperiodic $\ensuremath{\mathsf{2WST}}_{fo}$ is also captured by an aperiodic $\ensuremath{\mathsf{2WST}}_{sf}$. \end{lemma} \begin{proof} To prove this, we show that Aperiodic $\ensuremath{\mathsf{2WST}}{}_{fo} \subseteq$ Aperiodic $\ensuremath{\mathsf{2WST}}{}_\ensuremath{sf}$. Let $\mathcal{T}=(\Sigma, \Gamma, Q, q_0, \delta, F)$ be an aperiodic $\ensuremath{\mathsf{2WST}}{}_{fo}$. We define an aperiodic $\ensuremath{\mathsf{2WST}}{}_{sf}$ (with star free look around) $(T, A, B)$ where $T=(\Sigma, \Gamma, Q, q_0, \delta, F)$ capturing the same transformation. A transition of the $\ensuremath{\mathsf{2WST}}{}_{fo}$ is of the form $\delta(q,a, \varphi_1)=(q', z, \varphi_2)$. $\varphi_1(x)$ is an FO formula that acts as the guard of the transition, while $\varphi_2(x,y)$ is an FO formula that deterministically decides the jump to a position $y$ from the current position $x$. The languages of these formulae, $L(\varphi_1)$ and $L(\varphi_2)$ are aperiodic, and one can construct aperiodic Muller automata accepting $L(\varphi_1)$ and $L(\varphi_2)$. From Lemma \ref{lem-int-2}, we can write $L(\varphi_1)$ as a finite union of disjoint languages $R'_{\ell}.(a,1).R'_{r}$ with $R'_{\ell} \subseteq [\Sigma \times \{0\}]^*$ and $R'_{r} \subseteq [\Sigma \times \{0\}]^{\omega}$. It is easy to see that one can construct aperiodic automata accepting $R_{\ell}$ and $R_{m}$, the projections of $R'_{\ell}$ and $R'_{m}$ on $\Sigma$. We have to now show how to simulate the jumps of $\varphi_2$ by moving one cell at a time. First we augment $\varphi_2(x,y)$ in such a way that we know whether $y< x$ or $y >x$ or $y=x$. This is done by rewriting $\varphi_2(x,y)$ as $\exists y (y \sim x\wedge \varphi_2(x,y))$ where $\sim \in \{<, =, >\}$. Clearly, the new formula is also FO, and the language of the formula is aperiodic. If $y < x$, by Lemma \ref{lem-int-1}, the language of $\varphi_2(x,y)$ can be written as $R'_{\ell}(a,1,0) R'_m (b,0,1) R'_r$, where $R'_{\ell}, R'_m \subseteq [\Sigma \times \{0\} \times \{0\}]^*$ and $R'_r \subseteq [\Sigma \times \{0\} \times \{0\}]^{\omega}$. Let $R_{\ell}, R_m$ and $R_r$ be the projections of $R'_{\ell}, R'_m$ and $R'_r$ to $\Sigma$. We can construct an aperiodic look-behind automata for $R_{\ell} a R_m b$ and an aperiodic look ahead Muller automaton for $R_r$. This is possible since $R_{\ell},R_m $ are aperiodic. To walk cell by cell, instead of jumping, the $\ensuremath{\mathsf{2WST}}_{sf}$ does the following. (1) construct the aperiodic automaton that accepts the reverse of $R_m$ (this is possible since the reverse of an aperiodic language is aperiodic), (2) simulate this reverse automaton on each transition, and remember the state reached in this automaton in the finite control, while moving left cell by cell each time, (3) when an accepting state is reached in the reverse automaton, check the look-behind $R_{\ell}$, and look-ahead $R_m b R_r$. Note that there is an aperiodic look-ahead Muller automaton for $R_m b R_r$. If indeed at the position where we are at an accepting state of the reverse of $R_m$ automaton, the look-ahead and look-behind are satisfied, then we can stop moving left. A similar construction works when $y>x$. Clearly, each transition of the aperiodic $\ensuremath{\mathsf{2WST}}_{fo}$ can be simulated by a $\ensuremath{\mathsf{2WST}}_{sf}$ with star-free look around. The aperiodicity of $\ensuremath{\mathsf{2WST}}_{sf}$ follows from the fact the input language accepted by the $\ensuremath{\mathsf{2WST}}_{sf}$ is same as that of the $\ensuremath{\mathsf{2WST}}_{fo}$, and is aperiodic. \end{proof} \subsection{Formal Construction of FOT from Aperiodic SST} \label{app-fot-cons} We describe the construction of the \fot{} from the 1-bounded, aperiodic \sst{}. In the case of finite strings, if the output function of the accepting state is $X_1 \dots X_n$, then $(X_1^{in}, |s|)$ is the start node, on reading string $s$. The path from $(X_1^{in}, |s|)$ to $(X_1^{out}, |s|)$, followed by the path from $(X_2^{in}, |s|)$ to $(X_2^{out}, |s|)$ and so on gives the output. Unlike the finite string case, one of the difficulties here, is to specify the start node of the \fot{} since the strings are infinite. The first thing we do, in order to specify the start node, is to identify the position where some Muller set starts, and the rest of the run stays in that set. This is done for instance, in Proposition \ref{prop:fostates} by the formula $\psi^{Rec}_P$. We can easily catch the first such position where $\psi^{Rec}_P$ holds. This position will be labeled in the \fot{} with a unique symbol $\perp$. Let $O \subseteq \Gamma^*$ be the finite set of output strings $\gamma_i$ that appear in the variable updates of the \sst{}. That is, $O=\{\gamma_i \mid \rho(q,a)(X)=\gamma_0X_1 \gamma_1 \dots \gamma_{n-1} X_n \gamma_n\}$. We build an \fot{} that marks the first position where $\psi^{Rec}_P$ evaluates to true as as $\perp$, outputs the contents of $X_1, \dots, X_{n-1}$ first till $\perp$, and then of $X_n$, where $F(P)=X_1\dots X_{n-1}X_n$ is the output function of the \sst{}. The \fot{} is defined as $(\Sigma, O \cup \{\perp\}, \phi_{dom}, C, \phi_{pos}, \phi_{\prec})$ where: \begin{itemize} \item $\phi_{dom}=is\_string \wedge \exists i\psi^{Rec}_P(i)$ where the FO formula $is\_string$ is a simple FO formula that says every position has a unique successor and predecessor. The conjunction with $\exists i\psi^{Rec}_P(i)$ says that there is a position $i$ of the string from where $\psi^{Rec}_P$ is true. \item $C=\mathcal{X} \times \{in, out\}$, \item $\phi_{pos}=\{\phi_{\gamma}^c(i) \mid c \in C, \gamma \in O \cup \{\perp\}\}$ is such that \begin{itemize} \item $\phi_{\perp}^{(X_1,in)}(i)=\bigvee_{\{P \in F \mid F(P)=X_1 \dots X_n\}}[\psi^{Rec}_P(i) \wedge \forall j(j \prec i \rightarrow \neg\psi^{Rec}_P(j))]$. $i$ is the first position from where a Muller set starts to hold continuosuly. \item for $\gamma \in O$, we define $\phi_{\gamma}^c(i)=\neg \phi_{\perp}^c(i) \wedge \psi_{\gamma}^c(i)$, where $\psi_{\gamma}^c(i)$ is \end{itemize} \begin{itemize} \item true, if $(i=0$), $c=(X,in)$ and $\gamma=\epsilon$, \item false, if $(i=0$), $c=(X,in)$ and $\gamma \neq \epsilon$, \item $\bigvee_{q \in Q, a \in \Sigma}\phi_q(i-1) \wedge L_a(i-1)$ if $c=(X,in)$, $i > 0$ and $\rho(q,a)(X)=\gamma Y_1\gamma_1 Y_2 \dots \gamma_n$. The formula $\phi_q(x)$ is defined in Proposition \ref{prop:fostates}. \item $\bigvee_{q \in Q, a \in \Sigma}\phi_q(i) \wedge L_a(i)$ if $c=(X,out)$, $i \geq 0$ and $\rho(q,a)(Y)=\gamma_0 Y_1\gamma_1 Y_2 \dots Y_k\gamma_k \dots \gamma_n$, for a unique $Y \in \mathcal{X}$, and $Y_k=X$ and $\gamma=\gamma_k$ for some $1 \leq k \leq n$. Note that $Y \in \mathcal{X}$ is unique since the \sst{} is 1-bounded (the \sst{} output-structure does not contain useless variables, and if $X$ appears in the rightside of two variables $Y,Z$ then one of them will be useless for the output) \item $\bigvee_{q \in Q, a \in \Sigma}\phi_q(i) \wedge L_a(i)$ if $c=(X,out)$, $i \geq 0$ and $\gamma=\epsilon$ if $X$ does not appear in the update of any variable in $\rho(q,a)$. \end{itemize} \item We next define $\phi^{c,d}(i,j)=\bigvee_{P \in F}[\psi^{Rec}_P \wedge \psi^{c,d}_P(i,j)] $, where $\psi^{c,d}_P(i,j)$ is defined as follows. Let $F(P)=X_1X_2 \dots X_{n-1}X_n$. Let $k$ be the earliest position where $\psi^{Rec}_P$ holds. Note that all these edges are defined only when a copy is useful, that is, it contributes to the output. The notion of usefulness has been defined in section \ref{sec:sst-fot}. \begin{itemize} \item $true \wedge useful_X(i)$, if $c=(X,in)$, $d=(X,out)$ for some variable $X$, and $i=j=0$. The formula $useful_X(i)$ is defined in Appendix \ref{app:useful}. \item $\phi_q(i-1) \wedge L_a(i-1) \wedge useful_X(i)$ if $c=(X,in), d=(X,out)$ $i=j>0$ and $\rho(q,a)(X)=\gamma$. \item $\phi_q(i) \wedge L_a(i) \wedge useful_X(i)$ if $c=(X,out)$, $d=(Y,in)$ for some variables $X,Y$, $i=j\geq 0$, and for some $Z$ we have $\rho(q,a)(Z)=\gamma_0 Z_1 \gamma_1 \dots Z_n \gamma_n$ with $Z_{\ell}=X$ and $Z_{\ell+1}=Y$ for some $1 \leq \ell \leq n$ \item $\phi_q(i-1) \wedge L_a(i-1) \wedge useful_X(i)$ if $C=(X,in)$, $d=(Y, in)$ for some variables $X,Y$, $i >0$, $j=i-1$, $\rho(q,a)(X)=\gamma Y \gamma_1 Y_2 \dots Y_n \gamma_n$ \item $\phi_q(i) \wedge L_a(i) \wedge useful_X(i)$ if $c=(X,out)$, $d=(Y,out)$ for some variables $X,Y$, $i+1=j$ and $\rho(q,a)(Y)=\gamma X_1 \dots X \gamma_n$. \item We do the above between copies upto position $k$, which influences the values of variables $X_1, \dots, X_{n-1}$, which produces the variable flow upto position $k$. Beyond position $k$, the contents of variables $X_1, \dots, X_{n-1}$ remain unchanged. Hence, for all $i \geq k$, and $1 \leq j \leq n-1$, we dont have any edges from $(X_j,out)$ at position $i$ to $(X_j,out)$ at position $i+1$, and from $(X_j,in)$ at position $i+1$ to $(X_j,in)$ at position $i$. We simply add a transition from $(X_j,out)$ to $(X_{j+1},in)$ at position $k$ to seamlessly catenate the contents of $X_1, \dots, X_{n-1}, X_n$ at position $k$. \item From $(X_n,in)$ at position $k$ onwards, we simply follow the edges as defined above, to obtain at each position, the correct output $X_1X_2 \dots X_n$. If at position $k$, $X_n:=X_n \gamma_0 Y_1 \dots \gamma_{n-1} Y_n \gamma_n$ then we have the connections from $(X_n, in)$ at position $k$ to $(X_n, in)$ at position $k-1$, which will eventually reach $(X_n, out)$ at position $k-1$. This is then connected to $(Y_1, in)$ at position $k-1$, and so on till we reach $(Y_n,out)$ at position $k-1$. This is then connected to $(X_n, out)$ at position $k$, rendering the correct output at position $k$. The same thing repeats for position $k+1$ and so on. \end{itemize} \end{itemize} Thanks to Lemma \ref{lem:fopath}, the transitive closure between some copy $(X,d)$ and $(Y,d')$ is FO-definable as $\phi^{(X,d),(Y,d')}_{\prec}=path_{X,Y,d,d'}(x,y)$. This completes the construction of the \fot{}. \section{Proofs from Section \ref{sec:2wst-sst} : $\sst_{\ensuremath{sf}} \subseteq \sst{}$} \label{app:starfree} We now show that we can eliminate the star-free look-around from the $\sst_{sf}$ $(T,A,B)$ without losing expressiveness. Eliminating the look behind is easy: $B$ can be simulated by computing for each state $p_B \in P_B$, the state $p_B'$ of $P_B$ reached by $B$ starting in $p_B$ on the current prefix, and whenever $p_B'$ is a final state of $B$, the transition is triggered. In order to remove the look-ahead, we need to keep track of $P_i \in2^{P_A}$ at every step. On processing a string $s=a_1 a_2 a_3 \ldots \in \Sigma^{\omega}$ in $T$ starting with $P_0=\emptyset$, we obtain successively $P_1, P_2, \dots$ where $P_{i+1}= P_{i+1}' \cup \{p_{i+1}\}$ such that $\delta(q_i, r_{i+1}, a_{i+1}, p_{i+1}) = q_{i+1}$, and for all $p\in P_i$, $\delta_A(p,a_{i+1})\in P_{i+1}'$. Thus, starting with $P_0=\emptyset$ and $\delta(q_0, r_{1}, a_{1}, p_{1}) = q_{1}$, we have $P_1=\{p_1\}$, $P_2=\delta_A(p_1, a_2) \cup \{p_2\}$ and so on. A configuration of the \sst$_{sf}$ is thus a tuple $(q, (r'_1, \dots, r'_n), 2^{P_A})$ where $r'_i$ is the state reached in $B$ on reading the current prefix from state $r_i$, assuming $P_B=\{r_1, \dots, r_n\}$, and $2^{P_A}$ is a set of states in $P_A$ obtained as explained above. We say that $\rho$ is an accepting run of $s$ if $(q_0,(r_1, \dots, r_n), P_0)$ is an initial configuration, i.e. $q_0\in Q_0, P_0=\emptyset$ and after some point, we only see all elements of exactly one Muller set $M_i$ repeating infinitely often in $Q$ as well as in $P_A$ i.e. $\Omega(\rho)_{1} = M_i$ from domain of $F$ and $\Omega(\rho)_2 = M_j$ from $P_f$. Also, $(q_i, (r'_1, \dots, r'_n), P_i) \stackrel{r_{k}, a, p_{i+1}}{\longrightarrow} (q'_i, (r''_1, \dots, r''_n), P_{i+1})$ iff $\delta(q_i,a,r_k,p_{i+1})=q'_i$, $P_{i+1}= \delta^*_A(P_{i},a) \cup \{p_{i+1}\}$ and $\delta_B(r'_j,a)=r''_j$ for $1 \leq j \leq n$ and the state $r''_k$ is an accepting state of $B$. A configuration in the \sst$_{sf}$ is said to be \emph{accessible} if it can be reached from an initial configuration, and \emph{co-accessible} if from it accepting configurations can be reached. It is \emph{useful} if it is both accessible and co-accessible. Note that from the mutual-exclusiveness of look-arounds and the determinism of $A,B$, it follows that for any input string, there is at most one run of the $\mathsf{SST}_{\ensuremath{sf}}$ from and to useful configurations, as shown in Appendix \ref{app:unique}. The concept of substitutions induced by a run can be naturally extended from $\mathsf{SST}$ to $\mathsf{SST}_{\ensuremath{sf}}$. Also, we can define the transformation implemented by an $\mathsf{SST}_{\ensuremath{sf}}$ in a straightforward manner. The transition monoid of an $\mathsf{SST}_{\ensuremath{sf}}$ is defined by matrices indexed by configurations $(q_i, (r_1, \dots, r_n), P_i)\in Q\times P_B^n \times 2^{P_A}$, using the notion of run defined before, and the definition of aperiodicity of $\mathsf{SST}_{\ensuremath{sf}}$ follows that of $\mathsf{SST}$. \subsection{Uniqueness of Accepting Runs in $\mathsf{SST}_{\ensuremath{sf}}$} \label{app:unique} Let $a_1a_2\dots \in \Sigma^{\omega}$ and $\rho: (q_0, (r_1, \dots, r_n), P_0) \stackrel{a_1}{\rightarrow}(q_1, (r_1^1, \dots, r_n^1), P_1)\stackrel{a_2}{\rightarrow}(q_2, (r_1^2, \dots, r_n^2), P_2) \dots$ be an accepting run in the $\mathsf{SST}_{\ensuremath{sf}}$. We show that $\rho$ as well as the sequences of transitions associated with $\rho$ are unique. Given a sequence of transitions of the $\mathsf{SST}_{\ensuremath{sf}}$, it is clear that there is exactly one run since both $A, B$ are deterministic. Lets assume that the sequence of transitions are not unique, that is there is another accepting run $\rho'$ for $a_1a_2 \dots$. Let $i$ be the smallest index where $\rho$ and $\rho'$ differ. The $(i-1)$th configuration is then some $(q_i, (r'_1, \dots, r'_n), P_i)$ in both $\rho, \rho'$. Let us assume that we have two transitions $\delta(q_i, r_j, a_i, p_i)=q_{i+1}$ and $\delta(q_i, r_k, a_i, p'_i)=q'_{i+1}$ enabled such that $r_j \neq r_k$ or $p_i \neq p'_i$ or $q_{i+1} \neq q'_{i+1}$. Assume $r_j \neq r_k$. Since both are trigerred, we have the prefix upto now is in $L(B_{r_j}) \cap L(B_{r_k})$ with $r_j \neq r_k$, which contradicts mutual exclusiveness of look-behind. If $p_i \neq p'_i$, then since both runs are accepting, we have the infinite suffix in $L(A_{p_i}) \cap L(A_{p'_i})$, which contradicts the mutual exclusiveness of look-ahead. If $r_j=r_k$ and $p_i=p'_i$, but $q_{i+1} \neq q'_{i+1}$, then $\delta$ is not a function, which is again a contradiction. \subsection*{Variable Flow and Transition Monoid of $\mathsf{SST}_{\ensuremath{sf}}$} \textbf{Variable Flow and Transition Monoid}. Let $P_A, P_B$ represent the states of the (deterministic) lookahead and look-behind automaton $A,B$, and $Q$ denote states of the $\mathsf{SST}_{\ensuremath{sf}}$. The transition monoid of an $\mathsf{SST}_{\ensuremath{sf}}$ depends on its configurations and variables. It extends the notion of transition monoid for $\mathsf{SST}$ with look-behind, look-ahead states components but is defined only on {\it useful} configurations $(q,(r'_1, \dots, r'_n), P)$. A configuration $(q,(r'_1, \dots, r'_n), P)$ is {\it useful} iff it is {\it accessible} and {\it co-accessible} : that is, $(q,(r'_1, \dots, r'_n), P)$ is reachable from the initial configuration $(q_0,(r_1, \dots, r_n),\varnothing)$ (here, $P_B=\{r_1, \dots, r_n\}$) and will reach a configuration $(q',(r''_1, \dots, r''_n),P')$ from where on, some muller subset of $Q$ is witnessed continuously, and some muller subset of $P_A$ is witnessed continuously. Note that given two useful configurations $(q,(r'_1, \dots, r'_n),P)$, $(q',(r''_1, \dots, r''_n), P')$ and a string $s\in\Sigma^*$, there exists \emph{at most} one run from $(q,(r'_1, \dots, r'_n),P)$ to $(q',(r''_1, \dots, r''_n), P')$ on $s$. Indeed, since $(q,(r'_1, \dots, r'_n), P)$ and $(q',(r''_1, \dots, r''_n), P')$ are both useful, there exists $s_1,s_2\in\Sigma^*$ such that $(q_0, (r_1, \dots, r_n),\varnothing)\rightsquigarrow^{s_1} (q,(r'_1, \dots, r'_n),P)$ and $(q',(r''_1, \dots, r''_n),P') \rightsquigarrow^{s_2} (q'', (r'''_1, \dots, r'''_n), P'')$ such that from $(q'', (r'''_1, \dots, r'''_n), P'')$, we settle in some Muller set of both $Q$ and $P_A$ reading some $w \in \Sigma^{\omega}$. If there are two runs from $(q,(r'_1, \dots, r'_n),P)$ to $(q',(r''_1, \dots, r''_n),P')$ on $s$, then there are two accepting runs for $s_1ss_2w$, which contradicts the fact that accepting runs are unique. We denote by $\textsf{useful}(T,A,B)$ the useful configurations of $(T,A,B)$. Thanks to the uniqueness of the sequence of transitions associated with the run of an $\mathsf{SST}_{\ensuremath{sf}}$ from and to useful configurations on a given string, one can extend the notion of variable flow naturally by considering, as for $\mathsf{SST}$, the composition of the variable updates along the run. Assume that $T$ has $j$ muller sets and $B$ has $k$ muller sets. A string $s\in \Sigma^*$ maps to a square matrix $M_s$ of dimension $|Q\times (P_B \times \dots \times P_B) \times 2^{P_A}|\cdot |\mathcal{X}|$ and is defined as \begin{itemize} \item $M_{s}[(q, (r'_1, \dots, r'_n),P), X][(q', (r''_1, \dots, r''_n), P'), X']=n, \alpha_1, \alpha_2$ if there exists a run $\rho$ from \\ $(q, (r'_1, \dots, r'_n), P)$ to $(q', (r''_1, \dots, r''_n), P')$ on $s$ such that $n$ copies of $X$ flows to $X'$ over the run $\rho$, and $(q, (r_1, \dots, r_n), P)$ and $(q', (r''_1, \dots, r''_n), P')$ are both useful (which implies that the sequence of transitions of $\rho$ from $(q, (r_1, \dots, r_n), P)$ to $(q', (r''_1, \dots, r''_n), P')$ is unique, as seen before), and \item $\alpha_1 \in \{0,1,\kappa\}^j$ and $\alpha_2 \in 2^{[\{0,1,\kappa\}^k]}$. $\alpha_1$ is a $j$-tuple keeping track for each of the $j$ muller sets of $T$, whether the set of states seen on reading a string $s$ is a muller set, a strict subset of it, or has seen a state outside of the muller set. Likewise, $\alpha_2$ is a set of $k$-tuples doing the same thing based on the transition from set $P$ to set $P'$. Note that since we keep a subset of states of $P_A$ in the configuration, we need to keep track of this information for each state in the set. Thus, $\alpha_1$ keeps track of the run from $q$ and whether it witnesses a full muller set (1), a partial muller set ($\kappa$), or goes outside of a muller set (0). Likewise, $\alpha_2$ keeps track of the same for each $p \in P_i$, the run from $p$. If a run is accepting, then $\alpha_2$ will eventually become the singleton $\{(0, \dots, 0, 1, 0, \dots, 0)\}$ corresponding to some muller set of $P_A$. \item $M_{s}[(q, (r'_1, \dots, r'_n),P), X][(q',(r''_1, \dots, r''_n),P'), X']=\bot$, otherwise. \end{itemize} \begin{lemma}\label{lem:aperiodicSSTLA} For all aperiodic 1-bounded $\mathsf{SST}_{\ensuremath{sf}}$ with star-free look-around, there exists an equivalent aperiodic 1-bounded SST. \end{lemma} \begin{proof} Let $(T,A,B)$ be an $\mathsf{SST}_{\ensuremath{sf}}$, with $A = (P_A, \Sigma, \delta_A, P_f)$ a deterministic lookahead muller automaton, $B=(P_B, \Sigma, \delta_B)$ be a deterministic look-behind automaton. Let $T = (\Sigma, \Gamma, Q, q_0,\delta, \mathcal{X}, \rho, F)$. Without loss of generality, we make the following \emph{unique successor} assumption \begin{itemize} \item For all states $q,q',q''\in Q$, and for all states $p, p' \in P_A$, and for all states $r, r' \in P_B$, and for any symbol $a \in \Sigma$, whenever $p \neq p'$ or $r \neq r'$, and if $\delta(q,r,a,p)=q'$, $\delta(q,r',a,p')=q''$, then $q'\neq q''$. \end{itemize} If this is not the case, say we have $p \neq p', r=r'$, and $q'=q''$. Then considering the state of $T$ as $(q,r)$, we obtain $\delta((q,r),a,p)=(q',r)=\delta((q,r),a,p')$. However, it is easy to define transitions $\delta((q,r),a,p)=(q',r)$, $\delta((q,r),a,p')=(q'',r)$ by duplicating the transitions of $q', q''$, without affecting anything else, especially the aperiodicity. The same argument can be given when $r \neq r'$ or both ($p \neq p'$ and $r \neq r'$). Thus, for our convenience, we assume the \emph{unique successor} assumption without loss of generality. \subsection{Elimination of look-around from $(T,A,B)$ : Construction of $T'$} We construct an aperiodic and 1-bounded \sst{} $T'$ equivalent to \sst$_{sf}$ $(T,A,B)$. As explained in definition of $\mathsf{SST}_{\ensuremath{sf}}$, the unique run of a string $s$ on $(T,A,B)$ is not only a sequence of $Q$-states, but also a collection of the look ahead states $2^{P_A}$, and maintains the reachable state from each state of $P_B$. At any time, the current state of $Q$, the $n$-tuple of reachable states of $P_B$, and the collection of look-ahead states $P \subseteq P_A$ is a configuration. For brevity of notation, let $\eta$ (we also use $\zeta, \eta'$ in the sequel) denote the $n$-tuple of $P_B$ states. A configuration $(q_1,\eta_1,P_1)$, on reading $a$, evolves into $(q_2,\eta_2,P_2 \cup\{p_2\})$, where $\delta(q_1,r_2,a,p_2)=q_2$ is a transition in the $\mathsf{SST}_{\ensuremath{sf}}$ and $\delta_A(P_1,a)=P_2$, where $\delta_A$ is the transition function of the look ahead automaton $A$, and the state reached from $r_2$ is an accepting state of $P_B$. Note that the transition monoid of the $\mathsf{SST}_{\ensuremath{sf}}$ is aperiodic and 1-bounded by assumption. We now show how to remove the look-around, resulting in an equivalent $\sst{}$ $T'$ whose transition monoid is aperiodic and 1-bounded. While defining $T'$, we put together all the states resulting from transitions of the form $(q,r,a,p,q')$ and $(q,r',a,p',q'')$ in the $\mathsf{SST}_{\ensuremath{sf}}$. We define $T'=(\Sigma, \Gamma,Q',q'_0, \delta', \mathcal{X}', \rho',Q_f')$ with: \begin{itemize} \item $Q' = 2^{\textsf{useful}(T,A,B)}$ where $\textsf{useful}(T,A,B)$ are the useful configurations of $(T,A,B)$ ;\\ (Note that we can pre-compute $\textsf{useful}(T,A,B)$ once we know $(T,A,B)$), \item $q'_0=\{(q_0,\eta,\emptyset)\}$ ($\eta=(r_1, \dots, r_n)$ where $P_B=\{r_1, \dots, r_n\}$. We assume wlog that, $(T,A,B)$ accepts at least one input. Therefore $(q_0,\eta,\emptyset)$ is useful), \item $Q'_f$, the set of muller sets of $T'$, is defined by the set of pairs $(S,R)$ where $S$ is any of the $j$ muller sets in $F$, and $R$ is any of the $k$ muller sets in $P_f$. \item $\mathcal{X}'=\{X_{q'} \mid X \in \mathcal{X}, q' \in \textsf{useful}(T,A,B)\}$, \item The transitions are defined as follows: $\delta'(S,a)=\bigcup_{(q,\eta,P)\in S}\Delta((q,\eta,P),a)$ where \\ $\Delta((q,\eta,P),a)=\{(q', \eta', P' \cup\{p'\}) \mid \delta(q,r,a,p') =q'$ and $\delta_A(P,a)=P'\}\cap \text{useful}(T,A,B)$. Each component of $\eta'$ is obtained from $\eta$ using $\delta_B$. \end{itemize} Before defining the update function, we first assume a total ordering $\preceq_{\text{useful}(T,A,B)}$ on $\text{useful}(T,A,B)$. For all $(p,\eta,P)$, we define the substitution $\sigma_{(p,\eta,P)}$ as $X \in \mathcal{X} \mapsto X_{(p,\eta, P)}$. Let $(S,a,S')$ be a transition of $T'$. Given a state $(q',\eta', P') \in S'$, there might be several predecessor states $(q_1, \eta_1, P_1),\dots,(q_k,\eta_k, P_k)$ in $S$ on reading $a$. The set $\{(q_1,\eta_1, P_1),\dots,(q_k, \eta_k, P_k)\} \subseteq S$ is denoted by $Pre_S((q',\eta',P'),a)$. Formally, it is defined by $\{ (q,\eta,P)\in S\ |\ (q',\eta',P')\in \Delta((q,\eta,P),a)\}$. We consider only the variable update of the transition from the minimal predecessor state. Indeed, since any string has at most one accepting run in the $\mathsf{SST}_{\ensuremath{sf}}$ $(T,A,B)$ (and at most one associated sequence of transitions), if two runs reach the same state at some point, they will anyway define the same output and therefore we can drop one of the variable update, as shown in \cite{FiliotTrivediLics12}. Formally, the variable update $\rho'(S,a,S')(X_{(q',\eta',P')})$, for all $X_{(q',\eta',P')}\in\mathcal{X}'$ is defined by $\epsilon$ if $(q',\eta',P') \notin S'$, and by $\sigma_{(q,\eta,P)}\circ \rho(q,r,a,p,q')(X)$, where $(q,\eta,P) = \text{min}\ \{ (t,\zeta,R)\in S\ |\ (q',\eta',P')\in \Delta((t,\zeta,R),a)\}$, and $\delta(q,r,a,p) = q'$ (by the \emph{unique successor} assumption the look-around states $p,r$ are unique). It is shown in \cite{FiliotTrivediLics12} that indeed $T'$ is equivalent to $T$. We show here that the transition monoid of $T'$ is aperiodic and $1$-bounded. For all $S\in Q'$, and $s \in \Sigma^*$, define $\Delta^*(S,s)=\{ (q',\eta',P') \mid \exists (q,\eta,P) \in S$ such that $(q,\eta,P) \rightsquigarrow^s_{T,A,B} (q',\eta',P')\}\cap\text{useful}(T,A,B)$. \subsection{Connecting Transition Monoids of $T'$ and $(T,A,B)$} \begin{lemma} \label{lem:int} Let $M_{T'}$ be the transition monoid of $T'$ and $M_{(T,A,B)}$ the transition monoid of $(T,A,B)$. Let $S_1,S_2\in Q'$, $X_{(q,\eta,P)},Y_{(q',\eta',P')}\in \mathcal{X}'$ and $s\in\Sigma^*$. \\ Then $M_{T',s}[S_1,X_{(q,\eta,P)}][S_2,Y_{(q',\eta',P')}]=i, \alpha_1, \alpha_2$ iff $S_2 = \Delta^*(S_1,s)$ and one of the following hold: \begin{enumerate} \item either $i=0$ and, $(q,\eta,P)\not\in S_1$ or $(q',\eta',P')\not\in S_2$, or \item $(q,\eta,P) \in S_1$, $(q',\eta',P')\in S_2$, $(q,\eta,P)$ is the minimal ancestor in $S_1$ of $(q',\eta',P')$ (i.e. $(q,\eta,P) = \text{min}\ \{ (t,\zeta,R)\in S_1\ |\ (q',\eta',P')\in \Delta^*((t,\zeta,R),s)\}$), and\\ $M_{(T,A,B),s}[(q,\eta,P),X][(q',\eta',P'),Y] = i, \alpha_1, \alpha_2$. \end{enumerate} \end{lemma} \begin{proof} It is easily shown that $M_{T',s}[S_1,X_{(q,\eta,P)}][S_2,Y_{(q',\eta',P')}]=i, \alpha_1, \alpha_2$ with $i \geq 0$ iff $S_2 = \Delta^*(S_1,s)$. Let us show the two other conditions. Assume that $M_{T',s}[S_1,X_{(q,\eta,P)}][S_2,Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$. The variable update function is defined in $T'$ as follows: after reading string $s$, from state $S_1$, all the variables $Z_{(t,\zeta,R)}$ such that $(t,\zeta,R)\not\in S_2$ are reset to $\epsilon$ (and therefore no variable can flow from $S_1$ to them). In particular, if $(q',\eta', P')\not\in S_2$, then no variable can flow in $Y_{(q',\eta',P')}$ and $i=0$. Now, assume that $(q',\eta',P')\in S_2$, and consider the sequence of states $S_1,S'_1,S'_2,\dots,S'_k,S_2$ of $T'$ on reading $s$. By definition of the variable update, the variables that are used to update $Y_{(q',\eta',P')}$ on reading the last symbol of $s$ from $S'_k$ are copies of the form $Z_{(t,\zeta,R)}$ such that $(t,\zeta,R)$ is the minimal predecessor in $S'_k$ of $(q',\eta',P')$ (by $\Delta$). By induction, it is easily shown that if some variable $Z_{(t,\zeta,R)}$ flows to $Y_{(q',\eta',P')}$ from $S_1$ to $S_2$ on reading $s$, then $(t,\zeta,R)$ is necessarily the minimal ancestor (by $\Delta^*$) of $(q',\eta',P')$ on reading $s$. In particular if $(q,\eta,P)\not\in S_1$, then $i=0$. Finally, if $i>0$, then necessarily $(q,\eta,P)$ is the minimal ancestor in $S_1$ of $(q',\eta',P')$ on reading $s$, from $S_1$ to $S_2$, and since $T'$ mimics the variable update of $(T,A,B)$ on the copies, we get that $M_{(T,A,B),s}[(q,\eta,P),X][(q',\eta',P'),Y]=i, \alpha_1, \alpha_2$. The converse is shown similarly. \end{proof} \subsection{1-boundedness and aperiodicity of $T'$} 1-boundedness is an obvious consequence of the claim and the fact that $(T,A,B)$ is $1$-bounded. Let us show that $M_{T'}$ is aperiodic. We know that $M_{(T,A,B)}$ is aperiodic. Therefore there exists $n\in\mathbb{N}$ such that for all strings $s\in\Sigma^*$, $M_{(T,A,B),s}^n = M_{(T,A,B),s}^{n+1}$. We first show that $\forall$ $S_1,S_2\in Q'$, and all strings $s\in\Sigma^*$, $\Delta^*(S_1,s^n) = S_2$ iff $\Delta^*(S_1,s^{n+1}) = S_2$. Indeed, \begin{itemize} \item $S_2 = \Delta^*(S_1,s^n)$, iff $$S_2 = \{ (q',\eta', P')\in\textsf{useful}(T,A,B)\ |\ \exists (q,\eta,P)\in S_1,\ (q,\eta,P)\rightsquigarrow^{s^n}_{T,A,B},(q',\eta',P')\},$$ iff \item $S_2 = \{ (q',\eta',P')\in\textsf{useful}(T,A,B)\ |\ \exists (q,\eta,P)\in S_1,\ M_{(T,A,B),s}^n[(q,\eta,P),X][(q',\eta',P'),Y]=i, \alpha_1, \alpha_2$, $i \geq 0\text{ for some }X,Y\in\mathcal{X}\}$, iff \item by aperiodicity of $M_{(T,A,B)}$, $$S_2 = \{ (q',\eta',P')\in\textsf{useful}(T,A,B)\ |\ \exists (q,\eta,P)\in S_1,$$ $$~~~~~~M_{(T,A,B),s}^{n+1}[(q,\eta,P),X][(q',\eta',P'),Y]=i, \alpha_1, \alpha_2, i \geq 0 \\ \text{ for some }X,Y\in\mathcal{X}\}$$ iff \item $S_2 = \Delta^*(S_1,s^{n+1})$. \end{itemize} Let $S_1,S_2\in Q'$ and $X_{(q,\eta,P)}, Y_{(q',\eta',P')}\in \mathcal{X}$. Let also $s\in\Sigma^*$. We now show that the matrices corresponding to $s^n, s^{n+1}$ coincide, using Lemma \ref{lem:int}. By the first condition of Lemma \ref{lem:int}, we have \begin{quote} $M_{T',s}^n[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$ and condition $(1)$ of Lemma \ref{lem:int} holds, iff \\ $M_{T',s}^{n+1}[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$ and condition $(1)$ of Lemma \ref{lem:int} holds. \end{quote} \begin{itemize} \item Indeed, $M_{T',s}^n[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = 0, \alpha_1, \alpha_2$ and, $(q,\eta,P)\not\in S_1$ or $(q',\eta',P')\not\in S_2$ iff by Lemma \ref{lem:int}, $\Delta^*(S_1,s^n) = S_2$, and $(q,\eta,P)\not\in S_1$ or $(q',\eta',P')\not\in S_2$, iff \\ by what we just showed, $\Delta^*(S_1,s^{n+1}) = S_2$, and $(q,\eta,P)\not\in S_1$ or $(q',\eta',P')\not\in S_2$, iff by Lemma \ref{lem:int}, $M_{T',s}^{n+1}[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = 0, \alpha_1, \alpha_2$ and condition $(1)$ of Lemma \ref{lem:int} holds. \end{itemize} Let us now look at condition $(2)$ of Lemma \ref{lem:int}, and show that \begin{quote} $M_{T',s}^{n}[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$ and condition $(2)$ of Lemma \ref{lem:int} holds, iff \\ $M_{T',s}^{n+1}[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$ and condition $(2)$ of Lemma \ref{lem:int} holds. \end{quote} \begin{itemize} \item We only show one direction, the other being proved exactly similarly. Assume that \\ $M_{T',s}^{n}[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$ and $(q,\eta,P)\in S_1$, $(q',\eta',P')\in S_2$, and $(q,\eta,P)$ is the minimal ancestor in $S_1$ of $(q',\eta',P')$, and $M_{(T,A,B),s}^n[(q,\eta,P),X][(q',\eta',P'),Y] = i, \alpha_1, \alpha_2$. By Lemma \ref{lem:int}, we have $\Delta^*(S_1,s^n) = S_2$, and therefore $\Delta^*(S_1,s^{n+1}) = S_2$. Now, we have $(q,\eta,P) = \text{min}\ \{ (t,\zeta,R)\in S_1\ |\ (q',\eta',P')\in \Delta^*((t,\zeta, R),s^n)\}$. Since $\Delta^*((t,\zeta,R),s^n) = \Delta^*((t,\zeta,R),s^{n+1})$ for all $(t,\zeta,R)\in S_1$, we have $(q,\eta,P) = \text{min}\ \{ (t,\zeta,R)\in S_1\ |\ (q',\eta',P')\in \Delta^*((t,\zeta,R),s^{n+1})\}$. Finally, $M_{(T,A,B),s}^{n+1}[q,\eta,P,X][q',\eta',P',Y] = M_{(T,A,B),s}^{n}[q,\eta,P,X][q',\eta',P',Y]$ (by aperiodicity of $(T,A,B)$). By Lemma \ref{lem:int}, we obtain $M_{T',s}^{n+1}[S_1,X_{(q,\eta,P)}][S_2, Y_{(q',\eta',P')}] = i, \alpha_1, \alpha_2$ and hence condition $(2)$ of Lemma \ref{lem:int} is satisfied. \end{itemize} Using Lemma \ref{lem:int} and the aperiodicity of $M_{(T,A,B)}$, we obtain the aperiodicity of $M_{T'}$. \end{proof} \section{Proofs from Section \ref{subsec:sst}} \label{app:sst-basics} \subsection{Example of SST} \label{eg:sst} \begin{figure}[h] \tikzstyle{trans}=[-latex, rounded corners] \begin{center} \scalebox{0.9}{ \begin{tikzpicture}[->,>=stealth',shorten >=1pt,auto, semithick,scale=.8] \tikzstyle{every state}=[fill=gold] \node[initial,state, initial text={}] at (6, 4) (A1) {$1$} ; \node[state] at (13,4) (B1) {$2$} ; \path (A1) edge [loop above] node {$\#\left|\begin{array}{llllllll}x:=x\#\\y:=\varepsilon\\z:=\varepsilon\end{array}\right.$} (A1); \path (A1) edge [bend left] node [above] {$\begin{array}{lllll} \vspace{-2mm}\\ \alpha|(x,y,z):=(x,\alpha y \alpha, z \alpha)\end{array}$} (B1); \path (B1) edge [loop above] node {$\alpha\left|\begin{array}{lllll} x:=x\\ y:=\alpha y \alpha \\ z := z\alpha \end{array}\right.$} (A1); \path (B1) edge [bend left] node [below] {$\begin{array}{l} \#|(x,y,z):=(xy\#,\varepsilon,\varepsilon) \\ \vspace{-2mm}\end{array}$} (A1); \end{tikzpicture} } \end{center} \vspace{-1em} \caption{Transformation $f_1$ given as streaming string transducers with $F(\set{2}) = x z$ is the output associated with Muller set $\set{2}$. $\alpha$ stands for $a,b$. \label{appfig:fo-example}} \vspace{-1em} \end{figure} \begin{example} The transformation $f_1$ introduced is definable by the \sst{} in Figure~\ref{appfig:fo-example}. Consider the successive valuations of $x, y,$ and $z$ upon reading the string $ab\#a^\omega$. \[ \begin{array}{r|cccccccccccccccccccccccccccccccccccccc} & & \!\!\!\!a\!\!\!\! & & \!\!\!\!b\!\!\!\! & & \!\!\!\!\#\!\!\!\! & & \!\!\!\!a\!\!\!\! & & \!\!\!\!a\!\!\!\! &\dots\\ \hline x & \varepsilon & & \varepsilon & & \varepsilon & & baab\# & & baab\# & & baab\# \\ y & \varepsilon & & aa & & baab & & \varepsilon & & aa & & aaaa \\ z & \varepsilon & & a & & ab & & \varepsilon & & a & & aa \\ \end{array} \] Notice that the limit of $x z$ exists and equals $baab\#a^\omega$. \end{example} \subsection{Transition Monoid for SSTs} \begin{lemma} \label{sst-monoid} $(M_T,\times,\textbf{1})$ is a monoid, where $\times$ is defined as matrix multiplication and the identity element $\textbf{1}$ is the matrix with diagonal elements $(\emptyset,\emptyset,\dots,\emptyset)$ and all non-diagonal elements being $\bot$. \end{lemma} \begin{proof} Consider any matrix $M_s$ where $s \in \Sigma^*$. Assume the \sst{} has $g$ variables $X_1, \dots, X_g$, and $m$ states $\{p_1, \dots, p_m\}$. Let there be an ordering $(p_i, X_j) < (p_i, X_l)$ for $j < l$ and $(p_i, X_j) < (p_k, X_j)$ for $i < k$ used in the transition monoid. Consider a row corresponding to some $(p,X_i)$. The only entries that are not $\perp$ in this row are of the form $[(p,X_i)(r,X_1)], \dots, [(p,X_i)(r,X_g)]$, for some state $r$, such that there is a run of $s$ from $p$ to $r$, $p, r \in \{p_1, \dots, p_m\}$. The $(x_1, \dots, x_n)$ component of all entries $[(p,X_i)(r,X_1)], \dots, [(p,X_i)(r,X_g)]$ are same. Let it be $(\kappa_1, \dots, \kappa_n)$. Let $[(p,X_i)(r,X_j)]=k_{ij}(\kappa_1, \dots, \kappa_n)$. Consider $M_s.\textbf{1}$. The $[(p,X)(r,-)]$ entries of the product are obtained from the $(p,X)$th row of $M_s$ and the $(r,-)$th column of $\textbf{1}$. The $(r,-)$th column of $\textbf{1}$ has exactly one $1(\emptyset, \dots, \emptyset)$, while all other elements are $\bot$. Let $p=p_c$ and $r=p_h$, with $c<h$ ($h<c$ is similar). Then the $[(p,X)(r,Y)]$ entry for $X=X_i, Y=X_j$ is of the form $\bot+ \dots + \bot + k_{ij}(\kappa_1, \dots, \kappa_n).1(\emptyset, \dots, \emptyset)+k_{i+1~j}(\kappa_1, \dots, \kappa_n).\bot$ $+ \dots +k_{gj}(\kappa_1, \dots, \kappa_n).\bot + \dots +\bot.$ Clearly, this is equal to $k_{ij}(\kappa_1, \dots, \kappa_n)$. Similarly, it can be shown that the $[(p,X_i)(r,X_j)]$th entry of $M_s$ is preserved in $\textbf{1}.M_s$. Associativity can also be checked easily. \end{proof} The mapping $M_{\bullet}$, which maps any string $s$ to its transition matrix $M_s$, is a morphism from $(\Sigma^*, ., \epsilon)$ to $(M_T, \times, \textbf{1})$. We say that the transition monoid $M_T$ of an \sst{} $T$ is $n$-bounded if in all entries $(j, (x_1, \dots,x_n))$ of the matrices of $M_T$, $j \leq n$. Clearly, any $n$-bounded transition monoid is finite. A streaming string transducer is \emph{aperiodic} if its transition monoid is aperiodic. An $\omega$-streaming string transducer with $n$ Muller acceptance sets is 1-bounded if its transition monoid is 1-bounded. That is, for all strings $s$, and all pairs $(p,Y)$, $(q,X)$, $M_s[p,Y][q,X] \in \{\bot\} \cup \{ (i, (x_1, \dots,x_n)) \mid i \leq 1\}$. \subsection{Example of Transition Monoid} \label{app-example} \begin{figure}[h] \begin{center} \begin{tikzpicture}[->,>=stealth',shorten >=1pt,auto,node distance=1cm, semithick,scale=0.9,every node/.style={scale=0.9}] \tikzstyle{every state}=[fill=gold,minimum size=1em] \node[state,fill=gold,accepting] at (8,-3) (C) {$q$} ; \node[state,fill=gold,accepting] at (2,-3) (B) {$r$} ; \node[state,initial, initial where=above,initial text={},fill=gold] at (5,-3) (D) {$t$} ; \path(B) edge[loop left] node {$a \mid X:=Xb$} (B); \path(C) edge[loop right] node {$b\mid Y:=YX$} (C); \path(B) edge[bend left] node {$b$} (D); \path(D) edge[bend left] node {$b\mid X:=bY$} (B); \path(C) edge[bend left] node {$a\mid X:=bX$} (D); \path(D) edge[bend left] node {$a\mid Y:=aX$} (C); \node[state,initial,initial text={},initial where=above,fill=gold,accepting] at (12,-3) (A1) {$u$} ; \node[state,fill=gold,accepting] at (15,-3) (B1) {$v$} ; \path(A1) edge[bend left] node {$a$} (B1); \path(B1) edge[bend left] node {$b$} (A1); \path(B1) edge[loop below] node {$a$} (B1); \path(A1) edge[loop below] node {$b$} (A1); \end{tikzpicture} \end{center} \caption{Muller accepting set of \sst{} on the left= $\{\{q\},\{r\}\}$. Also, $F(q)=XY$, $F(r)=X$. Muller accepting set for \sst{} on the right=$\{u,v\}$;$X:=X$ and $Y:=Y$ on all edges, $F(\{u,v\})=X$.} \label{fig:tm-ex1} \end{figure} In figure \ref{fig:tm-ex1} consider the strings $ab$ and $bb$ for the automaton on the left. The transition monoids are \[ M_{ab}=\begin{blockarray}{ccccccc} & \matindex{(t,X)} & \matindex{(t,Y)} & \matindex{(q,X)} & \matindex{(q,Y)} & \matindex{(r,X)} & \matindex{(r,Y)} \\ \begin{block}{c(cccccc)} \matindex{(t,X)} & \bot & \bot & 1,(0,0) & 2,(0,0) &\bot &\bot \\ \matindex{(t,Y)} & \bot & \bot & 0, (0,0) & 0, (0,0) & \bot & \bot \\ \matindex{(q,X)} & \bot & \bot & \bot & \bot & 0, (0,0) & 0, (0,0)\\ \matindex{(q,Y)} & \bot & \bot & \bot & \bot & 1, (0,0) &1, (0,0) \\ \matindex{(r,X)} & 1,(0,0) & 0, (0,0) & \bot & \bot & \bot & \bot\\ \matindex{(r,Y)} & 0,(0,0) & 1,(0,0) & \bot & \bot & \bot &\bot \\ \end{block} \end{blockarray} \] \[ M_{bb}=\begin{blockarray}{ccccccc} & \matindex{(t,X)} & \matindex{(t,Y)} & \matindex{(q,X)} & \matindex{(q,Y)} & \matindex{(r,X)} & \matindex{(r,Y)} \\ \begin{block}{c(cccccc)} \matindex{(t,X)} & 0,(0,0) & 0,(0,0) & \bot & \bot &\bot &\bot \\ \matindex{(t,Y)} & 1,(0,0) & 1,(0,0) & \bot & \bot & \bot & \bot \\ \matindex{(q,X)} & \bot & \bot & 1,(1,0) & 2, (1,0) &\bot & \bot\\ \matindex{(q,Y)} & \bot & \bot & 0,(1,0) & 1,(1,0) & \bot & \bot \\ \matindex{(r,X)} & \bot & \bot & \bot & \bot & 0,(0,0) & 0,(0,0)\\ \matindex{(r,Y)} & \bot & \bot & \bot & \bot & 1,(0,0) &1,(0,0) \\ \end{block} \end{blockarray} \] It can be checked that $M_{abbb}=M_{ab}M_{bb}$. Likewise, the transition monoid for $a,b$ for the automaton on the right is \[ M_{a}=\begin{blockarray}{ccccc} & \matindex{(u,X)} & \matindex{(u,Y)} & \matindex{(v,X)} & \matindex{(v,Y)} \\ \begin{block}{c(cccc)} \matindex{(u,X)} & \bot & \bot & 1(1) & 0(1) \\ \matindex{(u,Y)} & \bot & \bot & 0(1) & 1(1) \\ \matindex{(v,X)} & \bot & \bot & 1(v) & 0(v) \\ \matindex{(v,Y)} & \bot & \bot & 0(v) & 1(v) \\ \end{block} \end{blockarray} \] \[ M_{b}=\begin{blockarray}{ccccc} & \matindex{(u,X)} & \matindex{(u,Y)} & \matindex{(v,X)} & \matindex{(v,Y)} \\ \begin{block}{c(cccc)} \matindex{(u,X)} & 1(u) & 0(u) & \bot & \bot \\ \matindex{(u,Y)} & 0(u) & 1(u) & \bot & \bot \\ \matindex{(v,X)} & 1(1) & 0(1) & \bot & \bot \\ \matindex{(v,Y)} & 0(1) & 1(1) & \bot & \bot \\ \end{block} \end{blockarray} \] \begin{proposition}\label{prop:fodomain} The domain of an aperiodic \sst{} is FO-definable. \end{proposition} \begin{proof} Let $T = (\Sigma, \Gamma, Q, q_0, Q_f, \delta, \mathcal{X}, \rho, F)$ be an aperiodic SST and $M_T$ its (aperiodic) transition monoid. Let us define a function $\varphi$ which associates with each matrix $M\in M_T$, the $|Q|\times |Q|$ Boolean matrix $\varphi(M)$ defined by $\varphi(M)[p][q] =(x_1, \dots, x_n)$ iff there exist $X,Y\in\mathcal{X}$ such that $M_T[p,X][q,Y]=(k,(x_1, \dots,x_n))$, with $k \geq 0$. Clearly, $\varphi(M_T)$ is the transition monoid of the underlying input automaton of $T$ (ignoring the variable updates). The result follows, since the homomorphic image of an aperiodic monoid is aperiodic. \end{proof} \subsection{Aperiodic Streaming String Transducers} \section{Introduction} \label{sec:introduction} \input{introduction} \section{Preliminaries} \label{sec:prelims} \input{prelims} \section{Aperiodic Transformations} \label{sec:aperiodic} \input{model.tex} \section{FOTs $\equiv$ Aperiodic 2WST$_{sf}$ } \label{sec:fot-2wst} \input{fot-2wst.tex} \section{Aperiodic SST $\subset$ FOT} \label{sec:sst-fot} \label{sec:2wst-sst} \input{sst-fot-short} \vspace{-0.5em} \section{Conclusion} \vspace{-0.5em} We extended the notion of aperiodicity from finite string transformations to that on infinite strings. We have shown a way to generalize transition monoids for deterministic Muller automata to streaming string transducers and two-way finite state transducers that capture the FO definable global transformations. A interesting and natural next step is to investigate LTL-definable transformations, their connection with FO-definable transformations, and their practical applications in verification and synthesis. \subsection{First-Order Logic Definable Transformations} \label{subsec:fotrans} Courcelle~\cite{Cour94} initiated the study of structure transformations using MSO logic. His main idea was to define a transformation $(w, w') \in R$ by defining the string model of $w'$ using a finite number of copies of positions of the string model of $w$. The existence of positions, various edges, and position labels are then given as $\mathrm{MSO}(\Sigma)$ formulas. We study a restriction of his formalism to use first-order logic to express string transformations. \begin{definition} An \emph{FO string transducer} is a tuple $T {=} (\Sigma, \Gamma, \phi_{\mathrm{dom}}, C, \phi_{\mathrm{pos}}, \phi_{\preceq})$ where: \begin{itemize} \item $\Sigma$ and $\Gamma$ are finite sets of input and output alphabets; \item $\phi_\mathrm{dom}$ is a closed $\mathrm{FO}(\Sigma)$ formula characterizing the domain of the transformation; \item $C {=} \set{1, 2, \ldots, n}$ is a finite index set; \item $\phi_{\mathrm{pos}} {=} \set{ \phi^c_\gamma(x) : c \in C \text{ and } \gamma \in \Gamma}$ is a set of $\mathrm{FO}(\Sigma)$ formulae with a free variable $x$; \item $\phi_{\preceq} {=} \set{\phi^{c, d}_\preceq(x, y) : c, d \in C}$ is a set of $\mathrm{FO}(\Sigma)$ formulae with two free variables $x$ and $y$. \end{itemize} The transformation $\inter{T}$ defined by $T$ is as follows. A string $s$ with $\struc{s} = (\mathrm{dom}(s), \preceq, (L_a)_{a \in \Sigma})$ is in the domain of $\inter{T}$ if $s \models \phi_{\mathrm{dom}}$ and the output string $w$ with structure \\ $M = (D, \preceq^M, (L^M_\gamma)_{\gamma\in\Gamma})$ is such that \begin{itemize} \item $D = \set{ v^c \::\: v \in \mathrm{dom}(s), c \in C \text{ and } \phi^c(v)}$ is the set of positions where $\phi^c(v) \rmdef \lor_{\gamma \in \Gamma} \phi^c_\gamma(v)$; \item $\preceq^M {\subseteq} D {\times} D$ is the ordering relation between positions and it is such that for $v, u \in dom(s)$ and $c, d \in C$ we have that $v^c \preceq^M u^d$ if $w \models \phi^{c, d}_\preceq(v, u)$; and \item for all $v^c \in D$ we have that $L_\gamma^M(v^c)$ iff $\phi^c_\gamma(v)$. \end{itemize} \end{definition} Observe that the output is unique and therefore FO transducers implement functions. A string $s \in \Sigma^{\omega}$ can be represented by its string-graph with $dom(s)=\{i \in \mathbb{N}\}$, $\preceq=\{(i,j) \mid i \leq j\}$ and $L_a(i)$ iff $s[i] = a$ for all $i$. From now on, we denote the string-graph of $s$ as $s$ only. We say that an FO transducer is a \emph{string-to-string} transducer if its domain is restricted to string graphs and the output is also a string graph. We say that a string-to-string transformation is FO-definable if there exists an FO transducer implementing the transformation. We write $\fot{}$ for the set of FO-definable string-to-string $\omega$-transformations. \begin{example} Figure~\ref{fig:fo-example}(c) shows a transformation for an \fot{} that implements the transformation $f_1: \Sigma^* \{a,b\}^{\omega} \to \Sigma^{\omega}$, where $\Sigma=\{a,b, \#\}$, by replacing every maximal $\#$ free string $u$ into $\overline{u}u$. Let $\texttt{is\_string}_{\#}$ be an FO formula that defines a string that contains a $\#$, and let $\texttt{reach}_{\#}(x)$ be an FO formula that is true at a position which has a $\#$ at a later position. To define the \fot{} formally, we have $\phi_{dom}=\texttt{is\_string}_{\#}$, $\phi^1_{\gamma}(x) = \phi^2_{\gamma}(x) = L_{\gamma}(x) \wedge \neg L_{\#}(x) \wedge \texttt{reach}_{\#}(x)$, since we only keep the non $\#$ symbols that can ``reach'' a $\#$ in the input string in the first two copies. $\phi^3_{\gamma}(x) = L_{\#}(x) \vee (\neg L_{\#}(x) \wedge \neg \texttt{reach}_{\#}(x))$, since we only keep the $\#$'s, and the infinite suffix from where there are no $\#$'s. The full list of formulae $\phi^{i,j}$ can be seen in Appendix \ref{app:fot-example}. \end{example} \subsection{Two-way Transducers (\ensuremath{\mathsf{2WST}}{})} \label{sec:2wst} A \ensuremath{\mathsf{2WST}}{} is a tuple $T=(Q, \Sigma, \Gamma, q_0, \delta, F)$ where $\Sigma, \Gamma$ are respectively the input and output alphabet, $q_0$ is the initial state, $\delta$ is the transition function and $F \subseteq 2^Q$ is the acceptance set. The transition function is given by $\delta : Q \times \Sigma \rightarrow Q \times \Gamma^* \times \{1,0,-1\}$. A configuration of the \ensuremath{\mathsf{2WST}}{} is a pair $(q,i)$ where $q \in Q$ and $i \in \mathbb N$ is the current position of the input string. A run $r$ of a \ensuremath{\mathsf{2WST}}{} on a string $s \in \Sigma^{\omega}$ is a sequence of transitions $(q_0,i_0{=}0) \xrightarrow{a_1/c_1,dir} (q_1, i_1) \xrightarrow{a_2/c_2,dir} (q_2, i_2) \cdots$ where $a_i \in \Sigma$ is the input letter read and $c_i \in \Gamma^*$ is the output string produced during a transition and $i_j$s are the positions updated during a transition for all $j \in dom(s)$. $dir$ is the direction, $\{1,0,-1\}$. W.l.o.g. we can consider the outputs to be over $\Gamma \cup \{\epsilon\}$. The output $out(r)$ of a run $r$ is simply a concatenation of the individual outputs, i.e. $c_1c_2 \dots \in \Gamma^{\infty}$. We say that the transducer reads the whole string $s$ when $\sup \set{i_n \mid 0 \leq n <|r|} {=} \infty$. The output of $s$, denoted $T(s)$ is defined as $out(r)$ only if $\Omega(r) \in F$ and $r$ reads the whole string $s$. We write $\inter{T}$ for the transformation captured by $T$. \noindent{\bf Transition Monoid.} The transition monoid of a \ensuremath{\mathsf{2WST}}{} $T=(Q, \Sigma, \Gamma, q_0, \delta, \{F_1, \dots, F_n\})$ is the transition monoid of its underlying automaton. However, since the \ensuremath{\mathsf{2WST}}{} can read their input in both directions, the transition monoid definition must allow for reading the string starting from left side and leaving at the left (left-left) and similar other behaviors (left-right, right-left and right-right). Following~\cite{CD15}, we define the behaviors $\Bb_{xy}(w)$ of a string $w$ for $x, y \in \{\ell,r\}$. $\Bb_{\ell r}(w)$ is a set consisting of pairs $(p,q)$ of states such that starting in state $p$ in the left side of $w$ the transducer leaves $w$ in right side in state $q$. In the example in figure \ref{fig:fo-example}(a), we have $\Bb_{\ell r}(ab\#) = \set{(t,t), (p,t), (q,t)}$ and $\Bb_{r r}(ab\#) = \set{(q,t), (t,t), (p,q)}$. Two words $w_1, w_2$ are ``equivalent'' if their left-left, left-right, right-left and right-right behaviors are same. That is, $\Bb_{xy}(w_1)=\Bb_{xy}(w_2)$ for $x, y \in \{\ell,r\}$. The transition monoid of $T$ is the conjunction of the 4 behaviors, which also keeps track, in addition, the set of states witnessed in the run, as shown for the deterministic Muller automata earlier. For each string $w \in \Sigma^*$, $x, y \in \{\ell,r\}$, and states $p,q$, the entries of the matrix $M_{u}^{xy}[p][q]$ are of the form $\bot$, if there is no run from $p$ to $q$ on word $u$, starting from the side $x$ of $u$ and leaving it in side $y$, and is $(x_1,\ldots x_n)$ otherwise, where $x_i$ is defined exactly as in section \ref{ap-omega}. For equivalent words $u_1,u_2$, we have $M_{u_1}^{xy}[p][q]= M_{u_2}^{xy}[p][q]$ for all $x,y \in \{\ell, r\}$ and states $p,q$. Addition and multiplication of matrices are defined as in the case of Muller automata. See Appendix \ref{app:2way} for more details. Note that behavioral composition is quite complex, due to left-right movements. In particular, it can be seen from the example that $\Bb_{\ell r}(ab\#a\#)=\Bb_{\ell r}(ab\#)\Bb_{\ell \ell}(a\#)\Bb_{r r}(ab\#)\Bb_{\ell r}(a\#)$. Since we assume that the \ensuremath{\mathsf{2WST}}{} $T$ is deterministic and completely reads the input string $\alpha \in \Sigma^{\omega}$, we can find a unique factorization $\alpha= [\alpha_0\dots \alpha_{p_1}][\alpha_{p_1+1} \dots \alpha_{p_2}]\dots$ such that the run of $\mathcal{A}$ on each $\alpha$-block progresses from left to right, and each $\alpha$-block will be processed completely. That is, one can find a unique sequence of states $q_{p_1}, q_{p_2}, \dots$ such that the \ensuremath{\mathsf{2WST}}{} starting in initial state $q_0$ at the left of the block $\alpha_0\dots \alpha_{p_1}$ leaves it at the right in state $q_{p_1}$, starts the next block $\alpha_{p_1+1} \dots \alpha_{p_2}$ from the left in state $q_{p_1}$ and leaves it at the right in state $q_{p_2}$ and so on. We consider the languages $L^{xy}_{pq}$ for $x,y \in \{\ell,r\}$, where $\ell,r$ respectively stand for left and right. $L^{\ell \ell}_{pq}$ stands for all strings $w$ such that, starting at state $p$ at the left of $w$, one leaves the left of $w$ in state $q$. Similarly, $L^{r\ell}_{pq}$ stands for all strings $w$ such that starting at the right of $w$ in state $p$, one leaves the left of $w$ in state $q$. In figure \ref{fig:fo-example}(a), note that starting on the right of $ab\#$ in state $t$, we leave it on the right in state $t$, while we leave it on the left in state $p$. So $ab\# \in L^{r r}_{tt}, L^{r \ell}_{tp}$. Also, $ab \# \in L^{r r}_{pq}$. A \ensuremath{\mathsf{2WST}}{} is said to be \emph{aperiodic} iff for all strings $u \in \Sigma^*$, all states $p,q$ and $x,y \in \{l,r\}$, there exists some $m \geq 1$ such that $u^m \in L_{pq}^{xy}$ iff $u^{m+1} \in L_{pq}^{xy}$. \noindent{\bf Star-Free Lookaround.} We wish to introduce aperiodic \ensuremath{\mathsf{2WST}}{} that are capable of capturing FO-definable transformations. However, as we discussed earlier (see page \pageref{two-way} in the paragraph on two-way transducers) \ensuremath{\mathsf{2WST}}{} without look-ahead are strictly less expressive than MSO transducers. To remedy this we study aperiodic \ensuremath{\mathsf{2WST}}{}s enriched with star-free look-ahead (star-free look-back can be assumed for free). An aperiodic $\ensuremath{\mathsf{2WST}}{}$ with star-free look-around ($\ensuremath{\mathsf{2WST}}{}_{\ensuremath{sf}}$) is a tuple $(T, A, B)$ where $A$ is an aperiodic Muller look-ahead automaton and $B$ is an aperiodic look-behind automaton, resp., and $T = (\Sigma, \Gamma, Q, q_0, \delta, F)$ is an aperiodic \ensuremath{\mathsf{2WST}}{} as defined earlier except that the transition function $\delta: Q \times Q_B \times \Sigma \times Q_A \to Q \times \Gamma \times \set{-1, 0, +1}$ may consult look-ahead and look-behind automata to make its decisions. Let $s \in \Sigma^\omega$ be an input string, and $L(A, p)$ be the set of infinite strings accepted by $A$ starting in state $p$. Similarly, let $L(B,r)$ be the set of finite strings accepted by $B$ starting in state $r$. We assume that $\ensuremath{\mathsf{2WST}}_{\ensuremath{sf}}$ are \emph{deterministic} i.e. for every string $s \in \Sigma^\omega$ and every input position $i \leq |s|$, there is exactly one state $p \in Q_A$ and one state $r \in Q_B$ such that $s(i)s(i+1)\ldots \in L(A, p)$ and $s(0)s(1)\ldots s(i-1) \in L(B, r)$. If the current configuration is $(q, i)$ and $\delta(q, r, s(i), p) = (q', z, d)$ is a transition, such that the string $s(i)s(i+1)\ldots \in L(A, p) $ and $s(0)s(1)\ldots s(i-1) \in L(B, r)$, then $\ensuremath{\mathsf{2WST}}_{\ensuremath{sf}}$ writes $z \in \Gamma$ on the output tape and updates its configuration to $(q', i+d)$. Figure~\ref{fig:fo-example}(a) shows a \ensuremath{\mathsf{2WST}}{} with star-free look-ahead $\texttt{reach}_{\#}(x)$ capturing the transformation $f_1$ (details in App.~\ref{app:2way}). \subsection{Streaming $\omega$-String Transducers (SST)}\label{subsec:sst} Streaming string transducers(\sst{}s) manipulate a finite set of string variables to compute their output. In this section we introduce aperiodic \sst{}s for infinite strings. Let $\mathcal{X}$ be a finite set of variables and $\Gamma$ be a finite alphabet. A substitution $\sigma$ is defined as a mapping ${\sigma : \mathcal{X} \to (\Gamma \cup \mathcal{X})^*}$. A valuation is defined as a substitution $\sigma: \mathcal{X} \to \Gamma^*$. Let $\mathcal{S}_{\mathcal{X}, \Gamma}$ be the set of all substitutions $[\mathcal{X} \to (\Gamma \cup \mathcal{X})^*]$. Any substitution $\sigma$ can be extended to $\hat{\sigma}: (\Gamma \cup \mathcal{X})^* \to (\Gamma \cup \mathcal{X})^*$ in a straightforward manner. The composition $\sigma_1 \sigma_2$ of two substitutions $\sigma_1$ and $\sigma_2$ is defined as the standard function composition $\hat{\sigma_1} \sigma_2$, i.e. $\hat{\sigma_1}\sigma_2(x) = \hat{\sigma_1}(\sigma_2(x))$ for all $x \in \mathcal{X}$. We say that a string $u \in (\Gamma \cup \mathcal{X})^*$ is \emph{copyless{}} (or linear) if each $x \in \mathcal{X}$ occurs at most once in $u$. A substitution $\sigma$ is copyless{} if $\hat{\sigma}(u)$ is copyless{}, for all linear $u \in (\Gamma \cup \mathcal{X})^*$ . \begin{definition} A \emph{streaming $\omega$-string transducer} (\sst{}) is a tuple $T = (\Sigma, \Gamma, Q, q_0, \delta, \mathcal{X}, \rho, F)$ \begin{itemize} \item $\Sigma$ and $\Gamma$ are finite sets of input and output alphabets; \item $Q$ is a finite set of states with initial state $q_0$; \item $\delta:Q \times \Sigma \to Q$ is a transition function and $\mathcal{X}$ is a finite set of variables; \item $\rho : (Q \times \Sigma) \to \mathcal{S}_{\mathcal{X}, \Gamma}$ is a variable update function to copyless{} substitutions; \item $F: 2^Q \rightharpoonup \mathcal{X}^*$ is an output function such that for all $P \in \mathrm{dom}(F)$ the string $F(P)$ is copyless{} of form $x_1 \dots x_n$, and for $q, q' \in P$ and $a \in \Sigma$ s.t. $q' = \delta(q, a)$ we have \begin{itemize} \item $\rho(q, a)(x_i) = x_i$ for all $i < n$ and $\rho(q, a)(x_n) = x_n u$ for some $u \in (\Gamma \cup \mathcal{X})^*$. \end{itemize} \end{itemize} \end{definition} The concept of a run of an \sst{} is defined in an analogous manner to that of a Muller automaton. The sequence $\seq{\sigma_{r, i}}_{0 \leq i \leq |r|}$ of substitutions induced by a run $r = q_0 \xrightarrow{a_1} q_1 \xrightarrow{a_2} q_2 \ldots$ is defined inductively as the following: $\sigma_{r, i} {=} \sigma_{r, i{-}1} \rho(q_{i-1}, a_{i})$ for $0 < i \leq |r|$ and $\sigma_{r, 0} = x \in X \mapsto \varepsilon$. The output $T(r)$ of an infinite run $r$ of $T$ is defined only if $F(r)$ is defined and equals $T(r) \rmdef \lim_{i \to \infty} \seq{\sigma_{r, i}(F(r))}$, when the limit exists. If not, we pad $\bot^{\omega}$ to the obtained finite string to get $\lim_{i \to \infty} \seq{\sigma_{r, i}(F(r)) \bot^\omega}$ as the infinite output string. The assumptions on the output function $F$ in the definition of an \sst{} ensure that this limit always exist whenever $F(r)$ is defined. Indeed, when a run $r$ reaches a point from where it visits only states in $P$, these assumptions enforce the successive valuations of $F(P)$ to be an increasing sequence of strings by the prefix relation. The padding by unique letter $\bot$ ensures that the output is always an $\omega$-string. The output $T(s)$ of a string $s$ is then defined as the output $T(r)$ of its unique run $r$. The transformation $\inter{T}$ defined by an SST $T$ is the partial function $\set{(s, T(s)) \::\: T(s) \text{ is defined}}$. See Appendix \ref{eg:sst} for an example. We remark that for every \sst{} $T = (\Sigma, \Gamma, Q, q_0, \delta, \mathcal{X}, \rho, F)$, its domain is always an $\omega$-regular language defined by the Muller automaton $(\Sigma, Q, q_0, \delta, \mathrm{dom}(F))$, which can be constructed in linear time. However, the range of an \sst{} may not be $\omega$-regular. For instance, the range of the \sst-definable transformation $a^n\#^\omega \mapsto a^nb^n\#^\omega$ ($n\geq 0$) is not $\omega$-regular. \noindent{\bf Aperiodic Streaming String Transducers.} We define the notion of aperiodic \sst{}s by introducing an appropriate notion of transition monoid for transducers. The transition monoid of an \sst{} $T$ is based on the effect of a string $s$ on the states as well as on the variables. The effect on variables is characterized by, what we call, flow information that is given as a relation that describes the number of copies of the content of a given variable that contribute to another variable after reading a string $s$. Let $T = (\Sigma, \Gamma, Q, q_0, \delta, \mathcal{X}, \rho, F)$ be an \sst{}. Let $s$ be a string in $\Sigma^*$ and suppose that there exists a run $r$ of $T$ on $s$. Recall that this run induces a substitution $\sigma_r$ that maps each variable $X \in \mathcal{X}$ to a string $u \in (\Gamma \cup \mathcal{X})^*$. For string variables $X,Y\in \mathcal{X}$, states $p,q\in Q$, and $n \in \mathbb N$ we say that $n$ copies of $Y$ flow to $X$ from $p$ to $q$ if there exists a run $r$ on $s \in \Sigma^*$ from $p$ to $q$, and $Y$ occurs $n$ times in $\sigma_r(X)$. We extend the notion of transition monoid for the Muller automata as defined in Section~\ref{sec:prelims} for the transition monoid for SSTs to equip it with variables. Formally, the transition monoid $\mathcal{M}_T {=} (M_T, \times, \textbf{1})$ of an \sst{} $T = (\Sigma, \Gamma, Q, q_0, \delta, \mathcal{X}, \rho, \set{F_1, \ldots, F_n})$ is such that $M_T$ is a set of $|Q\times \mathcal{X}|\times|Q \times \mathcal{X}|$ square matrices over $(\mathbb N \times (\set{0, 1} \cup 2^Q)^n) \cup \set{\bot}$ along with matrix multiplication $\times$ defined for matrices in $M_T$ and identity element $\textbf{1} \in M_T$ is the matrix whose diagonal entries are $(1, (\emptyset, \emptyset, \ldots, \emptyset))$ and non-diagonal entries are all $\bot$'s. Formally $M_T {=} \set{M_s \::\: s \in \Sigma^*}$ is defined using matrices $M_s$ for strings $s\in \Sigma^*$ s.t. $M_s[(p, X)][(q, Y)] {=} \bot$ if there is no run from state $p$ to state $q$ over $s$ in $T$, otherwise $M_s[(p, X)][(q, Y)] = (k, (x_1, \dots, x_n)) \in (\mathbb N \times (\set{0, 1} \cup 2^Q)^n)$ where $x_i$ is defined exactly as in section \ref{ap-omega}, and $k$ copies of variable $X$ flow to variable $Y$ from state $p$ to state $q$ after reading $s$. We write $(p,X) \rightsquigarrow^u_{\alpha} (q,Y)$ for $M_u[(p, X)][(q, Y)] = \alpha$. It is easy to see that $M_\epsilon = \textbf{1}$. The operator $\times$ is simply matrix multiplication for matrices in $M_T$, however we need to define addition $\oplus$ and multiplication $\odot$ for elements $(\set{0, 1} \cup 2^Q)^n \cup \set{\bot}$ of the matrices. We have $\alpha_1 \odot \alpha_2 = \bot$ if $\alpha_1 = \bot$ or $\alpha_2 = \bot$, and if $\alpha_1 = (k_1, (x_1, \ldots, x_n))$ and $\alpha_2 = (k_2, (y_1, \ldots, y_n))$ then $\alpha_1 \odot \alpha_2 = (k_1\times k_2, (z_1, \ldots, z_n))$ s.t. for all $1 \leq i \leq n$ $z_i$ are defined as in~(\ref{eq1:product}) from Section~\ref{ap-omega}. Note that due to determinism of the \sst{}s we have that for every matrix $M_s$ and every state $p$ there is at most one state $q$ such that $M_s[p][q] \not = \bot$ and hence the only addition rules we need to introduce is $\alpha \oplus \bot = \bot \oplus \alpha = \alpha$, $0 \oplus 0=0$, $1 \oplus 1=1$ and $\kappa \oplus \kappa=\kappa$ for $\kappa \subseteq Q$. It is easy to see that $(M_T, \times, \textbf{1})$ is a monoid and we give a proof in Appendix~\ref{app:sst-basics}. We say that the transition monoid $M_T$ of an \sst{} $T$ is $1$-bounded if in all entries $(j, (x_1, \dots,x_n))$ of the matrices of $M_T$, $j \leq 1$. A streaming string transducer is \emph{aperiodic} if its transition monoid is aperiodic. \subsection{Aperiodic Monoids for $\omega$-String Languages} A monoid $\mathcal{M}$ is an algebraic structure $(M, \cdot, e)$ with a non-empty set $M$, a binary operation $\cdot$, and an identity element $e \in M$ such that for all $x, y, z \in M$ we have that $(x \cdot (y \cdot z)) {=} ((x \cdot y) \cdot z)$, and $x \cdot e = e \cdot x$ for all $x \in M$. We say that a monoid $(M, \cdot, e)$ is \emph{finite} if the set $M$ is finite. A monoid that we will use in this paper is the \emph{free} monoid, $(\Sigma^*,\cdot, \epsilon)$, which has a set of finite strings over some alphabet $\Sigma$ with the empty string $\epsilon$ as the identity. We define the notion of acceptance of a language via monoids. A morphism (or homomorphism) between two monoids $\mathcal{M} = (M, \cdot, e)$ and $\mathcal{M}' = (M', \times, e')$ is a mapping $h: M \rightarrow M'$ such that $h(e) = e'$ and $h(x \cdot y) = h(x) \times h(y)$. Let $h: \Sigma^* \rightarrow M,$ be a morphism from free monoid $(\Sigma^*, \cdot, \epsilon)$ to a finite monoid $(M, \cdot, e)$. Two strings $u, v \in \Sigma^{*}$ are said to be similar with respect to $h$ denoted $u \sim_h v$, if for some $n \in \mathbb{N} \cup \{\infty\}$, we can factorize $u, v$ as $u=u_1u_2 \dots u_n$ and $v=v_1 v_2 \dots v_n$ with $u_i, v_i \in \Sigma^+$ and $h(u_i)=h(v_i)$ for all $i$. Two $\omega$-strings are $h$-similar if we can find factorizations $u_1u_2 \dots$ and $v_1v_2 \dots$ such that $h(u_i) =h(v_i)$ for all $i$. Let $\cong$ be the transitive closure of $\sim_h$. $\cong$ is an equivalence relation. Note that since $M$ is finite, the equivalence relation $\cong$ is of finite index. For $w \in \Sigma^{\infty}$ we define $[w]_h$ as the set $\set{u \mid u \cong w}$. We say that a morphism $h$ accepts a language $L \subseteq \Sigma^{\infty}$ if $w \in L$ implies $[w]_h \subseteq L$ for all $w \in \Sigma^{\infty}$. We say that a monoid $(M, ., e)$ is \emph{aperiodic}~\cite{Strau94} if there exists $n \in \mathbb N$ such that for all $x \in M$, $x^n = x^{n+1}$. Note that for finite monoids, it is equivalent to require that for all $x\in M$, there exists $n\in \mathbb N$ such that $x^n= x^{n+1}$. A language $L \subseteq \Sigma^{\infty}$ is said to be aperiodic iff it is recognized by some morphism to a finite and aperiodic monoid (See Appendix~\ref{app:example-aperiodic}). \subsection{First-Order Logic for $\omega$-String Languages} A string $s \in \Sigma^{\omega}$ can be represented as a relational structure $\struc{s} {=} (\mathrm{dom}(s), \preceq^s, (L^s_a)_{a \in \Sigma})$, called the string model of $s$, where $\mathrm{dom}(s) = \set{1, 2, \ldots}$ is the set of positions in $s$, $\preceq^s$ is a binary relation over the positions in $s$ characterizing the natural order, i.e. $(x, y) \in \preceq^s$ if $x \leq y$; $L^s_a$, for all $a \in \Sigma$, are the unary predicates that hold for the positions in $s$ labeled with the alphabet $a$, i.e., $L^s_a(i)$ iff $s[i]=a$, for all $i\in \mathrm{dom}(s)$. When it is clear from context we will drop the superscript $s$ from the relations $\preceq^s$ and $L^s_a$. Properties of string models over the alphabet $\Sigma$ can be formalized by first-order logic denoted by $\mathrm{FO}(\Sigma)$. Formulas of $\mathrm{FO}(\Sigma)$ are built up from variables $x, y, \ldots$ ranging over positions of string models along with \emph{atomic formulae} of the form $x{=} y, x {\preceq} y$, and $L_a(x)$ for all $a \in \Sigma$ where formula $x{=}y$ states that variables $x$ and $y$ point to the same position, the formula $x \preceq y$ states that position corresponding to variable $x$ is not larger than that of $y$, and the formula $L_a(x)$ states that position $x$ has the label $a \in \Sigma$. Atomic formulae are connected with \emph{propositional connectives} $\neg$, $\wedge$, $\lor$, $\to$, and \emph{quantifiers} $\forall$ and $\exists$ that range over node variables and we use usual semantics for them. We say that a variable is \emph{free} in a formula if it does not occur in the scope of some quantifier. A \emph{sentence} is a formula with no free variables. We write $\phi(x_1, x_2, \ldots, x_k)$ to denote that at most the variables $x_1, \ldots, x_k$ occur free in $\phi$. For a string $s\in \Sigma^*$ and for positions $n_1, n_2, \ldots, n_k \in \mathrm{dom}(s)$ we say that $s$ with valuation $\nu = (n_1, n_2, \ldots, n_k)$ satisfies the formula $\phi(x_1, x_2, \ldots, x_k)$ and we write $(s, \nu) \models \phi(x_1, x_2, \ldots, x_k)$ or $s \models \phi(n_1, n_2, \ldots, n_k)$ if formula $\phi$ with $n_i$ as the interpretation of $x_i$ is satisfied in the string model $\struc{s}$. The language defined by an FO sentence $\phi$ is $L(\phi) \rmdef \set{s \in \Sigma^{\omega} \::\: \struc{s} \models \phi}$. We say that a language $L$ is FO-definable if there is an FO sentence $\phi$ such that $L = L(\phi)$. The following is a well known result. \begin{theorem}\cite{Strau94} \label{thm:fo-aperiodic} A language $L\subseteq \Sigma^*$ is \ensuremath{\mathrm{FO}}-definable iff it is aperiodic. \end{theorem} \subsection{Aperiodic Muller Automata for $\omega$-String Languages} \label{ap-omega} A deterministic Muller automaton (DMA) is a tuple $\mathcal{A} = (Q, q_0, \Sigma, \delta, F)$ where $Q$ is a finite set of states, $q_0 \in Q$ is the initial state, $\Sigma$ is an input alphabet, $\delta: Q \times \Sigma \to Q$ is a transition function, and $F \subseteq 2^Q$ are the accepting (Muller) sets. For states $q, q' \in Q$ and letter $a \in \Sigma$ we say that $(q, a, q')$ is a transition of the automaton $\mathcal{A}$ if $\delta(q,a ) = q'$ and we write $q \xrightarrow{a} q'$. We say that there is a run of $\mathcal{A}$ over a finite string $s = a_1 a_2 \ldots a_n \in \Sigma^*$ from state $p$ to state $q$ if there is a finite sequence of transitions $\seq{(p_0, a_1, p_1), (p_1, a_2, p_2), \ldots, (p_{n-1}, a_n, p_n)} \in (Q \times \Sigma \times Q)^*$ with $p = p_0$ and $q = p_n$. We write $L_{p, q}$ for the set of finite strings such that there is a run of $\mathcal{A}$ over $w$ from $p$ to $q$. We say that there is a run of $\mathcal{A}$ over an $\omega$-string $s = a_1 a_2 \ldots \in \Sigma^\omega$ if there is a sequence of transitions $\seq{(q_0, a_1, q_1), (q_1, a_2, q_2), \ldots} \in (Q \times \Sigma \times Q)^\omega$. For an infinite run $r$, we denote by $\Omega(r)$ the set of states that occur infinitely often in $r$. We say that an $\omega$-string $w$ is accepted by a Muller automaton $\mathcal{A}$ if the run of $\mathcal{A}$ on $w$ is such that $\Omega(r) \in F$ and we write $L(\mathcal{A})$ for the set of all $\omega$-strings accepted by $\mathcal{A}$. \\ A Muller automaton $\mathcal{A}$ is \emph{aperiodic} iff there exists some $m {\geq} 1$ s.t. $u^m {\in} L_{p,q}$ iff $u^{m+1} {\in} L_{p,q}$ for all $u \in \Sigma^*$ and $p, q \in Q$. Another equivalent way to define aperiodicity is using the transition monoid, which, to the best of our knowledge, has not been defined in the literature for Muller automata. Given a DMA $\mathcal{A} {=} (Q, q_0, \Sigma, \Delta, \set{F_1, \ldots, F_n})$, we define the transition monoid $\mathcal{M}_\mathcal{A} {=} (M_\mathcal{A}, \times, \textbf{1})$ of $\mathcal{A}$ as follows: $M_\mathcal{A}$ is a set of $|Q|\times|Q|$ square matrices over $(\set{0, 1} \cup 2^Q)^n \cup \set{\bot}$. Matrix multiplication $\times$ is defined for matrices in $M_\mathcal{A}$ with identity element $\textbf{1} \in M_\mathcal{A}$, where $\textbf{1}$ is the matrix whose diagonal entries are $(\emptyset, \emptyset, \ldots, \emptyset)$ and non-diagonal entries are all $\bot$'s. Formally, $M_\mathcal{A} {=} \set{M_s \::\: s \in \Sigma^*}$ is defined using matrices $M_s$ for strings $s\in \Sigma^*$ s.t. $M_s[p][q] {=} \bot$ if there is no run from $p$ to $q$ over $s$ in $\mathcal{A}$. Otherwise, let $P$ be the set of states (excluding $p$ and $q$) witnessed in the unique run from $p$ to $q$. Then $M_s[p][q] = (x_1, \dots, x_n) \in (\set{0, 1} \cup 2^Q)^n$ where (1) $x_i = 0$ iff $\exists r \in P \cup \{p,q\}$, $r \notin F_i$; (2) $x_i=1$ iff $P \cup \{p,q\}=F_i$, and (3) $x_i=P \cup \{p,q\}$ iff $P \cup \{p,q\} \subset F_i$. It is easy to see that $M_\epsilon = \textbf{1}$, since $\epsilon$ takes a state to itself and nowhere else. The operator $\times$ is simply matrix multiplication for matrices in $M_\mathcal{A}$, however we need to define addition $\oplus$ and multiplication $\odot$ for elements $(\set{0, 1} \cup 2^Q)^n \cup \set{\bot}$ of the matrices. We have $\alpha_1 \odot \alpha_2 = \bot$ if $\alpha_1 = \bot$ or $\alpha_2 = \bot$, and if $\alpha_1 = (x_1, \ldots, x_n)$ and $\alpha_2 = (y_1, \ldots, y_n)$ then $\alpha_1 \odot \alpha_2 = (z_1, \ldots, z_n)$ s.t.: \begin{equation} \resizebox{0.7 \textwidth}{!} { \tag{$\star$} \label{eq1:product} $z_i = \begin{cases} 0 & \text{if $x_i = 0$ or $y_i = 0$}\\ 1 & \text{if $(x_i = y_i = 1)$ or if $(x_i, y_i \subset F_i \text { and } x_i \cup y_i = F_i)$}\\ 1 & \text{if ($x_i = 1$ and $y_i \subset F_i$) or ($y_i = 1$ and $x_i \subset F_i$)} \\ x_i \cup y_i & \text{if $x_i, y_i \subset F_i$ and $x_i \cup y_i \subset F_i$} \end{cases} $ } \end{equation} Due to determinism, we have that for every matrix $M_s$ and every state $p$ there is at most one state $q$ such that $M_s[p][q] \not = \bot$ and hence the only addition rule we need to introduce is $\alpha \oplus \bot = \bot \oplus \alpha = \alpha$. It is easy to see that $(M_\mathcal{A}, \times, \textbf{1})$ is a monoid (a proof is deferred to the Appendix \ref{app:basic-muller}). It is straightforward to see that a Muller automaton is aperiodic if and only if its transition monoid is aperiodic. Appendix \ref{app:aper-equiv} gives a proof showing that a language $L \subseteq \Sigma^{\omega}$ is aperiodic iff there is an aperiodic DMA accepting it. \subsection{FO-definability of variable flow} To construct the \fot{}, we make use of the output structure for \sst{}. It is an intermediate representation of the output, and the transformation of any input string into its \sst-output structure will be shown to be FO-definable. For any \sst{} $T$ and string $s\in\mathrm{dom}(T)$, the \sst-output structure of $s$ is a relational structure $G_T(s)$ obtained by taking, for each variable $X\in\mathcal{X}$, two copies of $\mathrm{dom}(s)$, respectively denoted by $X^{in}$ and $X^{out}$. For notational convenience we assume that these structures are labeled on the edges. A pair $(X,i)$ is \emph{useful} if the content of variable $X$ before reading $s[i]$ will be part of the output after reading the whole string $s$. This structure satisfies the following \emph{invariants}: for all $i\in dom(s)$, $(1)$ the nodes $(X^{in}, i)$ and $(X^{out}, i)$ exist only if $(X,i)$ is useful, and $(2)$ there is a directed path from $(X^{in}, i)$ to $(X^{out}, i)$ whose labels are same as variable $X$ computed by $T$ after reading $s[i]$. \input{outputgraph.tex} \noindent We define \sst-output structures formally in Appendix \ref{app:sst-output}, however, the illustration above shows an \sst-output structure. We show only the variable updates. Dashed arrows represent variable updates for useless variables, and therefore does not belong the \sst-output structure. The path from $(X^{in},6)$ to $(X^{out},6)$ gives the contents of $X$ ($ceaaafbc$) after 6 steps. We write $O_T$ for the set of strings appearing in right-hand side of variable updates. We next show that the transformation that maps an $\omega$-string $s$ into its output structure is FO-definable, whenever the SST is 1-bounded and aperiodic. Using the fact that variable flow is FO-definable, we show that for any two variables $X,Y$, we can capture in FO, a path from $(X^d, i)$ to $(Y^e, j)$ for $d, e\in \{in, out\}$ in $G_T(s)$ and all positions $i, j$. \begin{lemma}\label{lem:fopath} Let $T$ be an \textbf{aperiodic,1-bounded} \sst{} $T$. For all $X,Y\in \mathcal{X}$ and all $d,d'\in \{in,out\}$, there exists an FO[$\Sigma$]-formula $\text{path}_{X,Y,d,d'}(x,y)$ with two free variables such that for all strings $s\in\mathrm{dom}(T)$ and all positions $i,j\in dom(s)$, $s\models \text{path}_{X,Y,d,d'}(i,j)$ iff there exists a path from $(X^d,i)$ to $(Y^{d'},j)$ in $G_T(s)$. \end{lemma} The proof of Lemma \ref{lem:fopath} is in Appendix \ref{app:fopath}. As seen in Appendix (in Proposition \ref{prop:fostates}) one can write a formula $\phi_q(x)$ (to capture the state $q$ reached) and formula $\psi^{Rec}_P$ (to capture the recurrence of a Muller set $P$) in an accepting run after reading a prefix. For each variable $X \in \mathcal{X}$, we have two copies $X^{in}$ and $X^{out}$ that serve as the copy set of the \fot{}. As given by the \sst{} output-structure, for each step $i$, state $q$ and symbol $a$, a copy is connected to copies in the previous step based on the updates $\rho(q,a)$. The full details of the \fot{} construction handling the Muller acceptance condition of the \sst{} are in Appendix \ref{app-fot-cons}. \section{Aperiodic $\ensuremath{\mathsf{2WST}}_{\ensuremath{sf}}$ $\subset$ Aperiodic SST} \label{sec:2wst-sst} \input{2wst-sst} \subsection{FO-definability of variable flow} \begin{proposition}\label{prop:foflow} Let $T$ be an aperiodic,1-bounded \sst{} $T$ with set of variables $\mathcal{X}$. For all variables $X,Y\in \mathcal{X}$, there exists an FO-formula $\phi_{X\rightsquigarrow Y}(x,y)$ with two free variables such that, for all strings $s\in dom(T)$ and any two positions $i\leq j\in dom(s)$, $s\models \phi_{X\rightsquigarrow Y}(i,j)$ iff $(q_i,X)\rightsquigarrow^{s[i{+}1{:}j]}_1 (q_j,Y)$, where $q_0\dots q_n \dots$ is the accepting run of $T$ on $s$. \end{proposition} Let $X\in\mathcal{X}$, $s\in\mathrm{dom}(T)$, $i\in \mathrm{dom}(s)$. Let $P \in 2^Q$ be a Muller accepting set with output $F(P)=X_1 \dots X_n$. Let $r = q_0\dots q_{n}\dots$ be an accepting run of $T$ on $s$. Then there is a position $j$ such that $\forall k >j$, only states in $P$ are visited. On all transitions beyond $j$, $X_1, \dots, X_{n-1}$ remain unchanged, while $X_n$ is updated as $X_nu$ for some $u \in (\mathcal{X} \cup \Gamma)^*$. We say that the pair $(X,i)$ is \emph{useful} if the content of variable $X$ before reading $s[i]$ will be part of the output after reading the whole string $s$. Formally, for an accepting run $r = q_0\dots q_{j}q_{j+1}\dots$, such that only states from the Muller set $P$ are seen after position $j$, we say that $(X,i)$ is useful for $s$ if $(q_{i-1}, X)\rightsquigarrow^{s[i{:}j] }_1(q,X_k)$ for some variable $X_k \in F(P)$, $k \leq n$, or if $(q_{i-1}, X)\rightsquigarrow^{s[i{:}\ell] }_1(q,X_n)$, with $q \in P$ and $\ell >j$. Thanks to Proposition \ref{prop:foflow}, this property is FO-definable (Proofs of propositions \ref{prop:foflow} and \ref{prop:contribution} are in Appendix \ref{app:varflow}). \begin{proposition}\label{prop:contribution} For all $X\in\mathcal{X}$, there exists an FO-formula $\text{useful}_X(i)$ s.t. for all strings $s\in dom(T)$ and all positions $i\in \mathrm{dom}(s)$, $s\models \text{useful}_X(i)$ iff $(X,i)$ is useful for string $s$. \end{proposition} The output structure for SSTs over finite words has been introduced in \cite{FKT14}. It is an intermediate representation of the output, and the transformation of any input string into its \sst-output structure will be shown to be FO-definable. For any \sst{} $T$ and string $s\in\mathrm{dom}(T)$, the \sst-output structure of $s$ is a relational structure $G_T(s)$ obtained by taking, for each variable $X\in\mathcal{X}$, two copies of $\mathrm{dom}(s)$, respectively denoted by $X^{in}$ and $X^{out}$. For notational convenience we assume that these structures are labeled on the edges. This structure satisfies the following \emph{invariants}: for all $i\in dom(s)$, $(1)$ the nodes $(X^{in}, i)$ and $(X^{out}, i)$ exist only if $(X,i)$ is useful, and $(2)$ there is a directed path from $(X^{in}, i)$ to $(X^{out}, i)$ whose sequence of labels is equal to the value of the variable $X$ computed by $T$ after reading $s[i]$. The condition on usefulness of nodes implies that \sst-output structures consist of a single directed component, and therefore they are edge-labeled string structures. \input{outputgraph} As an example of \sst-output structure consider Fig. \ref{fig:outputgraph}. We show only the variable updates. Dashed arrows represent variable updates for useless variables, and therefore does not belong the \sst-output structure. Initially the variable content of $Z$ is equal to $\epsilon$. It is represented by the $\epsilon$-edge from $(Z^{in}, 0)$ to $(Z^{out}, 0)$ in the first column. Then, variable $Z$ is updated to $Zc$. Therefore, the new content of $Z$ starts with $\epsilon$ (represented by the $\epsilon$-edge from $(Z^{in},1)$ to $(Z^{in},0)$, which is concatenated with the previous content of $Z$, and then concatenated with $c$ (it is represented by the $c$-edge from $(Z^{out},0)$ to $(Z^{out},1)$). Note that the invariant is satisfied. The content of variable $X$ at position 5 is given by the label of the path from $(X^{in},5)$ to $(X^{out},5)$, which is $c$. Also note that some edges are labelled by strings with several letters, but there are finitely many possible such strings. In particular, we denote by $O_T$ the set of all strings that appear in right-hand side of variable updates. \sst-output structures are defined formally in Appendix \ref{app:sst-output}. We next show that the transformation that maps an $\omega$-string $s$ into its output structure is FO-definable, whenever the SST is 1-bounded and aperiodic. Using the fact that variable flow is FO-definable, we show that for any two variables $X,Y$, we can capture in FO, a path from $(X^d, i)$ to $(Y^e, j)$ for $d, e\in \{in, out\}$ in $G_T(s)$ and all positions $i, j$. We are in state $q_i$ at position $i$. For example, there is a path from $(X^{in},3)$ to $(Z^{in},1)$ in Figure \ref{fig:outputgraph}. However, $Z$ at position 1 does not flow into $X$ at position 3. Note that this is because of the update $X:=XY$ at position 6, and $Z$ at position 1 flows into $Y$ at position 5, and $X$ at position 3 flows into $X$ at position 5, and $X$ and $Y$ are catenated in order at position 5 to define $X$ at position 6. Similarly, there is a path from $(Z^{d},1)$ to $(Y^{out},5)$ for $d \in \{in,out\}$. This is because $Z$ at position 1 flows into $Z$ at position 4, and this value of $Z$ is used in updating $Y$ at position 5 as $Y:=bZc$. Lastly, there is a path from $(Y^{in},5)$ to $(Z^{d},2)$. This is because $Z$ at position 2 flows into $Y$ at position 5 by the update $Y:=YbZc$. Thus, the SST output graphs have a nice property, which connects a path from $(X^d,i)$ to $(Y^{d'},j)$ based on the variable flow, and the catenation of variables in updates. Formally, let $T$ be an \textbf{aperiodic,1-bounded} \sst{} $T$. Let $s\in dom(T)$, $G_T(s)$ its \sst-output structure and $r=q_0\dots q_n\dots$ the accepting run of $T$ on $s$. For all variables $X,Y\in \mathcal{X}$, all positions $i,j\in \mathrm{dom}(s)\cup\{0\}$, all $d,d'\in\{in,out\}$, there exists a path from node $(X^{d},i)$ to node $(Y^{d'},j)$ in $G_T(s)$ iff $(X,i)$ and $(Y,j)$ are both useful and one of the following conditions hold: either \begin{enumerate} \item $(q_j,Y)\rightsquigarrow^{s[j{+}1{:}i]}_1(q_i,X)$ and $d = in$, or \item $(q_i,X)\rightsquigarrow^{s[i{+}1{:}j]}_1 (q_j,Y)$ and $d' = out$, or \item there exists $k\geq max(i,j)$ and two variables $X',Y'$ such $(q_i,X)\rightsquigarrow^{s[i{{+}1:}k]}_1 (q_k,X')$, $(q_j,Y)\rightsquigarrow^{s[j{+}1{:}k]}_1 (q_k,Y')$ and $X'$ and $Y'$ are concatenated in this order\footnote{by concatenated we mean that there exists a variable update whose rhs is of the form $\dots X'\dots Y'\dots$} by $r$ when reading $s[k+1]$. \end{enumerate} Using the fact that variable flow is FO-definable, we have \begin{lemma}\label{lem:fopath} Let $T$ be an \textbf{aperiodic,1-bounded} \sst{} $T$. For all $X,Y\in \mathcal{X}$ and all $d,d'\in \{in,out\}$, there exists an FO[$\Sigma$]-formula $\text{path}_{X,Y,d,d'}(x,y)$ with two free variables such that for all strings $s\in\mathrm{dom}(T)$ and all positions $i,j\in dom(s)$, $s\models \text{path}_{X,Y,d,d'}(i,j)$ iff there exists a path from $(X^d,i)$ to $(Y^{d'},j)$ in $G_T(s)$. \end{lemma} The proof of Lemma \ref{lem:fopath} is in Appendix \ref{app:fopath}. We can now formally construct an FOT given the 1-bounded, aperiodic, SST. The sequence of variable substitutions of an aperiodic, 1-bounded SST over a string can be represented as the copy set in FOTs. As seen in Appendix (in Proposition \ref{prop:fostates}) one can write a formula $\phi_q(x)$ to capture the state $q$ reached in an accepting run after reading a prefix. For each variable $X \in \mathcal{X}$, we have two copies $X^{in}$ and $X^{out}$ that serve as the copy set of the FOT. As given by the SST output-structure, for each step $i$, state $q$ and symbol $a$, a copy is connected to copies in the previous step based on the updates $\rho(q,a)$. In the following, we detail this in FO. Lemma \ref{lem:fopath} then captures a path in the FOT. The full details of the FOT construction can be found in Appendix \ref{app-fot-cons}. \end{proof}
c5a48d8da4587e1c5a513dfdae25f189d040a288
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} While recurrent neural networks (RNNs) are seeing widespread success across many tasks, the fundamental architecture presents challenges to typical training algorithms. In particular, the problem of `vanishing/exploding gradients'~\cite{hochreiter1991untersuchungen} in gradient-based optimization persists, where gradients either vanish or diverge as one goes deeper into the network, resulting in slow training or numerical instability. The long short-term memory (LSTM) network~\cite{hochreiter1997long} was designed to overcome this issue. Recently, the use of norm-preserving operators in the transition matrix - the matrix of weights connecting subsequent \emph{internal} states - of the RNN have been explored~\cite{arjovsky2015unitary,mikolov2014learning,le2015simple}. Using operators with bounded eigenvalue spectrum should, as demonstrated by~\citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary}, bound the norms of the gradients in the network, assuming an appropriate non-linearity is applied. Unitary matrices satisfy this requirement and are the focus of this work. Imposing unitarity (or orthogonality) on this transition matrix is however challenging for gradient-based optimization methods, as additive updates typically do not preserve unitarity. Solutions include \emph{re-unitarizing} after each batch, or using a \emph{parametrization} of unitary matrices closed under addition. In this work we propose a solution in the second category, using results from the theory of Lie algebras and Lie groups to define a general parametrization of unitary matrices in terms of skew-Hermitian matrices (elements of the Lie algebra associated to the Lie group of unitary matrices). As explained in more detail below, elements of this Lie algebra can be identified with unitary matrices, while the algebra is closed under addition, forming a vector space over real numbers. While we are motivated by the issues of RNNs, and we consider an application in RNNs, our primary focus here is on a core question: how can unitary matrices be learned? \emph{Assuming} the choice of a unitary transition matrix is an appropriate modelling choice, the gradients on this operator should ultimately guide it towards unitarity, if it is possible under the parametrization or learning scheme used. It is therefore useful to know which approach is best. We distil the problem into its simplest form (a learning task described in detail later), so that our findings cannot be confounded by other factors specific to the RNN long-term memory task, before demonstrating our parametrization in that setting. \subsection{Related work} We draw most inspiration from the recent work of~\citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary}, who proposed a specific parametrization of unitary matrices and demonstrated its utility in standard long-term memory tasks for RNNs. We describe their parametrization here, as we use it later. Citing the difficulty of obtaining a general and efficient parametrization of unitary matrices, they use the fact that the unitary group is closed under matrix multiplication to form a composite operator: \begin{equation} U = \mathbf{D}_3 \mathbf{R}_2 \mathcal{F}^{-1} \mathbf{D}_2 \Pi \mathbf{R}_1 \mathcal{F} \mathbf{D}_1 \label{eqn:arjovsky} \end{equation} where each component is unitary and easily parametrized: \begin{itemize} \item $\mathbf{D}$ is a diagonal matrix with entries of the form $e^{i \alpha}$, $\alpha \in \mathbb{R}$ \item $\mathbf{R}$ is a complex reflection operator; $\mathbf{R} = \mathbb{I} - 2\frac{\mathbf{v}\mathbf{v}^{\dagger}}{\| \mathbf{v} \|^2}$ ($\dagger$ denotes Hermitian conjugate) \item $\mathcal{F}$ and $\mathcal{F}^{-1}$ are the Fourier and inverse Fourier transforms (or, in practice, their discrete matrix representations) \item $\Pi$ is a fixed permutation matrix \end{itemize} In total, this parametrization has $7n$ real learnable parameters ($2n$ for each reflection and $n$ for each diagonal operator), so describes a subspace of unitary matrices (which have $n^2$ real parameters). Nonetheless, they find that an RNN using this operator as its transition matrix outperforms LSTMs on the adding and memory tasks described first in~\citeauthor{hochreiter1997long}~\shortcite{hochreiter1997long}. This prompted us to consider other parametrizations of unitary matrices which might be more expressive or interpretable. \citeauthor{mikolov2014learning}~\shortcite{mikolov2014learning} constrain a part of the transition matrix to be close to the identity, acting as a form of long-term memory store, while \citeauthor{le2015simple}~\shortcite{le2015simple} \emph{initialize} it to the identity, and then use ReLUs as non-linearities. \citeauthor{henaff2016orthogonal}~\shortcite{henaff2016orthogonal} study analytic solutions to the long-term memory task, supporting observations and intuitions that orthogonal (or unitary) matrices would be appropriate as transition matrices for this task. They also study initializations to orthogonal and identity matrices, and consider experiments where an additional term in the loss function encourages an orthogonal solution to the transition matrix, without using an explicit parametrization. \citeauthor{saxe2013exact}~\shortcite{saxe2013exact} study exact solutions to learning dynamics in deep networks and find that orthogonal weight initializations at each layer lead to depth-independent learning (thus escaping the vanishing/exploding gradient problem). Interestingly, they attribute this to the eigenvalue spectrum of orthogonal matrices lying on the unit circle. They compare with weights initialized to random, scaled Gaussian values, which preserve norms in expectation (over values of the random matrix) and find orthogonal matrices superior. It therefore appears that preserving norms is \emph{not} sufficient to stabilize gradients over network depth, but that the eigenvalue spectrum must also be strictly controlled. In a related but separate vein, \citeauthor{krueger2015regularizing}~\shortcite{krueger2015regularizing} penalize the difference of difference of norms between subsequent hidden states in the network. This is \emph{not} equivalent to imposing orthogonality of the \emph{transition} matrix, as the norm of the hidden state may be influenced by the inputs and non-linearities, and their method directly addresses this norm. The theory of Lie groups and Lie algebras has seen most application in machine learning for its use in capturing notions of \emph{invariance}. For example, \citeauthor{miao2007learning}~\shortcite{miao2007learning}, learn infinitesimal Lie group generators (elements of the Lie algebra) associated with affine transformations of images, corresponding to visual perceptual invariances. This is different to our setting as our generators are already known (we assume the Lie group $U(n)$) and wish to learn the coefficients of a given transformation relative to that basis set of generators. However, our approach could be extended to the case where the basis of $\mathfrak{u}(n)$ is \emph{unknown}, and must be learned. As we find later (appendix B), the choice of basis can impact performance, and so may be an important consideration. \citeauthor{cohen2014learning}~\shortcite{cohen2014learning} learn commutative subgroups of $SO(n)$ (known as toroidal subgroups), motivated by learning the irreducible representations of the symmetry group corresponding to invariant properties of images. Their choice of group parametrization is equivalent to selecting a particular basis of the corresponding Lie algebra, as they describe, but primarily exploit the algebra to understand properties of toroidal subgroups. \citeauthor{tuzel2008learning}~\shortcite{tuzel2008learning} perform motion estimation by defining a regression function in terms of a function on the Lie algebra of affine transformations, and then learning this. This is similar to our approach in the sense that they do optimization in the Lie algebra, although as they consider two-dimensional affine transformations only, their parametrization of the Lie algebra is straight forward. Finally, \citeauthor{warmuthorthogonal}~\shortcite{warmuthorthogonal} describe an online learning algorithm for orthogonal matrices -- which are the real-valued equivalent to unitary matrices. They also claim that the approach is extends easily to unitary matrices. \subsection{Structure of this paper} We begin with an introduction to the relevant facts and definitions from the theory of Lie groups and Lie algebras, to properly situate this work in its mathematical context. Further exposition is beyond the scope of this paper, and we refer the interested reader to any of the comprehensive introductory texts on the matter. We explain our parametrization in detail and describe a method to calculate the derivative of the matrix exponential - a quantity otherwise computationally intractable. Then, we describe a simple but clear experiment designed to test our core question of learning unitary matrices. We compare to an approach using the parametrization of~\citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary} and one using polar decomposition to `back-project' to the closest unitary matrix. We use this experimental set-up to probe aspects of our model, studying the importance of the choice of basis (appendix B), and the impact of the restricted parameter set used by one of the alternate approaches. We additionally implement our parametrization in a recurrent neural network as a `general unitary RNN', and evaluate its performance on standard long-memory tasks. \section{The Lie algebra $\mathfrak{u}(n)$} \label{section:lie} \subsection{Basics of Lie groups and Lie algebras} A Lie group is a group which is also a differentiable manifold, with elements of the group corresponding to points on the manifold. The group operations (multiplication and inversion) must be smooth maps (infinitely differentiable) back to the group. In this work we consider the group $U(n)$: the set of $n \times n$ unitary matrices, with matrix multiplication. These are the complex-valued analogue to orthogonal matrices, satisfying the property \begin{equation} U^{\dagger} U = U U^{\dagger} = \mathbb{I} \label{eqn:unitary} \end{equation} where $\dagger$ denotes the conjugate transpose (or Hermitian conjugate). Unitary matrices preserve matrix norms, and have eigenvalues lying on the (complex) unit circle, which is the desired property of the transition matrix in a RNN. The differentiable manifold property of Lie groups opens the door for the study of the Lie \emph{algebra}. This object is the \emph{tangent space} to the Lie group at the identity (the group must have an identity element). Consider a curve through the Lie group $U(n)$ - a one-dimensional subspace parametrized by a variable $t$, where $U(t=0) = \mathbb{I}$ (this is a matrix $U(t)$ in $U(n)$ parametrised by $t$, not a group). Consider the defining property of unitary matrices (Equation~\ref{eqn:unitary}), and take the derivative along this curve: \begin{equation} U(t)^{\dagger}U(t) = \mathbb{I} \rightarrow \dot{U}(t)^{\dagger} U(t) + U^{\dagger}(t) \dot{U}(t) = 0 \end{equation} Taking $t \rightarrow 0 $, $U(t) \rightarrow \mathbb{I}$, we have \begin{equation} \dot{U}(0)^{\dagger} \mathbb{I} + \mathbb{I}^{\dagger} \dot{U}(0) = 0 \Rightarrow \dot{U}(0)^{\dagger} = - \dot{U}(0) \label{eqn:alg} \end{equation} The elements $\dot{U}(0)$ belong to the Lie algebra. We refer to this Lie algebra as $\mathfrak{u}(n)$, and an arbitrary element as $L$. Then Equation~\ref{eqn:alg} defines the properites of these Lie algebra elements; they are $n \times n$ skew-Hermitian matrices: $L^{\dagger} = -L$. As vector spaces, Lie algebras are closed under addition. In particular $\mathfrak{u}(n)$ is a vector space over $\mathbb{R}$, so a \emph{real} linear combination of its elements is once again in $\mathfrak{u}(n)$ (this is also clear from the definition of skew-Hermitian). We exploit this fact later. Lie algebras are also endowed with an operation known as the \emph{Lie bracket}, which has many interesting properties, but is beyond the scope of this work. Lie algebras are interesting algebraic objects and have been studied deeply, but in this work we use $\mathfrak{u}(n)$ because of the \emph{exponential map}. Above, it was shown that elements of the algebra can be derived from the group (considering infinitesimal steps away from the identity). There is a \emph{reverse} operation, allowing elements of the group to be recovered from the algebra: this is the \emph{exponential map}. In the case of matrix groups, the exponential map is simply the \emph{matrix exponential}: \begin{equation} \exp(L) = \sum_{j = 0}^{\infty}\frac{L^j}{j!} \end{equation} Very simply, $L \in \mathfrak{u}(n)$, then $\exp(L) \in U(n)$. While this map is not in \emph{general} surjective, it so happens that $U(n)$ is a compact, connected group and so $\exp$ \emph{is} indeed surjective \cite{terrytao}. That is, for any $U \in U(n)$, there exists \emph{some} $L \in \mathfrak{u}(n)$ such that $\exp(L) = U$. Notably, while orthogonal matrices also form a Lie group $O(n)$, with associated Lie algebra $\mathfrak{o}(n)$ consisting of skew-symmetric matrices, $O(n)$ is \emph{not} connected, and so the exponential map can only produce \emph{special} orthogonal matrices - those with determinant one - $SO(n)$ being the component of $O(n)$ containing the identity. \subsection{Parametrization of $U(n)$ in terms of $\mathfrak{u}(n)$} The dimension of $\mathfrak{u}(n)$ as a real vector space is $n^2$. This is readily derived from noting that an arbitrary $n \times n$ complex matrix has $2n^2$ free real parameters, and the requirement of $L^{\dagger} = -L$ imposes $n^2$ constraints. So, a set of $n^2$ linearly-independent skew-Hermitian matrices defines a basis for the space; $\{ T_j \}_{j = \{1, \dots, n^2 \}}$. Then any element $L$ can be written as \begin{equation} L = \sum_{j=1}^{n^2} \lambda_j T_j \end{equation} where $\{\lambda_j\}_{j = 1, \dots, n^2}$ are $n^2$ real numbers; the coefficients of $L$ with respect to the basis. Using the exponential map, \begin{equation} U = \exp(L) = \exp\left(\sum_{j=1}^{n^2} \lambda_j T_j \right) \label{eqn:parametrization} \end{equation} we see that these $\{\lambda_j\}_{j = 1, \dots, n^2}$ suffice as \emph{parameters} of $U$ (given the basis $T_j$). This is the parametrization we propose. It has two attractive properties: \begin{enumerate} \item It is a fully general parametrization, as the exponential map is surjective \item Gradient updates on $\{\lambda_j\}_{j = 1, \dots, n^2}$ preserve unitarity automatically, as the algebra is closed under addition \end{enumerate} This parametrization means gradient steps are taken in the vector space of $\mathfrak{u}(n)$, rather than the manifold of $U(n)$, which \emph{may} provide a flatter cost landscape - although confirming this intuition would require further analysis. This work is intended to explore the use of this parametrization for learning arbitrary unitary matrices. There are many possible choices of basis for $\mathfrak{u}(n)$. We went for the following set of sparse matrices: \begin{enumerate} \item $n$ diagonal, imaginary matrices: $T_a$ is $i$ on the $a$-th diagonal, else zero. \item $\frac{n(n-1)}{2}$ symmetric, imaginary matrices with two non-zero elements, e.g., for $n=2$, $\left(\begin{matrix} 0 & i \\ i & 0 \end{matrix}\right)$ \item $\frac{n(n-1)}{2}$ anti-symmetric, real matrices with two non-zero elements, e.g., for $n=2$, $\left(\begin{matrix} 0 & 1 \\ -1 & 0 \end{matrix}\right)$ \end{enumerate} We explore the effects of choice of basis in appendix B. \subsection{Derivatives of the matrix exponential} The matrix exponential appearing in Equation~\ref{eqn:parametrization} poses an issue for gradient calculations. In general, the derivative of the matrix exponential does not have a closed-form expression, so computing gradients is intractable. In early stages of this work, we used the method of finite differences to approximate gradients, which would prohibit its use in larger-scale applications (such as RNNs). In the appendix we describe an investigation into using random projections to overcome this limitation, which while promising turned out to yield minimal benefit. We therefore sought mathematical solutions to this complexity issue, which we describe here and in further detail in the appendix. Exploiting the fact that $L$ is skew-Hermitian, we can derive an analytical expression for the derivative of $U$ with respect to each of its parameters, negating the need for finite differences. This expression takes the form: \begin{equation} \frac{\partial U}{\partial \lambda_a} = W V_a W^{\dagger} \label{eqn:twomat} \end{equation} where $W$ is a unitary matrix of eigenvectors obtained in the eigenvalue decomposition of $U$; $U = W D W^{\dagger}$, ($D$ = diag($d_1$, $\dots$, $d_{n^2}$); $d_i$ are the eigenvalues of $U$). Each $V_a$ is a matrix defined component-wise \begin{align} i = j:& V_{ii} =(W^{\dagger} T_a W)_{ii} e^{d_i}\label{eqn:vii}\\ i \neq j:& V_{ij} = (W^{\dagger} T_a W)_{ij} \left(\frac{e^{d_i} - e^{d_j}}{d_i - d_j}\right)\label{eqn:vij} \end{align} Where $T_a$ is the basis matrix of the Lie algebra in the $a$-th direction. We provide the derivation, based on work from~\citeauthor{kalbfleisch1985analysis}~\shortcite{kalbfleisch1985analysis} and~\citeauthor{jennrich1976fitting}~\shortcite{jennrich1976fitting} in Appendix A. We can simplify the expression $W^{\dagger} T_a W$ for each $T_a$, depending on the type of basis element. In these expressions, $\mathbf{w}_a$ refers to the $a$-th \emph{row} of W. \begin{enumerate} \item $T_a$ purely imaginary; $W^{\dagger} T_a W = i \cdot \mathrm{outer}(\mathbf{w}^*_a, \mathbf{w}_a)$ \item $T_a$ symmetric imaginary, nonzero in positions $(r, s)$ and $(s, r)$: $W^{\dagger} T_{rs} W = i \cdot (\mathrm{outer}(\mathbf{w}^*_s, \mathbf{w}_r) + \mathrm{outer}(\mathbf{w}^*_r, \mathbf{w}_s))$ \item $T_a$ antisymmetric real, nonzero in positions $(r, s)$ and $(s, r)$: $W^{\dagger} T_{rs} W = \mathrm{outer}(\mathbf{w}^*_r, \mathbf{w}_s) - \mathrm{outer}(\mathbf{w}^*_s, \mathbf{w}_r)$ \end{enumerate} These expressions follow from the sparsity of the basis and are derived in appendix A. Thus, we reduce the calculation of $W^{\dagger} T_a W$ from two matrix multplications to at most two vector outer products. Overall, we have reduced the cost of calculating gradients to a single eigenvalue decomposition, and for each parameter two matrix multiplications (equation~\ref{eqn:twomat}), one or two vector outer products, and element-wise multiplication of two matrices (equations~\ref{eqn:vii},~\ref{eqn:vij}). As we see in the RNN experiments, this actually makes our approach faster than the (restricted)uRNN of~\cite{arjovsky2015unitary} for roughly equivalent numbers of parameters. \section{Supervised Learning of Unitary Operators} \label{section:learning} We consider the supervised learning problem of learning the unitary matrix $U$ that generated a $\mathbf{y}$ from $\mathbf{x}$; $\mathbf{y} = U\mathbf{x}$, given examples of such $\mathbf{x}$s and $\mathbf{y}$s. This is the core learning problem that needs to be solved for the state-transformation matrix in RNNs. It is similar to the setting considered in \citeauthor{warmuthorthogonal}~\shortcite{warmuthorthogonal} (they consider an online learning problem). We compare a number of methods for learning $U$ at different values of $n$. We further consider the case where we have artificially restricted the number of learnable variables in our parametrization (for the sake of comparison), and generate a pathological change of basis to demonstrate the relevance of selecting a good basis (appendix B). \subsection{Task} The experimental setup is as follows: we create a $n \times n$ unitary matrix $U$ (the next section describes how this is done), then sample vectors $\mathbf{x} \in \mathbb{C}^n$ with normally-distributed coefficients. We create $\mathbf{y}_j = U \mathbf{x}_j + \epsilon_j$ where $\epsilon \sim \mathcal{N}(0, \sigma^2)$. The objective is to recover $U$ from the $\{\mathbf{x}_j, \mathbf{y}_j\}$ pairs by minimizing the squared Euclidean distance between predicted and true $\mathbf{y}$ values; \begin{equation} U = \mathop{\mathrm{argmin}}_U \frac{1}{N} \sum_j^N \| \hat{\mathbf{y_j}} - \mathbf{y_j} \|^2 = \mathop{\mathrm{argmin}}_U \frac{1}{N} \sum_j^N \| U \mathbf{x}_j - \mathbf{y_j} \|^2 \end{equation} While this problem is easily solved in the batch setting using least-squares, we wish to learn $U$ through mini-batch stochastic gradient descent, to emulate a deep learning scenario. For each experimental run (a single $U$), we generate one million training $\{\mathbf{x}_j, \mathbf{y}_j\}$ pairs, divided into batches of size 20. The test and validation sets both contain $100,000$ examples. In practice we set $\sigma^2 = 0.01$ and use a fixed learning rate of $0.001$. For larger dimensions, we run the model through the data for multiple epochs, shuffling and re-batching each time. All experiments were implemented in Python. The code is available here: \texttt{https://github.com/ratschlab/uRNN}. For the matrix exponential, we use the scipy builtin \texttt{expm}, which uses Pade approximation~\cite{al2009new}. We make use of the fact that $iL$ is Hermitian to use \texttt{eigh} (also in scipy) to perform eigenvalue decompositions. \subsection{Generating the ground-truth unitary matrix} \label{section:generating} The $U$ we wish to recover is generated by one of three methods: \begin{enumerate} \item QR decomposition: we create a $n \times n$ complex matrix with normally-distributed entries and then perform a QR decomposition, producing a unitary matrix $U$ and an upper triangular matrix (which is discarded). This approach is also used to sample orthogonal matrices in~\citeauthor{warmuthorthogonal}~\shortcite{warmuthorthogonal}, noting a result from~\citeauthor{stewart1980efficient}~\shortcite{stewart1980efficient} demonstrating that this is equivalent to sampling from the appropriate Haar measure. \item Lie algebra: given the standard basis of $\mathfrak{u}(n)$, we sample $n^2$ normally-distributed real $\lambda_j$ to produce $U = \exp{\left( \sum_j \lambda_j T_j\right)}$ \item Unitary composition: we compose parametrized unitary operators as in ~\citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary} (Equation~\ref{eqn:arjovsky}). The parameters are sampled as follows: angles in $D$ come from $\mathcal{U}(-\pi, \pi)$. The complex reflection vectors in $\mathbf{R}$ come from $\mathcal{U}(-s, s)$ where $s = \sqrt{\frac{6}{2n}}$. \end{enumerate} We study the effects of this generation method on test-set loss in a later section. While we find no significant association between generation method and learning approach, in our experiments we nonetheless average over an equal number of experiments using each method, to compensate for possible unseen bias. \subsection{Approaches} We compare the following approaches for learning $U$: \begin{enumerate} \item \texttt{projection}: $U$ is represented as an unconstrained $n \times n$ complex matrix, but after each gradient update we \emph{project} it to the closest unitary matrix, using polar decomposition~\cite{keller1975closest}. This amounts to $2n^2$ real parameters. \item \texttt{arjovsky}: $U$ is parametrized as in Equation~\ref{eqn:arjovsky}, which comes to $7n$ real parameters. \item \texttt{lie\_algebra}: (we refer to this as $\mathfrak{u}(n)$) $U$ is parametrized by its $n^2$ real coefficients $\{ \lambda_j \}$ in the Lie algebra, as in Equation \ref{eqn:parametrization}. \end{enumerate} As baselines we use the \texttt{true} matrix $U$, and a \texttt{random} unitary matrix $U_R$ generated by the same method as $U$ (in that experimental run). We also implemented the algorithm described in~\citeauthor{warmuthorthogonal}~\shortcite{warmuthorthogonal} and considered both unitary and orthogonal learning tasks (our parametrization contains orthogonal matrices as a special case) but found it too numerically unstable and therefore excluded it from our analyses. \subsection{Comparison of Approaches} \label{section:results} \begin{table*} \small \begin{tabular}{ r || c | c c c | c} $n$ & \texttt{true} & \texttt{projection} & \texttt{arjovsky} & \texttt{lie algebra} & \texttt{rand} \\ \hline 3 & $6.004 \pm 0.005 \times 10^{-4}$ & 8 $\pm$ 1 & $\mathbf{6.005 \pm 0.003 \times 10^{-4}}$ & $\mathbf{6.003 \pm 0.003 \times 10^{-4}}$ & $12.5 \pm 0.4$ \\ 6 & $\sim$ 0.001 & $15 \pm 1$ & $0.09 \pm 0.01 $ & $\mathbf{0.03 \pm 0.01}$ & 24 $\pm$ 1 \\ 8 & $\sim$ 0.002 & $14 \pm 1$ & $1.17 \pm 0.06$ & $\mathbf{0.014 \pm 0.006}$ & $31.6 \pm 0.6$ \\ 14 & $\sim$ 0.003 & $24 \pm 4$ & $10.8 \pm 0.3$ & $\mathbf{0.07 \pm 0.02}$ & $52 \pm 1$ \\ 20 & $\sim$ 0.004 & $38 \pm 3$ & $29.0 \pm 0.5$ & $\mathbf{0.47 \pm 0.03}$ & $81 \pm 2$ \end{tabular} \caption{Loss (mean $l_2$-norm between $\hat{y}_i$ and $y_i$) on the test set for the different approaches as the dimension of the unitary matrix changes. \texttt{true} refers to the matrix used to generate the data, \texttt{projection} is the approach of `re-unitarizing' using a polar decomposition after gradient updates, \texttt{arjovsky} is the composition approach defined in Equation~\ref{eqn:arjovsky}, $\mathfrak{u}(n)$ is our parametrization (Equation~\ref{eqn:parametrization}) and \texttt{rand} is a random unitary matrix generated in the same manner as \texttt{true}. Values in bold are the best for that $n$ (excluding \texttt{true}). The error for \texttt{true} is typically very small, so we omit it.} \label{table:accuracies} \end{table*} Table~\ref{table:accuracies} shows the test-set loss for different values of $n$ and different approaches for learning $U$. We performed between 6 and 18 replicates of each experiment, and show bootstrap estimates of means and standard errors over these replicates. As we can see, the learning task becomes more challenging as $n$ increases, but our parametrization ($\mathfrak{u}(n)$) consistently outperforms the other approaches. \subsection{Restricting to $7n$ parameters} \label{section:restriction} As mentioned, \texttt{arjovsky} uses only $7n$ parameters. To check if this difference accounts for the differences in loss observed in Table~\ref{table:accuracies}, we ran experiments where we fixed all but $7n$ (selected randomly) of the $\{\lambda_j\}$ in the \texttt{lie\_algebra} parametrization. The fixed parameters retained their initial values throughout the experiment. We observe that, as suspected, restricting to $7n$ parameters results in a performance degradation equivalent to that of \texttt{arjovsky}. \begin{table*} \small \begin{tabular}{r || c c | c} $n$ & \texttt{arjovsky} & \texttt{lie\_restricted} & \texttt{lie\_unrestricted} \\\hline 8 & $1.2 \pm 0.1$ & $1.0 \pm 0.2$ & $0.04 \pm 0.01$\\ 14 & $11.6 \pm 0.3$ & $ 12.6 \pm 0.4 $ & $0.25 \pm 0.03$ \\ 20 & $27.8 \pm 0.7$ & $28.0 \pm 0.6$ & $0.19 \pm 0.03$ \end{tabular} \caption{We observe that restricting our approach to the same number of learnable parameters as that of \cite{arjovsky2015unitary} causes a similar degradation in performance on the task. This indicates that the relatively superior performance of our model is explained by its generality in capturing arbitrary unitary matrices.} \label{table:restrict} \end{table*} Table~\ref{table:restrict} shows the results for $n=8, 14, 20$. The fact that the restricted case is consistently within error of the \texttt{arjovsky} model supports our hypothesis that the difference in learnable parameters accounts for the difference in performance. This suggests that generalising the model of \citeauthor{arjovsky2015unitary} to allow for $n^2$ parameters may result in performance similar to our approach. However, how to go about such a generalisation is unclear, as a naive approach would simply use a composition of $n^2$ operators, and this would likely become computationally intractable. \subsection{Method of generating $U$} \label{section:method} As described, we used three methods to generate the true $U$. One of these produces $U$ in the subspace available to the composition parametrization (Equation~\ref{eqn:arjovsky}), so we were curious to see if this parametrization performed better on experiments using that method. We were also concerned that generating $U$ using the Lie algebra parametrization might make the task too `easy' for our approach, as its random initialization could lie close to the true solution. Figure~\ref{fig:boxplots} shows box-plots of the distribution of test losses from these approaches for the three methods, comparing our approach ($\mathfrak{u}(n)$) with that of \citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary}, denoted \texttt{arjovsky}. To combine results from experiments using different values of $n$, we first scaled test-set losses by the performance of \texttt{rand} (the random unitary matrix), so the y-axis ranges from 0 (perfect) to 1 (random performance). The dotted line denotes the average (over methods) of the test-set loss for \texttt{true}, similarly scaled. The right panel in Figure~\ref{fig:boxplots} shows a zoomed-in version of the $\mathfrak{u}(n)$ result where the comparison with \texttt{true} is more meaningful, and a comparison with the case where we have restricted to $7n$ learnable parameters (see earlier). \begin{figure} \centering \includegraphics[scale=0.55]{boxplot_all_inkscape.pdf} \caption{We ask whether the method used to generate $U$ influences performance for different approaches to \emph{learning} $U$. Error bars are bootstrap estimates of 95\% confidence intervals. To compare across different $n$'s, we normalise each loss by the loss of \texttt{rand} for that $n$, and reporrt fractions. The dotted line is the \texttt{true} loss, similarly normalised. the choice of method to generate $U$ does not appear to affect test-set loss for the different approaches. \emph{Right}: Finer resolution on the $\mathfrak{u}(n)$ result in left panel. We also include the case where we restrict to $7n$ learnable parameters.} \label{fig:boxplots} \end{figure} We do not observe a difference (within error) between the methods, which is consistent between $\mathfrak{u}(n)$ and \texttt{arjovsky}. Our concern that using the Lie algebra to generate $U$ would make the task `too easy' for $\mathfrak{u}(n)$ was seemingly unfounded. \section{Unitary Recurrent Neural Network for Long Memory Tasks} To demonstrate that our approach is practical for use in deep learning, we incorporate it into a recurrent neural network to solve standard long-memory tasks. Specifically, we define a general unitary RNN with recurrence relation \begin{equation} \mathbf{h}_t = f\left( \beta U \mathbf{h}_{t-1} + V \mathbf{x}_t + \mathbf{b} \right) \end{equation} where $f$ is a nonlinearity, $\beta$ is a free scaling factor, $U$ is our unitary matrix parametrised as in equation~\ref{eqn:parametrization}, $\mathbf{h}_t$ is the hidden state of the RNN and $\mathbf{x}_t$ is the input data at `time-point' $t$. We refer to this as a `general unitary RNN' (guRNN), to distinguish it from the restricted uRNN of~\citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary}. We use the guRNN on two tasks: the `adding problem' and the `memory problem', first described in~\cite{hochreiter1997long}. For the sake of brevity we refer to~\cite{arjovsky2015unitary} for specific experimental details, as we use an identical experimental setup (reproduced in TensorFlow; see above github link for code). We compare our model (guRNN) with the restricted uRNN (ruRNN) parametrised as in equation~\ref{eqn:arjovsky}, a LSTM~\cite{hochreiter1997long}, and the IRNN of~\citeauthor{le2015simple}~\shortcite{le2015simple}. Figure~\ref{fig:RNN} shows the results for each task where the sequence length or the memory duration is $T=100$. \begin{figure} \includegraphics[scale=0.85]{rnn_portrait_roboto.pdf} \caption{We compare different RNN models on two standard long-memory learning tasks, described in~\cite{hochreiter1997long}. The state size for all models was $n=30$, except for the ruRNN~\cite{arjovsky2015unitary}, which had $n=512$ and $n=128$ for the adding and memory tasks respectively; the optimal hyperparameters reported in their work. For our model (guRNN), we use $f=\mathrm{relu}$ and $f=\mathrm{tanh}$ for the nonlinearities in the adding and memory tasks. We compare guRNN with (guRNN$_\beta$) or without (guRNN$_1$) a scaling factor $\beta$ in front of $U$ to compensate for the tendency of the nonlinearity to shrink gradients. We used $\beta = 1.4$ and $\beta = 1.05$. That $\mathrm{relu}$ requires a larger $\beta$ is expected, as this nonlinearity discards more gradient information. Gradient clipping to $[-1, 1]$ was used for the LSTM and IRNN~\cite{le2015simple}. Dotted lines denote random baselines. The learning rate was set to $\alpha=10^{-3}$ for all models except IRNN, which used $\alpha=10^{-4}$. We used RMSProp~\cite{Tieleman2012} with decay 0.9 and no momentum. The batch size was 20.} \label{fig:RNN} \end{figure} While our model guarantees unitarity of $U$, this is \emph{not} sufficient to prevent gradients from vanishing. Consider the norm of the gradient of the cost $C$ with respect to the data at time $\tau$, and use submultiplicativity of the norm to write; \begin{equation*} \left\| \frac{\partial C}{\partial \mathbf{x}_\tau} \right\| \leq \left\| \frac{\partial C}{\partial \mathbf{x}_T} \right\| \left( \prod_{t = \tau}^{T-1} \| f' \left( U \mathbf{h}_t + V \mathbf{x}_t + \mathbf{b}\right) \| \| U \|\right) \left\|\frac{\partial \mathbf{h}_\tau}{\mathbf{x}_\tau} \right\| \end{equation*} where $f'$ is a diagonal matrix giving the derivatives of the nonlinearity. Using a unitary matrix fixes $\| U \| = 1$, but beyond further restrictions (on $V$ and $\mathbf{b}$) does nothing to control the norm of $f'$, which is at \emph{most} 1 for common nonlinearities. Designing a nonlinearity to better preserve gradient norms is beyond the scope of this work, so we simply scaled $U$ by a constant multiplicative factor $\beta$ to counteract the tendency of the nonlinearity to shrink gradients. In Figure~\ref{fig:RNN} we denote this setup by guRNN$_\beta$. Confirming our intuition, this simple modification greatly improves performance on both tasks. Perhaps owing to our efficient gradient calculation (appendix A) and simpler recurrence relation, our model runs faster than that of~\cite{arjovsky2015unitary} (in our implementation), by a factor of 4.8 and 2.6 in the adding and memory tasks shown in Figure~\ref{fig:RNN} respectively. This amounts to the guRNN processing 61.2 and 37.0 examples per second in the two tasks, on a GeForce GTX 1080 GPU. \section{Discussion} \label{section:discussion} Drawing from the rich theory of Lie groups and Lie algebras, we have described a parametrization of unitary matrices appropriate for use in deep learning. This parametrization exploits the Lie group-Lie algebra correspondence through the exponential map to represent unitary matrices in terms of real coefficients relative to a given basis of the Lie algebra $\mathfrak{u}(n)$. As this map from $\mathfrak{u}(n)$ to $U(n)$ is surjective, the parametrization can describe any unitary matrix. We have demonstrated that unitary matrices can be learned with high accuracy using simple gradient descent, and that this approach outperforms a recently-proposed parametrization (from~\citeauthor{arjovsky2015unitary}~\shortcite{arjovsky2015unitary}) and significantly outperforms the approach of `re-unitarizing' after gradient updates. This experimental design is quite simple, designed to probe a core problem, before considering the broader setting of RNNs. Our experiments with general unitary RNNs using this parametrization showed that this approach is practical for deep learning. With a fraction of the parameters, our model outperforms LSTMs on the standard `memory problem' and attains comparable (although inferior) performance on the adding problem~\cite{hochreiter1997long}. Further work is required to understand the difference in performance between our approach and the ruRNN of~\cite{arjovsky2015unitary} - perhaps the $7n$-dimensional subspace captured by their parametrization is serendipitously \emph{beneficial} for these RNN tasks - although we note that the results presented here are not the fruit of exhaustive hyperparameter exploration. Of particular interest is the impressive performance of both uRNNs on the memory task, where the LSTM and IRNN appear to fail to learn. While our RNN experiments have demonstrated the utility of using a unitary operator for these tasks, we believe that the role of the nonlinearity in the vanishing and exploding gradient problem must not be discounted. We have shown that a simple scaling factor can help reduce the vanishing gradient problem induced by the choice of nonlinearity. More analysis considering the combination of nonlinearity and transition operator must be performed to better tackle this problem. The success of our parametrization for unitary operator learning suggests that the approach of performing gradient updates in the Lie algebra is particularly effective. As Lie groups describe many naturally-occuring symmetries, the Lie group-Lie algebra correspondence could be rich for further exploitation to enhance performance in tasks beyond our initial motivation of recurrent neural networks. \section{Appendix A: Derivation of derivative of the matrix exponential} This derivation draws elements from~\citeauthor{kalbfleisch1985analysis}~\shortcite{kalbfleisch1985analysis} and~\citeauthor{jennrich1976fitting}~\shortcite{jennrich1976fitting}. We have $U = \exp(L)$, and seek $dU$. For what follows, we simply require that $L$ be normal, so the results are more general than the unitary case. In this case, $L$ is skew-Hermitian, which is normal and therefore diagonalisable by unitary matrices. Thus, there exist $W \in U(n)$ and $D = \mathrm{diag}(d_1, \dots, d_n)$ such that $L = W D W^{\dagger}$, and therefore \begin{equation} U = W \tilde{D} W^{\dagger} \label{eqn:-1} \end{equation} where $\tilde{D} = \mathrm{diag}(e^{d_1}, \dots, e^{d_n})$. We assume we can calculate: $dL$, $W$, and $D$ and seek an expression for $dU$. Then using~\ref{eqn:-1}: \begin{equation} \begin{split} dU &= d(W \tilde{D} W^{\dagger}) \\ &= dW \tilde{D} W^{\dagger} + W d\tilde{D} W^{\dagger} + W \tilde{D} dW^{\dagger} \end{split} \end{equation} Pre-multiplying with $W^{\dagger}$ and post-multiplying with $W$: \begin{equation} W^{\dagger} dU W = W^{\dagger} dW \tilde{D} + d\tilde{D} + \tilde{D} dW^{\dagger} W \label{eqn:0} \end{equation} The last term can be simplified by differentiating both sides of $W^{\dagger} W = \mathbb{I}$ (this follows from unitarity of $W$); \begin{equation} W^{\dagger} W + W^{\dagger} dW = 0 \Rightarrow dW^{\dagger} W = -W^{\dagger} dW \label{eqn:1} \end{equation} and substituting back into~\ref{eqn:0} to get: \begin{equation} W^{\dagger} dU W = W^{\dagger} dW \tilde{D} - \tilde{D} W^{\dagger} dW + d\tilde{D} \end{equation} We can then say that $dU = W V W^{\dagger}$ where \begin{equation} V = W^{\dagger} dW \tilde{D} - \tilde{D} W^{\dagger} dW + d\tilde{D} \end{equation} Similarly, $dL = W A W^{\dagger}$ where (replacing $\tilde{D}$ with $D$) \begin{equation} A = W^{\dagger} dW D - D W^{\dagger} dW + dD \label{eqn:2} \end{equation} and also $A = W^{\dagger} dL W$. \subsection{Calculating $V$} We use the convention that repeated indices denote summation over that index, unless otherwise stated. Looking at the components of $V$; \begin{equation} V_{ij} = (W^{\dagger} dW \tilde{D})_{ij} - (\tilde{D} W^{\dagger} dW)_{ij} + d\tilde{D}_{ij} \end{equation} \subsubsection{Diagonal case ($i = j$): (no summation over $i$)} \begin{equation} V_{ii} = W^{\dagger}_{ia} dW_{ab} \tilde{D}_{bi} - \tilde{D}_{ia} W^{\dagger}_{ab} dW_{bi} + d\tilde{D}_{ii} \end{equation} Since $\tilde{D}_{bi} = \delta_{bi} \tilde{d}_i$, the first two terms cancel: \begin{equation} \begin{split} V_{ii} =& W^{\dagger}_{ia} dW_{ab} \delta_{bi} \tilde{d}_i - \delta_{ai} \tilde{d}_i W^{\dagger}_{ab} dW_{bi} + d\tilde{D}_{ii}\\ =& W^{\dagger}_{ia} dW_{ai} \tilde{d}_i - \tilde{d}_i W^{\dagger}_{ib} dW_{bi} + d\tilde{D}_{ii}\\ =& d\tilde{D}_{ii} \label{eqn:3} \end{split} \end{equation} Using~\ref{eqn:2} we get $A_{ii} = dD_{ii} = (W^{\dagger} dL W)_{ii}$ Recall that the diagonal elements of $\tilde{D}$ are the exponentiated versions of the diagonal elements of $D$, so $\tilde{D}_{ii} = e^{d_i}$. Then \begin{equation} d\tilde{D}_{ii} = d(d_i) e^{d_i} = dD_{ii} \tilde{D}_{ii} \end{equation} Inserting that into Equation~\ref{eqn:3}: \begin{equation} V_{ii} = dD_{ii} \tilde{D}_{ii} = (W^{\dagger} dL W)_{ii} \tilde{D}_{ii} = (W^{\dagger} dL W)_{ii} e^{d_i} \end{equation} This produces Equation~\ref{eqn:vii} in the main paper. \subsubsection{Off-diagonal case ($i \neq j$): (no summation over $i, j$)} In this case, the purely diagonal part vanishes. We get: \begin{equation} \begin{split} V_{ij} =& W^{\dagger}_{ia} dW_{ab} \delta_{bj} \tilde{d}_j - \delta_{ai} \tilde{d}_i W^{\dagger}_{ab} dW_{bj}\\ =& W^{\dagger}_{ia} dW_{aj} \tilde{d}_j - W^{\dagger}_{ib} dW_{bj} \tilde{d}_i \\ &= (W^{\dagger} dW)_{ij} (\tilde{d}_j - \tilde{d}_i) \end{split} \label{eqn:previj} \end{equation} Similarly, \begin{equation} A_{ij} = (W^{\dagger} dW)_{ij} (d_j - d_i) \end{equation} Remembering that this is all component-wise multiplication (no summation over $i$ and $j$), we can rearrange expressions to get: \begin{equation} (W^{\dagger} dW)_{ij} = \frac{A_{ij}}{d_j - d_i} = \frac{(W^{\dagger} dL W)_{ij}}{d_j - d_i} \end{equation} Combining this with~\ref{eqn:previj} and remembering $\tilde{d}_a = e^{d_a}$, we have, for $i \neq j$: \begin{equation} V_{ij} = (W^{\dagger} dL W)_{ij} \left(\frac{e^{d_i} - e^{d_j}}{d_i - d_j}\right) \end{equation} This is Equation~\ref{eqn:vij} in the main paper. \subsection{Efficiently calculating $W^{\dagger} dL W$} This section is specific to our work, as it relies on the choice of basis for $\mathfrak{u}(n)$. In our case, $dL$ is simple. $L$ is a linear combination of the parameters $\lambda_i$; \begin{equation} L = \sum_i^{n^2} \lambda_i T_i \end{equation} Where $T_i$ are the basis matrices of $\mathfrak{u}(n)$. Then \begin{equation} dL_a = \frac{\partial L}{\partial \lambda_a} = T_a \end{equation} We need $W^{\dagger} T_a W$ for all $a$. Since the $T_a$s are sparse, this is cheaper than performing $n^2$ full matrix multiplications, as we demonstrate now. In components; \begin{equation} (W^{\dagger} T_a W)_{i j} = W^{\dagger}_{i k} {T_a}_{k l} W_{l j} \end{equation} Cases: \subsubsection{$T_a$ diagonal, purely imaginary} $T_a$ is zero except for a $i$ in the $a$-th position on the diagonal. \begin{equation} \begin{split} (W^{\dagger} T_a W)_{i j} &= i W^{\dagger}_{i a} W_{a j} = i W^*_{a i} W_{a j}\\ \Rightarrow W^{\dagger} T_a W &= i \cdot \mathrm{outer}(\mathbf{w}^*_a, \mathbf{w}_a) \end{split} \end{equation} where $\mathbf{w}_a$ is the $a$-th \emph{row} of $W$. \subsubsection{$T_a$ symmetric, purely imaginary} $T_{rs}$ is zero except for $i$ in position $(r, s)$ and $(s, r)$. \begin{equation} \begin{split} (W^{\dagger} T_{rs} W)_{i j} &= i W^{\dagger}_{i k} (\delta_{ks, lr} + \delta_{kr, ls}) W_{l j} \\ &= i (W^{\dagger}_{is} W_{rj} + W^{\dagger}_{ir} W_{s j}) = i ( W^*_{si} W_{rj} + W^*_{ri} W_{sj}) \\ \Rightarrow W^{\dagger} T_{rs} W &= i \cdot (\mathrm{outer}(\mathbf{w}^*_s, \mathbf{w}_r) + \mathrm{outer}(\mathbf{w}^*_r, \mathbf{w}_s)) \end{split} \end{equation} \subsubsection{$T_a$ antisymmetric, purely real} $T_{rs}$ is zero except for $1$ in position $(r, s)$ and $-1$ in position $(s, r)$. \begin{equation} \begin{split} (W^{\dagger} T_{rs} W)_{i j} &= W^{\dagger}_{i k} (\delta_{kr, sl} - \delta_{ks, rl}) W_{l j} \\ &= W^{\dagger}_{ir} W_{sj} - W^{\dagger}_{is} W_{r j} = W^*_{ri} W_{sj} - W^*_{si} W_{rj} \\ \Rightarrow W^{\dagger} T_{rs} W &= \mathrm{outer}(\mathbf{w}^*_r, \mathbf{w}_s) - \mathrm{outer}(\mathbf{w}^*_s, \mathbf{w}_r) \end{split} \end{equation} These reproduce the expressions in the main paper. The outer product of two $n$-dimensional vectors is an $O(n^2)$ operation, and so this provides a (up to) factor $n$ speed-up on matrix multiplication. \section{Appendix B: Changing the basis of $\mathfrak{u}(n)$} \label{section:basis} The Lie group parametrization assumes a fixed basis of $\mathfrak{u}(n)$. Our intuition is that this makes some regions of $U(n)$ more `accessible' to the optimization procedure, elements whose coefficients are small given this basis. Learning a matrix $U$ which came from elsewhere in $U(n)$ may therefore be more challenging. We emulated this `change of basis` without needing to explicitly construct a new basis by generating a change of basis matrix, $M$. That is, if $V_j$ is the $j$-th element of the new basis, it is given by \begin{equation} V_j = \sum_k M_{jk} T_k \end{equation} If $\{\tilde{\lambda}\}_a$ are the coefficients of $L$ relative to the basis $V$, the coefficients relative to the old basis $T$ are given by: \begin{equation} \lambda_b = \sum_k \tilde{\lambda}_k M_{kb} = \tilde{\lambda}^T \cdot M \end{equation} A change of basis matrix must be full-rank. We generate one by sampling a square, $n^2 \times n^2$ matrix from a continuous uniform distribution $\mathcal{U}(-c, c)$ ($c$ is a constant we vary in experiments, see Figure~\ref{fig:basis}). This is very unlikely to be singular. We choose the $c$ range of the distribution such that $M$ will have `large' values relative to the true matrix $U$, whose parameters $\lambda$ (relative to $T$) are drawn from $\mathcal{N}(0, 0.01)$. Preliminary experiments suggested that the learning rate must be adjusted to compensate for the change of scale - evidence for this is visible in the first column of Figure~\ref{fig:basis}, where changing the basis without changing the learning rate results in an unstable validation set trace. Poor performance resulting from an inappropriate learning rate is not our focus here, so we performed experiments for different values of the learning rate. Figure~\ref{fig:basis} shows a grid of validation set losses as we vary the learning rate (columns) and the value of $c$ (rows). Our intuition is that if the performance under the change of basis is \emph{purely} driven by the difference in scale, using an appropriately-scaled learning rate should negate its affect. Each parameter $\lambda_j$ is scaled by a variable uniformly distributed between $(-c, c)$. The expectation value of the absolute value of this quantity is $c^2/2$, so we consider learning rates normalised by this factor. As seen in Figure~\ref{fig:basis}, the graphs on the diagonal are \emph{not} identical, suggesting that merely scaling the learning rate does not account for the change of learning behavior given a new basis - at least in expectation. Nonetheless, it is reassuring to observe that for all choices of $c$ explored, there exists a learning rate which facilitates learning, even if it markedly slower than the `ideal' case. While having a `misspecified' basis does appear to negatively impact learning, it can be largely overcome with choice of learning rate. \begin{figure} \includegraphics[scale=0.5]{basis_inkscape.pdf} \caption{ We consider the effects on learning of changing the basis (rows) and changing the learning rate (columns). For this experiment, $n=6$. The first row uses the original basis. Other rows use change of basis matrices sampled from $\mathcal{U}(-c, c)$ where $c = \{5, 10, 20\}$. The learning rates decrease from the `default' value of 0.001 used in the other experiments. Subsequent values are given by $\frac{0.001}{c^2}$ for the above values of $c$, in an attempt to rescale by the expected absolute value of components of the change of basis matrix. If the change of scale were solely responsible for the change in learning behavior, we would therefore expect the graphs on the diagonal to look the same.} \label{fig:basis} \end{figure} \bibliographystyle{aaai}
d261fc23e334cfefccec7e31d4469dba55b5c91d
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Our purpose is to classify all finite dimensional representations of the Leavitt path algebra $L(\Gamma)$ of a row-finite di(rected )graph $\Gamma$. The main tools we employ are that the category of $L(\Gamma)$-modules is equivalent to a full subcategory of quiver representations of $\Gamma$ satisfying a natural isomorphism condition (Theorem \ref{teorem}) and an explicit Morita equivalence defined by a graph theoretical (reduction) algorithm (Theorem \ref{Morita}).\\ From the quiver representation viewpoint a module $M$ is a functor from the small category $\Gamma$ (whose objects are the vertices in $\Gamma$ and whose morphisms are the paths in $\Gamma$) to the category of $\mathbb{F}$-vector spaces. This functor is an $L(\Gamma)$-module if and only if it satisfies the isomorphism condition (I) of Theorem \ref{teorem}. The reduction algorithm helps to reduce the complexity of the source category $\Gamma$, in particular the cycles relevant for finite dimensional indecomposable modules become loops (a single arrow starting and ending at the same vertex).\\ That the Leavitt path algebra $L(\Gamma)$ of a finite $\Gamma$ has a nonzero finite dimensional representation is equivalent to an easy graph theoretical condition: $\Gamma$ has a \textit{maximal sink} or a \textit{maximal cycle}, that is, a sink or a cycle such that there is no path from any other cycle to it. For Leavitt path algebras this property is properly squeezed between two important algebraic notions: it is implied by the cycles of $\Gamma$ being mutually disjoint (which is equivalent to $L(\Gamma)$ having polynomial growth \cite[Theorem 5]{aajz12} and also to all simple $L(\Gamma)$-modules being finitely presented \cite[Theorem 4.5]{ar14}). In turn, it implies that $L(\Gamma)$ has IBN (Invariant Basis Number), but it is equivalent to neither \cite[Corollary 6.5]{ko1}.\\ Let's describe the classification of finite dimensional $L(\Gamma)$-modules for a finite $\Gamma$ here (rather than a row-finite $\Gamma$) to avoid the annoying technicality of \textit{finitely many predecessors}. Finite dimensional simple $L(\Gamma)$-modules are of two types: (i) projective modules corresponding to maximal sinks; (ii) the others that are in one-to-one correspondence with pairs $(C,f(x))$ where $C$ is a maximal cycle in $\Gamma$ and $f(x) \in \mathbb{F}[x]$ is an irreducible polynomial with $f(0)=1$. When $F$ is algebraically closed, type (ii) is parametrized by a disjoint union of finitely many copies of $\mathbb{F}^\times:=\mathbb{F}\setminus \{0\}$, one for each maximal cycle in $\Gamma$. \\ Finite dimensional nonsimple indecomposable modules correspond to pairs $(C, f(x)^n)$ where $n> 1$, $C$ and $f(x)$ are as above. Hence, if $M$ and $N$ are finite dimensional simples then $Ext(M,N)=0$ unless $M$ and $N$ are isomorphic and of type (ii). Given a maximal sink or a pair $(C,f(x)^n)$, the corresponding indecomposable module and the associated quiver representation can be described explicitly. Once the indecomposables are classified, all finite dimensional representations are classified by Krull-Schmidt. \\ We refer to the excellent survey \cite{abr15} for the history and development of Leavitt path algebras. The precise definition of the Leavitt path algebra $L(\Gamma)$ of a digraph $\Gamma$ will be given in section 2 below where we also collect other definitions and facts that will be needed. In particular the concepts of the support subgraph, maximal sinks, maximal cycles and the theorem stating the equivalence of the category of unital $L(\Gamma)$-modules with a full subcategory of quiver representations of $\Gamma$ \cite[Theorem 3.2]{ko1} are relevant for our classification. \\ Section 3 is about a graph theoretic process we call the \textit{reduction algorithm} (defined on a row-finite digraph $\Gamma$) and its consequences. Reduction leaves the nonstable $K$-theory (i.e., the monoid $\mathcal{V}(L(\Gamma))$ of isomorphism classes of finitely generated projective $L(\Gamma)$-modules under direct sum) invariant. This enables us to show that dimension functions on a finite digraph form a free commutative monoid on the maximal sinks and cycles of $\Gamma$ (Theorem \ref{27}). The monoid of dimension functions also turn out to be isomorphic to the monoid homomorphisms from $\mathcal{V}(L(\Gamma))$ to the natural numbers $\mathbb{N}$ under addition.\\ Section 4 starts with Theorem \ref{Morita} showing that the Leavitt path algebras of a digraph and any reduction of it are Morita equivalent (which is given explicitly in terms of the corresponding quiver representations). This generalizes the fact that \textit{source elimination} is Morita invariant \cite[Proposition 10]{abr15}. We classify all finite dimensional representations by first reducing the problem to the (finite) support subgraph, then applying the reduction algorithm and finally pulling back to the Leavitt path algebra via the Morita equivalence of Theorem \ref{Morita}.\\ The category of finite dimensional $L(\Gamma)$-modules, for a row-finite digraph $\Gamma$, is equivalent to the direct sum of the categories of finite dimensional vector spaces indexed by the maximal sinks with finitely many predecessors and the categories of finite dimensional modules over the algebra of Laurent polynomials indexed by the maximal cycles with finitely many predecessors (Theorem \ref{notasyon}). In particular the category of finite dimensional representations of the Leavitt path algebra $L(\Gamma)$ of a finite digraph $\Gamma$ is tame in contrast to the finite dimensional quiver representations of $\Gamma$ which are wild unless the connected components of the undirected $\Gamma$ are Dynkin diagrams of types $A, \> D, \> E, \> \tilde{A}, \> \tilde{D}, \> \tilde{E}$ by a celebrated theorem of Gabriel \cite{dw05}.\\ \section{Preliminaries} \medskip \subsection{Leavitt Path Algebras} \medskip A \textbf{di}(rected )\textbf{graph} $\Gamma$ is a four-tuple $(V,E,s,t)$ where $V$ is the set of vertices, $E$ is the set of arrows (directed edges), $s$ and $t:E \longrightarrow V$ are the source and the target functions. The digraph $\Gamma$ is \textit{finite} if $E$ and $V$ are both finite. $\Gamma$ is \textit{row-finite} if $s^{-1}(v)$ is finite for all $v$ in $V$. Given $V' \subseteq V$ the \textit{induced subgraph} on $V'$ is $\Gamma':= (V',E',s',t')$ with $E':= s^{-1}(V') \cap t^{-1}(V')$ ; $\> s':= s\vert_{E'}$ ; $\> t':= t \vert_{E'}$. A subgraph is \textit{full} if it is the induced subgraph on its vertices. \begin{remark} A digraph is also called an "oriented graph" in graph theory, a "diagram" in topology and category theory, a "quiver" in representation theory, usually just a "graph" in $C^*$-algebras and Leavitt path algebras. The notation above for a digraph is standard in graph theory. However $Q=(Q^0,Q^1, s,t)$ is more common in quiver representations while $E=(E^0, E^1,s,r)$ is mostly used in graph $C^*$-algebras and in Leavitt path algebras. We prefer the graph theory notation which involves two more letters but no superscripts. As in quiver representations we view $\Gamma$ as a small category, so "arrow" is preferable to "edge", similarly for "target" versus "range". \end{remark} A vertex $v$ in $V$ is a \textit{sink} if $s^{-1}(v)= \emptyset$; it is a \textit{source} if $t^{-1}(v)= \emptyset$. An \textit{isolated vertex} is both a source and a sink. If $t(e)=s(e)$ then $e$ is a \textit{loop}. A \textbf{pseudo-source} is a vertex $v$ where $t^{-1}(v)$ consists of a single loop. A \textit{path} of length $n>0$ is a sequence $p =e_{1}\ldots e_{n}$ such that $t(e_{i})=s(e_{i+1})$ for $i=1,\ldots ,n-1.$ The source of $p$ is $s(p ):=s(e_{1})$ and the target of $p$ is $t(p ):=t(e_{n})$. A path $p$ of length 0 consists of a single vertex $v$ where $s(p) :=v $ and $t(p) := v$. We will denote the length of $p$ by $l(p)$. A path $C=e_1e_2 \cdots e_n$ with $n>0$ is a \textit{cycle} if $s(C )=t(C )$ and $s(e_{i})\neq s(e_{j})$ for $i\neq j$. An arrow $e \in E$ is an \textit{exit} of the cycle $C=e_1e_2 \cdots e_n$ if there is an $i$ such that $ s(e)=s(e_i)$ but $e\neq e_i$. The digraph $\Gamma$ is \textit{acyclic} if it has no cycles. \\ There is a preorder defined on the set of sinks and cycles in $\Gamma$: we say that a cycle $C$ \textbf{connects to} a sink $w$ denoted by $C \leadsto w$ if there is a path from $C$ to $w$. Similarly $C\leadsto D$ if there is a path from the cycle C to the cycle D. This is a partial order if and only if the cycles in $\Gamma$ are mutually disjoint. A cycle is \textbf{minimal} with respect to $\leadsto$ if and only if it has no exit (sinks are always minimal). A cycle $C$ is \textbf{maximal} if no other cycle connects to $C$ (in particular, a maximal cycle is disjoint from all other cycles). A sink $w$ is maximal if there is no cycle $C$ which connects to $w$. \\ The \textbf{graph monoid $S(\Gamma)$} is the (additive) commutative monoid generated by $V$ subject to the relations: $v =\sum_{e \in s^{-1}(v)} te$ for all $ v \in V $ with $ 0< \vert s^{-1}(v) \vert < \infty$.\\ Given a digraph $\Gamma,$ the \textit{extended digraph} of $\Gamma$ is $\tilde{\Gamma} := (V,E \sqcup E^*, s~,t~)$ where $E^* :=\{e^*~|~ e\in E \}$, the functions $s$ and $t$ are extended as $s(e^{\ast}):=t(e)$ and $t(e^{\ast }):=s(e) $ for all $e \in E$. Thus the dual arrow $e^*$ has the opposite orientation of $e$. We want to extend $*$ to an operator defined on all paths of $\tilde{\Gamma}$: Let $v^*:=v$ for all $v$ in $V$, $(e^*)^*:=e $ for all $e$ in $E$ and $p^*:= e_n^* \ldots e_1^*$ for a path $p=e_1 \ldots e_n$ with $e_1, \ldots , e_n $ in $E \sqcup E^* $. In particular $*$ is an involution, i.e., $**=id$. \\ The \textbf{ Leavitt path algebra} of a digraph $\Gamma$ with coefficients in the field $\mathbb{F}$, as defined in \cite{aa05} and \cite{amp07}, is the $\mathbb{F}$-algebra $L_{\mathbb{F}}(\Gamma)$ generated by $V \sqcup E \sqcup E^*$ satisfying:\\ \indent(V) $\quad \quad vw ~=~ \delta_{v,w}v$ for all $v, w \in V ,$ \\ \indent($E$) $\quad \quad s ( e ) e = e=e t( e) $ for all $e \in E\sqcup E^*$, \\ \indent(CK1) $\quad e^*f ~ = ~ \delta_{e,f} ~t(e)$ for all $e,f\in E$, \\ \indent(CK2) $\quad v~ = ~ \sum_{s(e)=v} ee^*$ for all $v$ with $0< \vert s^{-1}(v) \vert < \infty $.\\ We will usually suppress the subscript $\mathbb{F}$ when we denote our algebras. If the digraph $\Gamma$ is fixed and clear from the context we may abbreviate $L(\Gamma)$ to $L$. From now on we will omit the parentheses to reduce notational clutter when the source and target functions $s$, $t$ are applied. \\ The relations (V) simply state that the vertices are mutually orthogonal idempotents. The relations (E) say that $e \in seLte$ and $e^* \in teLse$ for every $e \in E$. The Leavitt path algebra $L$ as a vector space is $\bigoplus vLw$ where the sum is over all pairs $(v,w) \in V \times V$ since $V\sqcup E\sqcup E^*$ generates $L$. If we only impose the relations (V) and ($E$) then we obtain $\mathbb{F}\tilde{\Gamma}$, the \textbf{path }(or \textbf{quiver}) \textbf{algebra} of the extended digraph $\tilde{\Gamma}$ : The paths in $\tilde{\Gamma}$ form a vector space basis of $\mathbb{F} \tilde{\Gamma}$, the product $pq$ of two paths $p$ and $q$ is their concatenation if $tp=sq$ and 0 otherwise. \\ We get the\textit{ Cohn path algebra} $C(\Gamma)$ when we impose the relations (CK1) in addition to (V) and ($E$). Hence $L(\Gamma)$ is a quotient of $C(\Gamma)$, which is a quotient of $ \mathbb{F} \tilde{\Gamma}$. The abbreviation CK stands for Cuntz-Krieger. \\ For any arrow $e$ in $E$ we have $e^*e=te$ by (CK1). Consequently $p^*p=tp$ for any path $p$ of $\Gamma$. Hence for any two paths $p$ and $q$ of $\Gamma$ if $q=pr$ then $p^*q=p^*pr=r$, if $p=qr$ then $p^*q=(q^*p)^*=r^*$. As $e^*f=0$ when $e\neq f$ by (CK1), $p^*q=0$ unless the path $q$ is an initial segment of the path $p$ (i.e., $p=qr$) or $p$ is an initial segment of $q$ (i.e., $q=pr$). Thus the Cohn path algebra $C(\Gamma)$ and the Leavitt path algebra $L(\Gamma)$ are spanned by $\left\{pq^*\right\}$ where $p$ and $q$ are paths of $\Gamma$ with $tp=tq$. In fact this is a basis for $C(\Gamma)$ which can be shown by defining an epimorphism from $C(\Gamma)$ to a reduced semigroup algebra with this basis (we will not need this fact). In $L(\Gamma)$ however $\{pq^* : tp=tq \}$ is linearly dependent because of (CK2).\\ The algebras $ \mathbb{F}\Gamma$, $ \mathbb{F}\tilde{\Gamma}$, $C(\Gamma)$ and $L(\Gamma)$ are unital if and only if $V$ is finite, in which case the sum of all the vertices is the unit: It is clear that $\sum_{v \in V} v=1$ when $V$ is finite. For the converse, a given element in any these algebras is a finite linear combination of paths in $\tilde{\Gamma}$ and we can pick $v \in V$ which is not the source of any of these paths if $V$ is infinite. Now left multiplication by $v$ gives zero, so there is no unit in any of these algebras since $\mathbb{F}\Gamma$ is a subalgebra of $L(\Gamma)$ by \cite[Lemma 1.6]{goo09} and thus $v\neq 0$ in $L(\Gamma)$ for every $v \in V$, hence also in $C(\Gamma)$.\\ We will need the following easy and well-known fact: \begin{lemma} \label{ez} If $v=sC$ where $C$ is a cycle without exits in $\Gamma$ then $vL(\Gamma)v \cong \mathbb{F}[x,x^{-1}]$ via $x \leftrightarrow C^*$.\end{lemma} \begin{proof} Since $C$ has no exit $ee^*=se$ by (CK2) for any arrow $e$ on $C$. So $CC^*=v$ and also $C^*C=v$ by (CK1). Since $v=1$ in $vL(\Gamma)v$ we get $C^{-1}=C^*$ and a homomorphism from $\mathbb{F}[x,x^{-1}]$ to $vL(\Gamma)v$ is defined by sending $x$ to $C^*$. This homomorphism is onto: $vL(\Gamma)v$ is spanned by $\{ pq^* \> \vert \> sp=v=sq , \> tp=tq\}$. If $tp\neq v$ then $tp=ee^*$ by (CK2) where $e$ is the unique arrow with $se=tp$. So $pq^*=pe(qe)^*$ and we can repeat as needed to get $pq^*=C^mC^{*n}=C^{m-n}$ for some $m$ and $n$ in $\mathbb{N}$. This homomorphism is also one-to-one because $\{C^n\}_{n \geq 1}$ is linearly independent in the path algebra $\mathbb{F}\Gamma$, hence also in $vL(\Gamma)v \subseteq L(\Gamma)$. \end{proof}\\ There is a $\mathbb{Z}$-grading on $ \mathbb{F} \tilde{\Gamma}$ given by $deg(v)=0$ for $v$ in $V$, $deg(e)=1$ and $deg(e^*)=-1$ for $e$ in $E$. This defines a grading on all our algebras since the relations are all homogeneous. The linear extension of $*$ on paths induces a grade-reversing involutive anti-automorphism (i.e., $deg(\alpha^*)=-deg(\alpha)$ and $(\alpha \beta)^*=\beta^* \alpha^*$). Hence these algebras are $\mathbb{Z}$-graded $*$-algebras and the (graded) categories of left modules and right modules are equivalent for any of these algebras.\\ A subset $H$ of $V$ is \textit{hereditary} if for any path $p$, $sp \in H$ implies that $tp \in H$; $H$ is \textit{saturated} if $\{ te : se=v \} \subseteq H$ implies that $v \in H$, for every $v \in V$ with $0<\vert s^{-1}(v) \vert < \infty $ \cite{aa05}. If $I$ is an ideal of $L(\Gamma)$ and $p$ is a path in $\Gamma$ with $sp \in I$ then $tp=p^*p=p^*(sp)p \in I$, also if $\{ te : se=v \} \subseteq I$ then $v=\sum_{se=v} ee^*=\sum_{se=v} e(te)e^* \in I$, so $I\cap V$ is hereditary and saturated. We have a Galois connection between the subsets of $V$ and the ideals of $L(\Gamma)$ given by $S \mapsto (S)$ and $I \mapsto I\cap V$ which gives a bijection between hereditary saturated subsets of $V$ and graded ideals of $L(\Gamma)$ when $\Gamma$ is a row-finite digraph \cite[Theorem 5.3]{amp07}. \medskip \subsection{$L(\Gamma)$-Modules and Quiver Representations} \medskip As each $v$ in $V$ is an idempotent, $vL$ is a cyclic projective $L$-module. We have a homomorphism $\phi_v: \bigoplus_{se=v} teL\longrightarrow vL$ sending $(\alpha_e)$ to $ \sum e\alpha_e$ for all $v$ with $0< \vert s^{-1}(v) \vert < \infty $. The relations (CK1) and (CK2) imply that this homomorphism is an isomorphism whose inverse sends $\beta$ in $vL$ to $(e^*\beta)$ in $\bigoplus teL$. \\ In fact, $L(\Gamma)$ can be defined as the Cohn localization of the path algebra $\mathbb{F}\Gamma$, without mentioning $E^*$, (CK1) or (CK2), making the homomorphisms analogous to the $\phi_v$ above from $\bigoplus te\mathbb{F}\Gamma$ to $v\mathbb{F}\Gamma$ invertible \cite[Corollary 4.2]{ab10}, \cite[Proposition 3.3]{ko1}.\\ The isomorphisms $\phi_v$ also enable us to define a monoid homomorphism from the graph monoid $S(\Gamma)$ to $\mathcal{V} (L)$, the nonstable K-Theory of $L$, by sending $[v]$ to $[vL]$. A deep and important result in the subject is: \begin{theorem} (\cite[Theorem 3.5]{amp07}) The monoid homomorphism from $S(\Gamma)$ to $\mathcal{V} (L(\Gamma))$ defined above is an isomorphism. \end{theorem} We will work in the category $\mathfrak{M}_{L }$ of unital (right) modules over $L:=L_{\mathbb{F}} (\Gamma)$. However $L$ has a 1 if and only if the vertex set $V$ is finite. Even if $V$ is infinite, we define a unital $L$-module as a module $M$ with the property that $ML=M$, i.e., for any $m$ in $M$ we can find $\lambda_1, \lambda_2, \ldots , \lambda_n$ in $L$ and $m_1, m_2, \ldots, m_n$ in $M$ so that $m=m_1\lambda_1+m_2 \lambda_2 +\cdots + m_n\lambda_n$. This condition is equivalent to the standard definition of unital (when $L$ has a 1) since $m1=(m_1\lambda_1+m_2 \lambda_2 +\cdots + m_n\lambda_n)1=m_1\lambda_1 1 +m_2 \lambda_2 1+\cdots + m_n\lambda_n1=m_1\lambda_1+m_2 \lambda_2 +\cdots + m_n\lambda_n=m$. The projective modules $vL$ for all $v \in V$ are unital. The category of unital modules is an abelian category with sums since it is closed under taking quotients, submodules, extensions, (arbitrary) sums (but not infinite products: if $V$ is infinite then the $L$-module $L^V$ is not unital).\\ For any $M$ in $\mathfrak{M}_L$ and any $v$ in $V$ the linear map sending $f \in Hom_L (vL, M)$ to $f(v) \in Mv$ gives an isomorphism. Applying the cofunctor $Hom_L( \>\underline{ \> \> \> } \>, M)$ to $\phi_v$ we get an isomorphism $Mv \longrightarrow \bigoplus Mte$. Thus, assigning the vector space $Mv$ to each vertex $v$ and the linear transformation from $Mse$ to $Mte$ given by right multiplication by $e$ for all $e \in E$ defines a quiver representation of $\Gamma$ satisfying the isomorphism conditions imposed by the $\{\phi_v \}$. \\ Recall that a quiver representation is a functor from the small category $\Gamma$ with objects $V$ and morphisms given by paths in $\Gamma$. The linear transformation assigned to the path $p$ from $Msp$ to $Mtp$ is given by right multiplication by $p$. Conversely, any quiver representation of $\Gamma$ satisfying the isomorphism conditions above yields an $M $ in $\mathfrak{M}_L$: \begin{theorem} \label{teorem} \cite[Theorem 3.2]{ko1} If $\Gamma= (V,E,s,t)$ is a row-finite digraph then the category $\mathfrak{M}_{L}$ is equivalent to the full subcategory of quiver representations $\rho$ of $\Gamma$ satisfying the following condition $(I)$:\\ $$\textit{For every nonsink } v\in V, \> \> \> \> (\rho(e))_{se=v}: \rho ( v) \longrightarrow \bigoplus\limits_{se=v } \rho(te) \textit{ is an isomorphism.}$$ The module $M$ corresponding to the quiver representation $\rho$ is $M=\bigoplus_{v\in V} \rho(v)$. \end{theorem} With the quiver representation viewpoint there is no need to mention the generators $\{e^* : e \in E \}$ explicitly, they are implicit in the condition (I). Below we will frequently define modules by constructing the corresponding quiver representations.\\ A \textbf{dimension function} on a digraph $\Gamma$ is a function $d: V \longrightarrow \mathbb{N}$ satisfying: $d(v) =\sum_{se=v} d(te)$ for all $ v \in V $ with $ 0< \vert s^{-1}(v) \vert < \infty$. Hence dimension functions on $\Gamma$ correspond exactly to monoid homomorphisms from the graph monoid $S(\Gamma)\cong \mathcal{V} (L)$ to the additive monoid of natural numbers $\mathbb{N}$.\\ $M \in \mathfrak{M}_L$ is of \textit{finite type} if $dim^{\mathbb{F}} (Mv)< \infty$ for all $v \in V$. Then $d(v):=dim^{\mathbb{F}}(Mv)$ is a dimension function by condition (I). When $V$ is finite, finite type is the same as finite dimensional since $M =\oplus_{v \in V} Mv$ by Theorem \ref{teorem}. Theorem \ref{teorem} also implies that any dimension function can in fact be realized by an $L$-module \cite[Corollary 3.7]{ko1}. \medskip \subsection{Support Subgraph} \medskip For any $M \in \mathfrak{M}_L$ the \textbf{support subgraph} of $M$, denoted by $\Gamma_{_M}$, is the induced subgraph of $\Gamma$ on $V_{_M}:=\{v \in V \mid Mv \neq 0 \}$. It's easy to check that $V\setminus V_{_M}= \{ v \in V \> \vert \> Mv=0 \}$ is hereditary and saturated, hence the ideal $I_{_M}$ generated by $V\setminus V_{_M}$ is the largest graded ideal of $L(\Gamma)$ contained in $AnnM$. Conversely, $V\setminus H$ for any hereditary saturated subset $H$ of $V$ is the support subgraph of $L(\Gamma) / (H)$ regarded as an $L(\Gamma)$-module. Also $L(\Gamma_{_M})\cong L(\Gamma)/I_{_M}$ and we may regard $M$ as an $L(\Gamma_{_M})$-module whose $L(\Gamma)$-module structure is induced by the projection from $L(\Gamma)$ to $L(\Gamma_{_M})$. As an $L(\Gamma_M)$-module, $M$ has full support, that is, $Mv \neq 0$ for all $v \in V_M$ \cite[Section 5]{ko1}. \begin{lemma} \label{exits} \cite[Lemma 6.1]{ko1} If $M\in \mathfrak{M}_{L(\Gamma)}$ is of finite type then the cycles of its support subgraph $\Gamma_{_M}$ have no exits. \end{lemma} \begin{proof} If $v_1,..., v_n, v_{n+1}=v_1$ are consecutive vertices in a cycle of $\Gamma_{_M}$ then $dim(Mv_1) \geq dim(Mv_2) \geq \cdots \geq dim(Mv_n)\geq dim( Mv_1)$ by Theorem \ref{teorem}. Hence $dim(Mv_k) =dim(Mv_{k+1})$ for $k=1, \cdots, n$. It follows from Thoerem \ref{teorem} again that $Mte=0$ for $e\in s^{-1}(v_k)$ unless $te=v_{k+1}$. Thus the cycles of $\Gamma_{_M}$ have no exits. \end{proof}\\ For any two vertices $v$ and $w$ in $V$ we write $v \leadsto w$ if there is a path $p$ in $\Gamma$ from $v$ to $w$. This defines a preorder (a reflexive and transitive relation). If $v$ and $w$ are on a cycle then $v \leadsto w$ and $w \leadsto v$. Let $U$ be the set of sinks and cycles of $\Gamma$. There is an induced preorder on $U$, also denoted by $\leadsto$. (This is a partial order on $U$ if and only if the cycles of $\Gamma$ are disjoint, equivalently when $L(\Gamma)$ has polynomial growth \cite[Theorem 5]{aajz12}.) A sink or a cycle $u \in U$ is \textbf{maximal} if $u' \leadsto u$ only if $u' =u$. \\ The \textbf{predecessors} of $v$ in $V$ is $V_{\leadsto v} := \{ w \in V \mid w \leadsto v\} $. If $u$ and $w$ are two vertices on a cycle $C$ then they have the same predecessors, so $V_{\leadsto C} $ is well-defined. The \textbf{ predecessors subgraph} $\Gamma_{\leadsto v}$ is the induced subgraph on $V_{\leadsto v}$, similarly $\Gamma_{\leadsto C} := \Gamma_{\leadsto v}$ where $v$ is any vertex on $C$. \section{The Reduction Algorithm} This section is about the consequences of a geometric (graph theoretic) process we call the \textbf{reduction algorithm} defined on a row-finite digraph $\Gamma=(V,E)$: For a loopless nonsink $v \in V$, we replace each path $fg$ of length 2 such that $tf=v=sg$ with an arrow labeled $fg$ from $sf$ to $tg$ and delete $v$ and all arrows touching $v$. (Note that $fg$ denotes a path in $\Gamma$, but an arrow in its reduction.) In particular, if $v$ is a source but not a sink, then we delete $v$ and all arrows starting at $v$ without adding any new arrows. We may repeat this as long as there is a loopless non-sink. Any digraph obtained during this process is called a \textbf{reduction} of $\Gamma$. If $\Gamma$ is finite then after finitely many steps we will reach a \textbf{complete reduction} of $\Gamma$, which has no loopless nonsinks. A digraph in which every vertex is either a sink or has a loop is called \textbf{reduced}.\\ In the example below, $\Gamma_1$ and $\Gamma_2$ are reductions of the digraph $\Gamma_0$. The number of arrows from one vertex to another is indicated by the number above the arrow (so there are 3 arrows from $w$ to $x$). There are two reduction steps going from $\Gamma_1$ to $\Gamma_2$ and $\Gamma_2$ is a complete reduction of $\Gamma_0$. \begin{example} $$ \Gamma_0 \quad \quad \xymatrix{ & {\bullet}^{w} \ar@/^1pc/[r]^{3} & {\bullet}^{x} \ar@/^1pc/[l]^{2} \\ {\bullet}^{u} \ar@{->}[ur] \ar@{->}[dr]^{2} & \\ & \bullet^{v} \ar@{->}[r] & \bullet^{y} }$$ \bigskip $$\Gamma_1 \quad \xymatrix{ & {\bullet}^{w} \ar@/^1pc/[r]^{3} & {\bullet}^{x} \ar@/^1pc/[l]^{2} \\ & \bullet^{v} \ar@{->}[r] & \bullet^{y} }$$ \bigskip $$ \Gamma_2 \quad \xymatrix{ & {\bullet}^{w} \ar@(ur,dr)^{6} &\\ & \bullet^{y} }$$ \end{example} \bigskip \begin{remark} The reduction algorithm can be described in terms of the "six graph moves" originating in symbolic dynamics \cite{alps11}, \cite[Appendix 3]{abr15}: When the vertex $v$ to be eliminated is not a source, we perform an "in-split" where $t^{-1}(v)$ is partitioned so that each part consists of a single arrow. Then we perform a "contraction" on each one of these arrows. Under the additional hypothesis of $L(\Gamma)$ being purely infinite simple all six graph moves yield Morita equivalences \cite{alps11}, \cite[Proposition 10]{abr15}. In Theorem \ref{Morita} below we prove that the reduction algorithm yields a Morita equivalence without any hypothesis on $L(\Gamma)$. \\ When $v$ is a source, reduction is exactly the same as "source elimination" which is known to give a Morita equivalence \cite[Proposition 10]{abr15}. Source elimination can also be obtained as a composition of an in-split partitioning all arrows $s^{-1}(v)$ in the singletons followed by contractions of all these arrows. \end{remark} If $\Gamma'$ is a one step reduction eliminating the vertex $v$ of $\Gamma$ then there is a monomorphism $ \iota$ from $L(\Gamma')$ to $L(\Gamma)$ defined as: $\iota(w)=w$ for all $w \in V'$, $\iota (e)=e $ and $\iota (e^*)= e^*$ if $e \in E' \cap E$, finally $\iota (fg)=fg \> , \> \iota( (fg)^*)=g^*f^*$ if $fg$ is a new arrow. (Again the input $fg$ is an arrow of $\Gamma'$ while the output $fg$ is a path of length 2 in $\Gamma$.) This defines a homomorphism since the relations are satisfied. In fact $\iota$ is a graded homomorphism where $L(\Gamma')$ has the standard $\mathbb{Z}$-grading ($deg(w)=0 $ for all $w \in V' \> ,\> deg(e)=1 \> , \> deg(e^*)=-1$ for all $e \in E'$) but the grading in $L(\Gamma)$ is defined as $deg(w)=0 $ for all $w \in V$, $deg(e)=1, deg(e^*)=-1$ if $se\neq v$ and $deg(g)=deg(g^*)=0$ if $sg=v$. A graded ideal is generated by the vertices it contains (as a consequence of the one-to-one correspondence between graded ideals and hereditary, saturated subsets of vertices). Hence the graded ideal $Ker (\iota)$ is trivial since it contains no vertices, thus $\iota$ is one-to-one. \\ We may identify $L(\Gamma')$ with $Im(\iota)=L':=\{ \alpha \in L(\Gamma) \mid v\alpha=0=\alpha v \} \> $: Clearly $Im(\iota) \subseteq L'$. Conversely, $L'$ is spanned by elements of the form $pq^*$ where $p$ and $q$ are paths in $\Gamma$ with $tp=tq$ and $sp\neq v \neq sq $. If $tp=v$ then we can use $v=\sum_{g \in s^{-1}(v)} gg^*$ to express $pq^*=pvq^*$ as an element of $Im(\iota)$ thus $L'=Im(\iota)$.\\ If $\Gamma'$ is a reduction of $\Gamma$ then $\Gamma$ and $\Gamma'$ have the same set of sinks. A cycle may get shorter under reduction but it can not disappear. However, the number of cycles may increase as illustrated in the following example where the number of cycles increases from 2 to 3: $$\Gamma : \xymatrix{& {\bullet}^{u} \ar@/^1pc/[r] &\ar [l] {\bullet}^{v} \ar [r] & {\bullet}^{w} \ar@/^1pc/[l] }$$\\ $$\Gamma' : \qquad \xymatrix{ & {\bullet}^{u} \ar@(ul,dl)\ar@/^1pc/[r] & {\bullet}^{w} \ar@/^1pc/[l] \ar@(ur,dr)} \qquad $$\\ The digraph $\Gamma$ above also shows that complete reductions are not unique. If we chose to eliminate the vertices $u$ and $w$ (instead of $v$ as above) we obtain the rose with 2 petals which is not isomorphic to the $\Gamma'$ above. \\ However, if the cycles of a finite digraph $\Gamma$ are disjoint (that is, the Gelfand-Kirillov dimension of $L(\Gamma)$ is finite \cite[Theorem 5]{aajz12}) then the number of cycles does not change under reduction since each eliminated vertex effects at most one cycle. In fact, all complete reductions of $\Gamma$ are isomorphic when the cycles of $\Gamma$ are disjoint. \begin{proposition} If $\Gamma$ is a finite digraph then $L(\Gamma)$ is finite dimensional if and only if any complete reduction of $\Gamma$ consists only of isolated vertices. In this case $L(\Gamma)$ is isomorphic to a direct sum of matrix algebras and the number of summands equals the number of (isolated) vertices of a complete reduction of $\Gamma$. \end{proposition} \begin{proof} $L(\Gamma)$ is finite dimensional if and only if $\Gamma $ is finite and has no directed cycles, moreover $L(\Gamma)$ is isomorphic to a direct sum of matrix algebras with as many summands as the number of sinks of $\Gamma$ \cite[Corollaries 3.6 and 3.7]{aas07}. A closed path in a reduction of $\Gamma$ comes from a closed path in $\Gamma$. Hence any complete reduction of a finite acyclic $\Gamma$ will only have isolated vertices, corresponding to the sinks of $\Gamma$. Conversely, if any complete reduction of $\Gamma$ consists of isolated vertices then $\Gamma$ could not have had any cycles because a reduction may shorten a cycle, but it can not get rid of it. \end{proof} \\ Reduction preserves the monoid of $\Gamma$ enabling us to classify all dimension functions. \begin{lemma} If $\Gamma^{'}$ is a reduction of the digraph $\Gamma$, then $S(\Gamma) \cong S(\Gamma^{'}) $. \end{lemma} \begin {proof} It suffices to show that $S(\Gamma) \cong S(\Gamma^{'}) $ when $\Gamma^{'}$ is a one step reduction of $\Gamma$. If $v$ is the deleted loopless nonsink, then $V^{'} =V \setminus \left\{v \right\} $ is the vertex set of $\Gamma^{'}$. Let $\Phi :S(\Gamma^{'}) \longrightarrow S(\Gamma) $ be induced by the inclusion of $V^{'}$ into $V$. In $S(\Gamma^{'})$ $$ u=\sum_{se=u , ~ te \neq v} t(e) + \sum_{se=u, ~te=v} \left( \sum_{sf=v} tf \right) \quad \forall \> u \in V'$$ by the definition of $\Gamma^{'}$. (Note that $sf=v$ implies $tf \neq v$ since $v$ is loopless.) Since, in $S(\Gamma)$ we have $ u=\sum_{se=u} te$ and $v=\sum_{sf=v} tf$, $\Phi$ is a well-defined semigroup homomorphism. Now, $\Psi (u) =u $ if $u\neq v$ and $\Psi (v) =\sum_{sf=v} tf$ similarly defines $\Psi: S(\Gamma) \longrightarrow S(\Gamma^{'})$, which is the inverse of $\Phi$. \end{proof} \bigskip A sink $w$ becomes an isolated vertex in a complete reduction of finite $\Gamma$ if and only if $w$ is maximal (there are no cycles in $\Gamma$ from which there is a path to $w$). A cycle is maximal in $\Gamma$ if and only if it becomes a loop at a pseudo-source in a complete reduction. \\ \begin{lemma} \label{nonzero} If $\Gamma$ is a reduced digraph and $ d:V \rightarrow \mathbb{N}$ is a dimension function then vertices with $d(v)>0 $ are either isolated vertices or pseudo-sources. \end{lemma} \begin{proof} Let $T$ be the set of isolated vertices and pseudo-sources of $\Gamma$. If $v \notin T$ then there are at least two loops at $v$ or there is an arrow $e$ such that $te=v$. \\ If there are two loops $f, g$ at $v$ then $d(v) = \sum_{s(e)=v} d(te) \geq 2d(v)$. Hence we get $d(v)=0$. \\ If there is an arrow $e$ such that $te=v$ then $se$ is not a sink. Hence there exists a loop $h$ at $se$ since $\Gamma$ is a reduced digraph. Then $d(v)=0$ because $d(se)\geq d(se)+ d(te)= d(se)+d(v)$. \end{proof} \begin{lemma} \label{psource} Let $\Gamma$ be a reduced digraph. A dimension function is obtained by assigning arbitrary values to the isolated vertices and the pseudo-sources of $\Gamma$. Hence the monoid of homomorphisms from $S(\Gamma )$ to $\mathbb{N}$ is isomorphic to the free commutative monoid generated by the isolated vertices and the pseudo-sources of $\Gamma$. \end{lemma} \begin{proof} We will check that any assignment of arbitrary natural numbers to pseudo-sources and isolated vertices will define a dimension function (where $d(v):=0$ for all the remaining vertices). There are no relations to check for sinks (including isolated vertices). If $v$ is a pseudo-source then there is a unique loop at $v$ and for any other arrow $e$ with $se=v$, the vertex $te$ is neither a pseudo-source nor an isolated vertex. Thus $d(te)=0$ by definition. Hence the relation at $v$ is $d(v)=d(v)+0$, which holds for any choice of $d(v)$. \\ If $v$ is neither a sink nor a pseudo-source then $d(v)=0$ by definition and for any $e$ with $se=v$ its target $te$ is neither isolated nor a pseudo-source. Hence the relation at $v$ is $0=0$. Therefore any choice of natural numbers for pseudo-sources and isolated vertices (with the remaining vertices of the reduced digraph sent to 0) yields a homomorphism from $S(\Gamma)$ to $\mathbb{N}$. This gives an isomorphism from the free commutative monoid on the isolated vertices and the pseudo-sources of $\Gamma$ to the monoid of homomorphisms from $S(\Gamma)$ to $\mathbb{N}$ under addition. \end{proof} \\ \begin{theorem} \label{27} Let $\Gamma$ be a finite digraph. A dimension function is obtained by assigning arbitrary natural numbers to the maximal sinks and the maximal cycles of $\Gamma$. Hence the monoid of homomorphisms from $S(\Gamma) \cong \mathcal{V} (L(\Gamma))$ to $\mathbb{N}$ is isomorphic to the free commutative monoid generated by the maximal sinks and the maximal cycles of $\Gamma$. \end{theorem} \begin{proof} If $\Gamma'$ is a complete reduction of $\Gamma$ then there are one-to-one correspondences between the isolated vertices (respectively, the pseudo-sources) of $\Gamma'$ and the maximal sinks (respectively, the maximal cycles) of $\Gamma$: Under reduction the maximal sinks or cycles remain maximal because the new cycles created (if any) do not involve vertices from which a maximal sink or cycle can be reached. Now Lemma \ref{psource} gives the desired conclusion. \end{proof}\\ In particular, although a finite digraph may not have a unique complete reduction, all such must have the same number of isolated vertices and pseudo-sources.\\ Theorem \ref{27} which classifies all dimension functions of a finite digraph $\Gamma$ is essentially a refinement of \cite[Theorem 6.4]{ko1} which determines the existence of a nonzero dimension function. Any dimension function of $\Gamma$ comes from a finite dimensional representation of $L(\Gamma)$ by \cite[Corollary 3.7]{ko1} but a representation may not be uniquely determined by its dimension function. Theorem \ref{uzun1} in the next section will further refine Theorem \ref{27} by classifying all finite dimensional representations of $L(\Gamma)$ for a row-finite digraph $\Gamma$. \section{The Classification of The Finite Dimensional Representations of $L(\Gamma)$ } Let's first show that the reduction algorithm applied to a (row-finite) digraph gives a Morita equivalence at the level of Leavitt path algebras. \begin{theorem} \label{Morita} If $\Gamma'$ is a reduction of $\Gamma$ then $L(\Gamma)$ and $L(\Gamma')$ are Morita equivalent, that is, their $($unital$)$ module categories are equivalent. This equivalence preserves the subcategories of finite dimensional modules and modules of finite type. \end{theorem} \begin{proof} We may assume that $\Gamma'=(V',E', s',t')$ is a one step reduction of $\Gamma=(V,E,s,t)$ with $V'=V\setminus \{v\}$ as above. An $L(\Gamma)$-module $M$ is equivalent to a quiver representation $\rho$ of $\Gamma$ satisfying ($I$) by Theorem \ref{teorem}. We will construct the corresponding $L(\Gamma')$-module $M'$ as a quiver representation $\rho'$ of $\Gamma'$ as follows: $\rho'(w):=\rho(w)$ if $w \in V' \>$ and $\> \rho'(e):=\rho(e)$ if $e \in E\cap E'$. For a new arrow $fg$ we define $\rho'(fg)$ to be the composition $\rho'(sf) \stackrel{\rho(f)}{\longrightarrow} \rho(v) \stackrel{\rho(g)}{\longrightarrow} \rho'(tg) $. To see that $\rho'$ satisfies $(I)$ at $w \in V'$ we need to combine the isomorphisms $\rho(v) \longrightarrow \bigoplus_{sg=v } \rho(tg)$ and $\rho' (w) =\rho(w) \longrightarrow \bigoplus_{sf=w} \rho(tf)$ given by $(I)$ for $\rho$ at $v$ and $w$ (using that $\bigoplus_{sf=w} \rho(tf)$ can be broken up into summands with $tf=v$ and those with $tf \neq v$). The definition of $\rho(fg)$ yields the desired isomorphism at $w$. If $\phi$ is a morphism between quiver representations of $\Gamma$ satisfying (I) then the morphism $\phi'$ between the corresponding quiver representations of $\Gamma'$ is the restriction of $\phi$ to $\phi'$.\\ Given a quiver representation $\sigma$ of $\Gamma'$ we define the quiver representation $\tilde{\sigma}$ of $\Gamma$ as $\tilde{\sigma}(w):=\sigma(w)$ if $w \in V'$ and $\tilde{\sigma}(v):=\bigoplus_{sg=v} \sigma(tg)$ (note that $tg $ is in $V'$ since $v$ is a loopless non-sink); also $\tilde{\sigma}(e):=\sigma (e)$ if $e \in E \cap E'$, if $tf=v$ then $\tilde{\sigma}(f):\tilde{\sigma}(sf) \stackrel{(\sigma (fg))_{sg=v}}{\longrightarrow } \tilde{\sigma}(v) = \bigoplus_{sg=v} \sigma(tg)$ and if $sh=v$ then $\tilde{\sigma}(h)$ is the projection from $\tilde{\sigma}(v)=\bigoplus_{sg=v} \sigma (tg)$ to the summand $\sigma (th)$. The condition $(I)$ at $v$ is immediate from the definitions. The definitions of $\tilde{\sigma}(v)$ and $\tilde{\sigma}(f)$ where $sf=w$ and $tf=v$ above yield $(I)$ for $\tilde{\sigma}$ at $w\neq v$ (using $t'fg=tg$). If $\psi$ is a morphism between quiver representations of $\Gamma'$ then the morphism $\tilde{\psi}$ between the corresponding quiver representations of $\Gamma$ is given by $\tilde{\psi} (w):= \psi(w)$ for $w\neq v$ and $\tilde{\psi }(v):=\bigoplus_{sg=v} \psi (tg)$.\\ By construction $(\tilde{\sigma})' =\sigma$ and $(\tilde{\psi})'=\psi$. An isomorphism $\theta: \rho \longrightarrow \tilde{(\rho')}$ is given by $\theta_v=(\rho(te))_{se=v }$ and $\theta_w=id_{\rho(w)}$ for $w \neq v$. Since $\theta$ is compatible with $\tilde{(\phi')}$, the unital module categories of $L(\Gamma)$ and $L(\Gamma')$ are equivalent. Both constructions $\> \> \tilde{} \> \> $ and $ \> ' \>$ preserve finite support and finite dimensionality at every vertex, hence the subcategories of finite dimensional modules and modules of finite type are preserved. \end{proof}\\ \begin{remark} An easy example of a Morita equivalence which does not preserve the subcategory of finite dimensional modules is given by the Leavitt path algebras consisting of a single vertex (which is isomorphic to $\mathbb{F}$) and the infinite line graph with one sink $$\Gamma : \qquad \xymatrix{& \cdots \> \> \ar [r] &{\bullet} \ar [r] & {\bullet} \ar [r] & {\bullet} & }$$ (which is isomorphic to $M_{\infty}(\mathbb{F})$, infinite matrices with finitely many nonzero entries.) The Morita equivalence is given by tensoring with the $(\mathbb{F}, M_{\infty}(\mathbb{F}))$-bimodule $\mathbb{F}^{(\mathbb{N})}$, finite $\mathbb{F}$-sequences. The latter has no finite dimensional modules (by Theorem 6.4 of \cite{ko1} this also follows from Theorem \ref{uzun1} below) since $\Gamma$ has no cycles and one maximal sink but with infinitely many predecessors. \end{remark} The proof of Theorem \ref{Morita} above showing that the module categories of $L:=L(\Gamma)$ and $L':= L(\Gamma')$ are equivalent is constructive and explicit at the level of quiver representations. The quiver representation for the reduction $\Gamma'$ is obtained by restricting the quiver representation of $\Gamma$ to the vertices of $\Gamma'$. In the opposite direction the representation at the omitted vertex $v$ is recovered by taking the direct sum of the representations at $\{te \vert se=v \}$. A more typical Morita equivalence would be given by an $(L,L')$-bimodule $P$ (a progenerator for $\mathfrak{M}_{L'}$) and $(L,L')$-bimodule $Q$ (a progenerator for $\mathfrak{M}_{L}$) such that $ \underline{\> \> \> \> }\otimes_L P$ and $\underline{\>\> \> \> } \otimes_{L'} Q$ give the equivalences between the categories $\mathfrak{M}_{L'}$ and $\mathfrak{M}_{L}$ . Identifying $L'$ with the subalgebra $\{ \alpha \in L \mid v \alpha=0=\alpha v \} $ as above we have $P= \{ \alpha \in L \mid \alpha v=0 \} = L' \oplus vL'$ and $Q= \{ \alpha \in L \mid v\alpha =0 \} = \oplus_{_{v\neq w \in V}} wL $ . \\ Next we want to classify finite dimensional modules of the Leavitt path algebra $L(\Gamma)$ of a row-finite digraph $\Gamma$ by determining all indecomposable finite dimensional modules. First we reduce the problem to a finite digraph $\Lambda$ (the support subgraph of the module). Then we apply the reduction algorithm to $\Lambda$ to obtain a disjoint union of loops and isolated vertices. Now the corresponding module is the direct sum of submodules associated to an isolated vertex or a loop and the classification of these is standard linear algebra. The result can then be pulled back to $L(\Gamma)$ via the explicit Morita equivalence of Theorem \ref{Morita}. \begin{theorem} \label{uzun1} A finite dimensional $L_{\mathbb{F}}(\Gamma)$-module where $\Gamma$ is a row-finite digraph has a functorial direct sum decomposition with the support subgraph of each summand being the predecessors of a maximal sink or a maximal cycle with finitely many predecessors. A summand $M$ corresponding to a sink $w$ is uniquely and completely determined by $dim_{\mathbb{F}} (Mw)$ (up to isomorphism). This summand is indecomposable if and only if $dim_{\mathbb{F}}(Mw)=1$ if and only if it is simple. A summand corresponding to a cycle decomposes into a direct sum of primary submodules, each associated with an irreducible polynomial $f(x)$ in $\mathbb{F}[x]$ with $f(0)=1$. The indecomposable summands of a primary submodule are uniquely and completely determined (up to isomorphism) by a positive integer, this integer is 1 if and only if the indecomposable is simple. \end{theorem} \begin{proof} The support subgraph $\Lambda:=\Gamma_{_M}$ of a finite dimensional $L(\Gamma)$-module $M$ is finite and its cycles have no exits by Lemma \ref{exits}. The $L(\Gamma)$-module structure is the restriction of the $ L(\Gamma)/I_{_M}\cong L(\Lambda)$-module structure. A complete reduction $\bar{\Lambda}$ of $\Lambda$ is a disjoint union of loops (in one-to-one correspondence with the cycles of $\Lambda$) and isolated vertices (in one-to-one correspondence with the sinks of $\Lambda$). We will work with the corresponding $L(\bar{\Lambda})$-module via the Morita equivalence of Theorem \ref{Morita}.\\ $L(\bar{\Lambda})$ is isomorphic to the direct sum of its subalgebras given by the connected components of $\bar{\Lambda}$ which are loops or isolated vertices. Each subalgebra is of the form $L(\tilde{\Lambda})v$ and isomorphic to $\mathbb{F}$ if $v$ is a sink or $\mathbb{F}[x,x^{-1}]$ otherwise (where the loop $e_v$ at $v$ corresponds to $x$ and $e_v^*$ corresponds to $x^{-1}$). Similarly, any $L(\tilde{\Lambda})$-module N has a direct sum of decomposition $N=\oplus Nv$ where the subspaces $Nv$ are submodules in this case. The $L(\bar{\Lambda})= \oplus L(\bar{\Lambda})v$ module structure of $N=\oplus Nv$ is given by coordinate-wise multiplication (i.e., $Nv$ is an $L(\bar{\Lambda})$-module and all other factors $L(\bar{\Lambda})w$ with $w \neq v$ annihilate $Nv$).\\ The support subgraph $\bar{\Lambda}_{Nv}$ of the $L(\bar{\Lambda})$-submodule $Nv$ is just the isolated vertex $v$ if $v$ is a sink and it consists of the vertex $v$ and the loop $e_v$ at $v$ when $v$ is not a sink. Also $L(\bar{\Lambda}_{Nv})\cong L(\bar{\Lambda})v$ which is (isomorphic to) $\mathbb{F}$ or $\mathbb{F}[x,x^{-1}]$ depending on whether $v$ is a sink or not. When $v$ is a sink, $Nv$ is a finite dimensional vector space determined (up to isomorphism) by its dimension. \\ When $v$ is not a sink, $Nv$ is a finite dimensional $\mathbb{F}[x,x^{-1}]$-module which is a direct sum of its primary components each corresponding to an irreducible polynomial $f(x)$ in $\mathbb{F}[x]$ with $f(0)=1$. (These polynomials are the irreducible divisors of the characteristic polynomial of $e_v$ considered as an endomorphism of the vector space $Nv$.) The primary component corresponding to $f(x)$ is a direct sum of indecomposables, each specified by a positive integer $n$ (where $f(x)$ is the minimal polynomial of the linear transformation given by $e_v$ restricted to this primary submodule and $f(x)^n$ is its characteristic polynomial). The indecomposable is simple if and only if $n=1$.\\ The canonical decomposition $N=\oplus Nv$ of the $L(\bar{\Lambda})$-module gives a canonical decomposition of the corresponding $L(\Lambda)$-module via the explicit functorial Morita equivalence of Theorem \ref{Morita}. This decomposition can further be pulled back, via the epimorphism $L(\Gamma) \longrightarrow L(\Lambda)$, to a canonical decomposition of the original $L(\Gamma)$-module. When $v$ is a sink in $\bar{\Lambda}$, the support of the $L(\Gamma)$-module $M$ corresponding to the $L(\bar{\Lambda})$-module $Nv$ is the predecessors of the maximal sink $v$ of $\Gamma$. For each vertex $u$ of $\Gamma$, $Nu$ is isomorphic to a direct sum of copies of the vector space $Nv$ indexed by the paths from $u$ to $v$. For any arrow $e$ of $\Gamma$ each path from $te$ to $v$ yields a path from $se$ to $v$ starting with $e$. The linear transformation (projection) given by $e$ from $Mse$ to $Mte$ is identity on the summands corresponding to paths related as above and it is 0 otherwise.\\ If $v$ is not a sink in $\bar{\Lambda}$, so there is the loop $e_v$ based at $v$, then the support of the $L(\Gamma)$-module $M$ corresponding to $Nv$ is the predecessors of the maximal cycle $C_v$ in $\Gamma$ corresponding to the loop $e_v$ in $\bar{\Lambda}$. The vertex $v$ of $\bar{\Lambda}$ is also a vertex in $\Lambda$ (and hence $\Gamma$ since $\Lambda$ is a subgraph of $\Gamma$): it is the only vertex of the cycle $C_v$ which is not eliminated during the complete reduction of $\Lambda$ to $\bar{\Lambda}$. Let $\tilde{e}_v$ be the arrow on the cycle $C_v$ starting at $v$. Now $Mw \cong Nv$ for each vertex $w$ on $C_v$ and $Mu$ is (isomorphic to) the direct sum of copies of $Nv$ indexed by the paths from $u$ to $C_v$ (equivalently, the paths from $u$ to $v$ not containing $C_v$). The linear transformation corresponding to $\tilde{e}_v$ is given by $e_v$ while all other arrows on $C_v$ give $id_{Nv}$, the identity transformation. For any arrow $e$ of $\Gamma$, different from $\tilde{e}_v$ the linear transformation given by $e$ from $Mse$ to $Mte$ is the projection defined just as in the paragraph above. \\ The decompositions of the $L(\bar{\Lambda})$-module $Nv$ into primary summands and indecomposables carry over functorially to the $L(\Gamma)$-module $M$ and the simple $L(\bar{\Lambda})$-modules correspond to simple $L(\Gamma)$-modules. \end{proof}\\ Consequently, there are two kinds of finite dimensional indecomposable unital $L(\Gamma)$-modules when $\Gamma$ is row finite: if $M$ is of the first kind then $M$ is completely determined by a maximal sink $v$ with finitely many predecessors. The subspace $Mw$ for any vertex $w$ has dimension equal to the number of paths from $w$ to $v$ (hence the support of $M$ is the predecessors of $v$). The linear transformations given by the arrows are projections. These indecomposables are simple. \\ An indecomposable $M$ of the second kind is determined by a maximal cycle $C$ with finitely many predecessors, an irreducible $f(x) \in \mathbb{F}[x]$ with $f(0)=1$ and a positive integer $n$. Now, $Mv\cong \mathbb{F}[x]/ f(x)^n$ for any vertex $v$ on $C$. The linear transformation given by one of the arrows on $C$ corresponds to multiplication by $x$, the remaining arrows on $C$ give the identity transformation (isomorphism type of $M$ is independent of which arrow corresponds to multiplication by $x$). For any vertex $w$, $Mw$ is isomorphic to a direct sum of copies of $Mv$ indexed by the paths from $v$ to $C$ (which do not traverse $C$). The arrows that are not on the cycle $C$ give projections. These indecomposables are simple if and only if $n=1$.\\ If $M$ and $N$ are simple then any nontrivial extension of $M$ by $N$ is indecomposable. Our classification shows that the only indecomposables of length 2 correspond to pairs $(C,f(x)^2)$, hence: \begin{corollary} If $M$ and $N$ are finite dimensional simple $L(\Gamma)$-modules then $Ext(M,N)=0$ unless $N\cong M$ and $M$ corresponds to a pair $(C,f(x))$ as above. \end{corollary} The corollary above does not hold when we don't assume that $M$ and $N$ are finite dimensional. There are indecomposable $L(\Gamma)$-modules of arbitrary length with several non-isomorphic factors \cite{ko2}.\\ Any finite dimensional $L(\Gamma)$-module $M$ is, up to isomorphism, a unique (up to ordering) direct sum of the indecomposables described above. Thus Theorem 6.4 of \cite{ko1} stating that $L(\Gamma)$ has a nonzero finite dimensional module if and only if $\Gamma$ has a maximal sink or cycle with finitely many predecessors is now a consequence of Theorem \ref{uzun1}. While the existence of a nonzero finite dimensional representation of $L(\Gamma)$ depends only on the digraph $\Gamma$, the classification of all finite dimensional representations depends also on the coefficient field. \\ When the field $\mathbb{F}$ is algebraically closed the simple modules obtained from a maximal cycle (with finitely many predecessors) are parametrized by $\mathbb{F}^\times$, the nonzero elements of the field. The indecomposables require an additional parameter, a positive integer. In particular, the category of finite dimensional representations of the Leavitt path algebra $L(\Gamma)$ of a finite digraph $\Gamma$ is tame in contrast to the finite dimensional quiver representations of $\Gamma$ which are wild unless the connected components of the undirected $\Gamma$ are Dynkin diagrams of types $A, \> D, \> E, \> \tilde{A}, \> \tilde{D}, \> \tilde{E}$ by a theorem of Gabriel \cite{dw05}. The algebra $L(\Gamma)$ has finite representation type, that is, there are only finitely many finite dimensional indecomposables (up to isomorphism) if and only if $\Gamma$ has no maximal cycles with finitely many predecessors and finitely many maximal sinks with finitely many predecessors.\\ We can almost remove the qualification "with finitely many predecessors" if we consider modules of finite type instead of finite dimensional modules. Each maximal sink or cycle with infinitely many predecessors contributes $L(\Gamma)$-module of finite type. The converse, however, does not hold. There may be no sinks or cycles in $\Gamma$ but $L(\Gamma)$ may still have a module of finite type. For example $$\Gamma : \xymatrix{& {\bullet}^{v_1} \ar [r] ^{e_1} & {\bullet}^{v_2} \ar [r] ^{e_2} & {\bullet}^{v_3} \> \> \cdots & }$$ The module $M$ of finite type is given by $Mv_i:=\mathbb{F}$, $i=1,2, \cdots$ and all arrows being $id_{\mathbb{F}}$.\\ \begin{example} The two digraphs below have no maximal sinks or maximal cycles, hence their Leavitt path algebras have no nonzero finite dimensional representations. $$ \xymatrix{ &\> \> { \bullet} \ar@(ul,ur) \ar@(dl,dr) \ar@/^1pc/[r] \ar[r] & {\bullet} } \qquad \quad \xymatrix{& {\bullet} \ar@/^1pc/[r] &\ar [l] {\bullet} \ar [r] & {\bullet} \ar@/^1pc/[l] }$$\\ \end{example} \begin{example} Finite dimensional representations of the Leavitt path algebras of the five digraphs below are all equivalent since they have no maximal sinks and the predecessor subgraphs of their unique maximal cycle are isomorphic. \\ $$ \xymatrix{ &\> \> { \bullet} \ar@(ul,ur) } \qquad \quad \xymatrix{ &\> \> { \bullet} \ar[r] \ar@(ul,ur) &{ \bullet} } \qquad \qquad \quad \quad \qquad \xymatrix{ {\bullet} \ar@(ul,dl) \ar@{->}[r] \ar@{->}[dr] &\bullet \\ & \bullet }$$ $$ \xymatrix{ &\> \> { \bullet} \ar[r] \ar@(ul,ur) &{ \bullet} \ar@(ul,ur) \ar@(dl,dr) } \quad \qquad \quad \xymatrix{ &\> \> { \bullet} \ar[r] \ar@(ul,ur) &\> \> { \bullet} \ar[r] \ar@(ul,ur) & \>\> { \bullet} \ar[r] \ar@(ul,ur) &\> \> { \bullet} } $$\\\\ \end{example} A somewhat different viewpoint is to start with the finite dimensional $L(\Gamma)$-module $M$ and to determine all its invariants. We arbitrarily fix a base vertex $w$ for each maximal cycle with finitely many predecessors and let $W$ be the set of the base vertices of maximal cycles and the maximal sinks with finitely many predecessors. Regarding $M$ as an $L(\Gamma_{_M})$-module (where $\Gamma_{_M}=(V_{_M}, E_{_M})$ is the support subgraph of $M$), we get the corresponding $L(\bar{\Gamma}_{_M})$-module $N$ using the explicit Morita equivalence of Theorem \ref{Morita}. Here $\bar{\Gamma}_{_M}$ is the complete reduction of $\Gamma_{_M}$ in which all vertices not in $W$ are eliminated. The vertex set of $\bar{\Gamma}_{_M}$ is $V_{_M} \cap W$ since $\bar{\Gamma}_{_M}$ is a disjoint union of loops and isolated vertices as explained above.\\ In the decomposition $N=\oplus_{w \in W} Nw$ each $Nw$ is an $L(\bar{\Gamma}_{_M})$-submodule. Let $M_{\leadsto w}$ be the $L(\Gamma_{_M})$-module corresponding to the $L(\bar{\Gamma}_{_M})$-module $Nw$ via Theorem \ref{Morita}. Regarding $M_{\leadsto w}$ as an $L(\Gamma)$-module via the epimorphism $L(\Gamma) \longrightarrow L(\Gamma_{_M})$ we have $M=\oplus_{w \in W} M_{\leadsto w} \> $; this is the functorial decomposition of Theorem \ref{uzun1}. The support of $M_{\leadsto w}$ is $\Gamma_{\leadsto w} \> $ (note that these support subgraphs may not be disjoint for distinct $w$). \\ The submodule $M_{\leadsto w}$ is determined by the vector space $Mw$ if $w$ is a sink, it is determined by the vector space $Mw$ and the invertible linear operator on $Mw$ given by right multiplication with $C_w$, the cycle containing $w$ (considered as an element of $L(\Gamma)$). If $f:M \longrightarrow M'$ is an $L(\Gamma)$-module homomorphism where $M$ and $M'$ are both finite dimensional then we may consider $f$ an $L(\Lambda)$-morphism where $\Lambda=\Gamma_{_M} \cup \Gamma_{_{M'}} \>$. The complete reduction $\bar{\Lambda}$ of $\Lambda$ with vertex set $W \cap (V_{_M} \cup V_{_{M'}})$ is again a disjoint union of loops and isolated vertices. Since $\bar{f} (Nw) \subseteq N'w$ for the corresponding $L(\bar{\Lambda})$-morphism $\bar{f} : N \longrightarrow N'$ we have $f(M_{\leadsto w}) \subseteq M_{\leadsto w}' \>$.\\ The $L(\Gamma)$-morphism $f \mid_{_{M_{\leadsto w}}} : M_{\leadsto w} \longrightarrow {M'}_{\leadsto w}$ is determined by the linear transformation $f \mid_{_{Mw}} :Mw \longrightarrow M'w $ completely. When $w$ is not a sink the linear transformation $f \mid_{_{Mw}} $ intertwines with $C_w$ (equivalently, $f \mid_{_{Mw}}$ is an $\mathbb{F}[x,x^{-1}]$-morphism where $x$ acts as $C_w$). Conversely, for any finite subset of $W$, given arbitrary finite dimensional vector spaces $Mw$, invertible operators $C_w: Mw \longrightarrow Mw $ when $w$ is not a sink we can construct an $L(\bar{\Lambda})$-module $N$ with $Nw :=Mw$ and the linear transformation corresponding to $e_w$ (the loop based at w) is $C_w \>$. The corresponding $L(\Gamma)$-module $M$ (via Theorem \ref{Morita} and the epimorphism $L(\Gamma) \longrightarrow L(\Gamma_{_M})$) has the prescribed $Mw$ and $C_w$ for each $w \in W$. Similarly, an $L(\Gamma)$-morphism $f :M \longrightarrow M'$ with arbitrarily specified $f \mid_{_{Mw}}$ satisfying the intertwining conditions can also be realized.\\ Paraphrasing the discussion above: The category of finite dimensional $L(\Gamma)$-modules for a row-finite digraph $\Gamma$ is equivalent to the direct sum of the categories indexed by $W$ where each summand corresponding to a sink is the category of finite dimensional vector spaces and each remaining summand is the category of finite dimensional $\mathbb{F}[x,x^{-1}]$-modules. That is, the category of finite dimensional $L(\Gamma)$-modules for a row-finite digraph $\Gamma$ is equivalent to the category of finite dimensional $L(\Lambda)$-modules where $\Lambda$ is a disjoint union of loops (corresponding to the maximal cycles of $\Gamma$ with finitely many predecessors) and isolated vertices (corresponding to the maximal sinks of $\Gamma$ with finitely many predecessors): \begin{theorem} \label{notasyon} If $\Gamma$ is a row-finite digraph then $$\mathfrak{M}_L^{^{fd}} \backsimeq \Big( \bigoplus_{_{\mathcal{S}} } \mathfrak{M}_{\mathbb{F}}^{^{fd}} \Big) \oplus \Big(\bigoplus_{_{\mathcal{T}} } \mathfrak{M}_{\mathbb{F} [x,x^{-1}]}^{^{fd}}\Big)$$ where $L:=L_{\mathbb{F}}(\Gamma)$ , $\mathfrak{M}_A^{^{fd}}$ is the category of finite dimensional $A$-modules, $S$ is the set of maximal sinks with finitely many predecessors in $\Gamma$, $\mathcal{T}$ is the set of maximal cycles with finitely many predecessors in $\Gamma$ and $\backsimeq$ denotes equivalence of categories. \end{theorem} We can get a more module theoretic description of the classification of the indecomposable finite dimensional $L(\Gamma)$-modules by unraveling the correspondence between quiver representations and $L(\Gamma)$-modules of Theorem \ref{teorem}: Given a finite dimensional indecomposable $L(\Gamma)$-module its support $\Gamma_M$ contains either a unique sink $w$ or a unique cycle $C$. If $\Gamma_M$ has a sink $w$ then $M\cong wL(\Gamma)$, a finite dimensional projective $L(\Gamma)$-module which is also simple. These are the only projective indecomposable (or simple) finite dimensional $L(\Gamma)$-modules. Any finite dimensional projective $L(\Gamma)$-module is isomorphic to a direct sum of these $\{wL(\Gamma)\}$ where $w$ is a maximal sink with finitely many predecessors.\\ If $\Gamma_M$ contains a cycle $C$ then $C$ is the unique cycle in $\Gamma_M=\Gamma_{\leadsto C}$ and in $\Gamma_M$ $C$ has no exit. $M$ corresponds to a pair $(C,f(x))$ in the classification, we can recover the polynomial $f(x)$ as follows: if $v=sC$ and $\varphi$ is the linear operator on the vector space $Mv$ given by right multiplication with $C$ then $f(x)$ is $det(1-x\varphi) $, essentially the characteristic polynomial of $\varphi$.\\ Conversely, given $(C,f(x)^n)$ where $C$ is a maximal cycle with finitely many predecessors and $f(x) \in \mathbb{F}[x]$ an irreducible polynomial with constant term 1, the corresponding finite dimensional indecomposable $M$ is $\mathbb{F}[x,x^{-1}]/ (f(x)^n) \otimes vL(\Gamma_{\leadsto C})$ where $v=sC$ and tensor product is over $\mathbb{F}[x,x^{-1}] \cong vL(\Gamma_{\leadsto C})v$ (the isomorphism is by Lemma \ref{ez}). Note that $vL(\Gamma_{\leadsto C})$ is a $(vL(\Gamma_{\leadsto C})v, L(\Gamma_{\leadsto C}))$-bimodule, so $M$ is an $L(\Gamma_{\leadsto C})$-module and hence inherits an $L(\Gamma)$-module structure since $ L(\Gamma_{\leadsto C}) \cong L(\Gamma)/ (V\setminus V_{\leadsto C})$.\\ The dimension function $d(u):=dim^\mathbb{F}(Mu)$ can also be computed in terms of $(C,f(x))$: if $v=sC$ and $P_v^u$ denotes the set of paths $p$ such that $sp=u$, $tp=v$ and $p$ does not contain $C$ then $d(u)= \vert P_v^u\vert degf(x)$. Similarly, when $M$ corresponds to maximal sink $w$ with finitely many predecessors, i.e., $M\cong wL(\Gamma)$ then $d(u)=\vert P_w^u\vert$ where $P_w^u$ is the set of paths from $u$ to $w$.\\ Another description of the indecomposable module $M$ corresponding to the pair $(C, f(x))$ is $vL(\Gamma)/f(C^*)L(\Gamma)$ where $v=sC$ \cite{ko2}. A consequence is that all finite dimensional $L(\Gamma)$-modules are finitely presented. Also $M$ is a rational Chen module if and only if $f(x)=1-x$ and $M$ is a twisted rational Chen module if and only if $f(x)=1-\lambda x$ where $0\neq \lambda \in \mathbb{F}$ \cite{che15}, \cite{ko2}. If $f(x)$ is an irreducible polynomial (as always $f(0)=1$) then the corresponding module $M$ is simple but not a Chen module. \\ \textbf{Acknowledgement}\\ Both authors were partially supported by TUBITAK grant 115F511.
19c8e250703a18d97101e53afd35c420154906aa
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Outline} \maketitle \tocless\section{Introduction}{} \addtocontents{toc}{\protect\setcounter{tocdepth}{1}} There is a vast literature on the theory of {\em Newton polyhedra}, initiated by V.\ I.\ Arnold's hypothesis that `reasonable' invariants of objects (e.g.\ singularities, varieties, etc) associated to a `typical' (system of) analytic function(s) or polynomial(s) should be computable in terms of their {\em Newton diagrams} or {\em Newton polytopes} (\cref{Newton-definition}). In this article we revisit two of the original questions that shaped this theory, namely the question of computing the Milnor number\footnote{\label{milnor-footnote}Let $f \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, where $\ensuremath{\mathbb{K}}$ is an algebraically closed field. Then the {\em Milnor number} $\mu(f)$ of $f$ at the origin is the dimension (as a vector space over $\ensuremath{\mathbb{K}}$) of the quotient ring of $\ensuremath{\mathbb{K}}[[x_1, \ldots, x_n]]$ modulo the ideal generated by partial derivatives of $f$ with respect to $x_j$'s.} of the singularity at the origin of the hypersurface determined by a generic polynomial or power series, and the question of computing the number (counted with multiplicity) of isolated zeroes of $n$ generic polynomials in $n$ variables. The first question was partially solved in a classical work of Kushnirenko \cite{kush-poly-milnor} and a subsequent work of Wall \cite{wall}; Bernstein \cite{bern}, following the work of Kushnirenko \cite{kush-poly-milnor}, solved the second question for the `torus' $(\kk^*)^n$ (where $\ensuremath{\mathbb{K}}$ is an algebraically closed field and $\ensuremath{\mathbb{K}}^* := \ensuremath{\mathbb{K}} \setminus \{0\}$), and many other authors (including Khovanskii \cite{khovanus}, Huber and Sturmfels \cite{hurmfels-bern}, Rojas \cite{rojas-toric}) gave partial solutions for the case of the affine space $\ensuremath{\mathbb{K}}^n$. Extending the approach from Bernstein's proof in \cite{bern} of his theorem, we give a complete solution to the first problem, and complete the program of extending Bernstein's theorem to $\ensuremath{\mathbb{K}}^n$ (or more generally, to the complement of a union of coordinate subspaces of $\ensuremath{\mathbb{K}}^n$). \\ In \cite{kush-poly-milnor} Kushnirenko gave a beautiful expression for a lower bound on the generic Milnor number and showed that a polynomial (or power series) attains this bound in the case that its Newton diagram is {\em convenient}\footnote{Kushnirenko used the term {\em commode} in French; `convenient' is also widely used, see e.g.\ \cite{boubakri-greuel-markwig}.} (which means that the Newton diagram contains a point on each coordinate axis), and it is {\em Newton non-degenerate}, i.e.\ the following is true (see \cref{inndefinition} for a precise formulation): \begin{savenotes} \begin{align} \parbox{.84\textwidth}{% for each {\em weighted order}\footnote{A {\em weighted order} corresponding to weights $(\nu_1, \ldots, \nu_n) \in \ensuremath{\mathbb{Z}}^n$ is the map $\nu:\ensuremath{\mathbb{K}}[x_1, \ldots, x_n] \to \ensuremath{\mathbb{Z}}$ given by $\nu(\sum a_\alpha x^\alpha ) := \min\{\sum_{k=1}^n \alpha_k\nu_k: a_\alpha \neq 0\}$.} $\nu$ on $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ with positive weights, the partial derivatives of the corresponding {\em initial form}\footnote{Given a weighted order $\nu$, the {\em initial form} of $f = \sum_\alpha a_\alpha x^\alpha \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ is the sum of all $a_\alpha x^\alpha$ such that $\nu(x^\alpha) = \nu(f)$.} of $f$ do not have any common zero on $(\kk^*)^n$.} \tag{NND} \end{align} \end{savenotes} However, it is straightforward to construct examples which show that \begin{prooflist} \item \label{isolated-defect} if the Newton diagram is not convenient, Newton non-degeneracy of $f$ does not imply `finite determinacy' of $f$, i.e.\ it does not imply that the origin is an isolated singular point of $f = 0$ (take e.g.\ $f := x_1^q \cdots x_n^q$, where $q$ is an integer $\geq 2$ not divisible by $\charac(\ensuremath{\mathbb{K}})$), \item \label{necessary-defect} even if the Newton diagram is convenient, Newton non-degeneracy is not necessary for the Milnor number to be generic (see \cref{Milnor-not-Newton}), \item \label{existential-defect} if $\charac(\ensuremath{\mathbb{K}}) > 0$, then there are (convenient) diagrams $\Gamma$ such that no polynomial with Newton diagram $\Gamma$ is Newton non-degenerate (see \cref{no-(I)NND}). \end{prooflist} Wall \cite{wall} found another sufficient criterion, called {\em inner Newton non-degeneracy} by Boubakri, Greuel and Markwig \cite{boubakri-greuel-markwig}, for the Milnor number to be generic. He showed that inner Newton non-degeneracy implies finite determinacy, correcting defect \ref{isolated-defect}, and that Kushnirenko's formula computes the Milnor number for inner Newton non-degenerate polynomials. However, inner Newton non-degeneracy also suffers from defects \ref{necessary-defect} and \ref{existential-defect} (see \cref{Milnor-not-inner,no-(I)NND}), and it remained a topic of ongoing investigations to completely understand what makes the Milnor number generic (see e.g.\ the works of Boubakri, Greuel and Markwig \cite{boubakri-greuel-markwig}, Greuel and Nguyen \cite{greuel-nguyen}). \\ In \cite{kush-poly-milnor} Kushnirenko proved another beautiful result that the number (counted with multiplicity) of isolated solutions on $(\kk^*)^n$ of $n$ generic (Laurent) polynomials $f_1, \ldots, f_n$ with identical Newton polytope $\ensuremath{\mathcal{P}}$ is bounded by $n! \vol(\ensuremath{\mathcal{P}})$, and that the bound is attained if and only if the following non-degeneracy condition is satisfied (in which case we say that $f_1, \ldots, f_n$ are {\em BKK non-degenerate}) \begin{align} \parbox{.72\textwidth}{% for each non-trivial weighted order $\nu$ on $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, the corresponding initial forms of $f_1, \ldots, f_n$ do not have any common zero on $(\kk^*)^n$.} \tag{BKK} \label{bkk} \end{align} Bernstein \cite{bern} removed the restriction of $f_i$'s having identical Newton polytopes. He showed that the general upper bound for the number (counted with multiplicity) of isolated solutions on $(\kk^*)^n$ is the {\em mixed volume} (\cref{mixed-defn}) of Newton polytopes of $f_1, \ldots, f_n$, and that the bound is attained if and only if $f_1, \ldots, f_n$ are BKK non-degenerate. A number of authors have worked on extending Bernstein's result, also known as the Bernstein-Kushnirenko-Khovanskii bound (in short {\em BKK bound}), to the case of $\ensuremath{\mathbb{K}}^n$; in particular, Khovanskii \cite{khovanus} gave a formula for the extended BKK bound in the special case that the Newton polytopes are such that generic systems have no zeroes in any (proper) coordinate subspace of $\ensuremath{\mathbb{K}}^n$; Huber and Sturmfels \cite{hurmfels-bern} introduced the notion of {\em stable mixed volumes} and used it to give a formula in the case that the Newton polytopes are such that generic systems have no non-isolated zeroes in $\ensuremath{\mathbb{K}}^n$; Rojas \cite{rojas-toric} extended the method of Huber and Sturmfels to give a formula in terms of stable mixed volumes that works without any restriction on Newton polytopes\footnote{The formulae that appear in \cite[Theorem I]{rojas-wang} and \cite[Affine Point Theorem II]{rojas-toric} for the number of roots and intersection multiplicity are incorrect - see \cref{counter-section}. However the formula from \cite[Corollary 1]{rojas-toric} in terms of stable mixed volumes is correct - this can be seen either by directly adapting the proof of the result of Huber and Sturmfels \cite{hurmfels-bern} to arbitrary Newton polytopes and arbitrary characteristics, or by observing that our formula (\cref{bkk-thm}), which works for all Newton polytopes and characteristics, is in the case of zero characteristic equivalent to the formula in \cite[Corollary 1]{rojas-toric}.}. However, the conditions for the attainment of the bound had not been characterized - only some sufficient conditions were known, see e.g.\ \cite[Section 4]{khovanus}, \cite[Main Theorem 2]{rojas-toric}. \subsection{Main results}{} All the results of this article are valid over an algebraically closed field $\ensuremath{\mathbb{K}}$ of arbitrary characteristic. \subsubsection{Intersection multiplicity} Since the Milnor number of the singularity at the origin of $f = 0$ is simply the {\em intersection multiplicity} (\cref{intersection-multiplicity}) at the origin of the partial derivatives of $f$, the problem of understanding genericness of the Milnor number naturally leads to the problem of understanding genericness of intersection multiplicity at the origin of hypersurfaces determined by $n$ polynomials $f_1, \ldots, f_n$ with fixed Newton diagrams. We give a complete solution to this problem: \begin{thm}[\Cref{generic-thm}] \label{generic-thm-0} The intersection multiplicity is the minimum if and only if the following non-degeneracy condition holds: \begin{align} \parbox{.72\textwidth}{% for each weighted order $\nu$ on $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ with positive weights and for each coordinate subspace $K$ of $\ensuremath{\mathbb{K}}^n$, the corresponding initial forms of $f_1|_K, \ldots, f_n|_K$ do not have any common zero on $(\kk^*)^n$.} \tag{$\overline{\text{BKK}}_0$} \label{bkk-bar-0} \end{align} \end{thm} We also give a new formula for the minimum intersection multiplicity (\cref{multiplicity-thm}). Given a fixed collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$, it can be combinatorially characterized (\cref{mult-support-solution}) if the intersection multiplicity at the origin is infinite for every collection of polynomials with these Newton diagrams; if this is not the case, then the the set of polynomials satisfying \eqref{bkk-bar-0} is Zariski dense in the space of polynomials with these Newton diagrams (\cref{bkk-existence}). Finally, we give a criterion which is equivalent to \eqref{bkk-bar-0}, but is computationally less expensive (\cref{weak-lemma}). \subsubsection{Milnor number} Let $f$ be a polynomial or a power series in $n$ variables over $\ensuremath{\mathbb{K}}$, and $\Gamma_1, \ldots, \Gamma_n$ be the Newton diagrams of partial derivatives of $f$. Assume that the minimum intersection multiplicity at the origin of all polynomials with these Newton diagrams is finite (this can be determined combinatorially due to \cref{mult-support-solution}). Then applying \cref{generic-thm-0} we show that \begin{thm}[\Cref{milnor-thm}] The Milnor number of $f$ at the origin is the minimum if and only if $\dfd{1}, \ldots, \dfd{n}$ satisfy \eqref{bkk-bar-0}, in which case we say that $f$ is {\em partially Milnor non-degenerate}. \end{thm} In particular, while (inner) Newton non-degeneracy is determined by {\em partial derivatives of initial forms} of $f$, partial Milnor non-degeneracy is determined by {\em initial forms of partial derivatives} of $f$. Note that \begin{itemize}[wide] \item A priori it is not clear if partially Milnor non-degenerate polynomials exist. Indeed, the collection of polynomials $f_1, \ldots, f_n$ which arise as derivatives of some polynomial is (contained in) a proper closed subvariety of the space of all polynomials with Newton diagrams $\Gamma_1, \ldots, \Gamma_n$, and it is a priori conceivable that this subvariety is contained in the complement of the Zariski dense set of polynomials which satisfy \eqref{bkk-bar-0}. Assertion \eqref{milnor-existence} of \cref{milnor-thm} ensures that this is not the case.\\ \item In particular, partial Milnor non-degeneracy does not suffer from any of the defects \ref{isolated-defect}--\ref{existential-defect}. \\ \item It follows from Kushnirenko's \cite{kush-poly-milnor} and Wall's \cite{wall} works that if a polynomial is Newton non-degenerate and has convenient Newton polytope, or if it is inner non-degenerate, then it is partially Milnor non-degenerate; we also give direct proofs of these results (\cref{newton-wall-lemma}). In \cref{newton-wall-lemma} we also show that if the singularity at the origin of $f = 0$ is isolated, then Newton non-degeneracy implies partial Milnor non-degeneracy, i.e.\ Newton non-degeneracy plus finite determinacy guarantee the minimality of the Milnor number at the origin; this seems to be a new observation. \\ \item In particular our results give an effective solution (\cref{kush-solution}), valid for fields of arbitrary characteristic, to the following problem considered and solved in the characteristic zero case by Kushnirenko \cite{kush-poly-milnor}: \begin{problem} \label{kush-problem} Given a subset $\Sigma$ of $(\ensuremath{\mathbb{Z}}_{\geq 0})^n$, determine if there exists a power series $f$ in $n$ variables supported at $\Sigma$ such that the Milnor number at the origin is finite. If there exists such a function, then also determine the minimum possible Milnor number of such $f$. \end{problem} \end{itemize} \subsubsection{Extension of the BKK bound to $\ensuremath{\mathbb{K}}^n$} Given a fixed collection $\mathscr{P} := (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ of $n$ convex integral polytopes in $\ensuremath{\mathbb{R}}^n$, we characterize those systems of polynomials supported at these polytopes for which the number (counted with multiplicity) of isolated zeroes on $\ensuremath{\mathbb{K}}^n$ (or more generally, on the complement of the union of a given set of coordinate subspaces of $\ensuremath{\mathbb{K}}^n$) is the maximum (\cref{generic-bkk-thm}). More precisely, we (combinatorially) divide the coordinate subspaces of $\ensuremath{\mathbb{K}}^n$ into two (disjoint) groups ${\tilde \scrT}_\mscrP$ and ${\tilde \scrT'}_\mscrP$ such that for a given system $f = (f_1, \ldots, f_n)$ of polynomials supported at $\mathscr{P}$, \begin{itemize} \item for every $K \in {\tilde \scrT'}_\mscrP$, if $f$ has non-isolated roots on the `torus'\footnote{If $K = \bigcap_{j \in J} \{x_j= 0\}$, then the `torus' of $K$ is $K \cap \{\prod_{i \not\in J} x_i \neq 0\}$. \label{torus-footnote}} of $K$, then the number of its isolated roots on $\ensuremath{\mathbb{K}}^n$ is less than that of a generic system. \item if $K \in {\tilde \scrT}_\mscrP$, then the existence of non-isolated roots of $f$ on the torus of $K$ does not necessarily imply that the number of isolated roots of $f$ on $\ensuremath{\mathbb{K}}^n$ is less than that of a generic system. \end{itemize} \begin{thm}[A special case of \cref{generic-bkk-thm}] \label{generic-bkk-thm-0} The number (counted with multiplicity) of isolated zeroes on $\ensuremath{\mathbb{K}}^n$ of a system $f = (f_1, \ldots, f_n)$ of polynomials supported at $\mathscr{P}$ is the maximum if and only if the following two conditions hold: \begin{defnlist \item \label{TPrime-property} for all $K' \in {\tilde \scrT'}_\mscrP$, and for all weighted order $\nu$ on $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ which is either `centered at infinity' on $K'$ or `centered at the torus' of some $K \in {\tilde \scrT}_\mscrP$, the corresponding initial forms of $f_1|_{K'}, \ldots, f_n|_{K'}$ have no common zero on $(\kk^*)^n$, and \item \label{TP-property} for all $K \in {\tilde \scrT}_\mscrP$ and for all weighed orders $\nu$ on $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ centered at the torus of some $K' \in {\tilde \scrT'}_\mscrP$, the corresponding initial forms of $f_1|_K, \ldots, f_n|_K$ have no common zero on $(\kk^*)^n$. \end{defnlist} \end{thm} We also give an equivalent version of the non-degeneracy criterion from \cref{generic-bkk-thm-0} which is computationally less expensive (\cref{weak-bkkriterion}), and a new formula for the maximum possible number of isolated zeroes on $\ensuremath{\mathbb{K}}^n$ (\cref{bkk-thm}). \subsection{Idea of proof of non-degeneracy criteria, and differences from Bernstein's theorem} \label{idea-section \subsubsection{} \label{sufficient-sketch} Our proof of the correctness of non-degeneracy criteria for the intersection multiplicity at the origin and for the extension of BKK bound follows Bernstein's \cite{bern} polynomial homotopy approach for the proof of correctness of BKK non-degeneracy. In particular, the basic idea of the proof of {\bf sufficiency} of the non-degeneracy criteria is as follows: given a sytem $f = (f_1, \ldots, f_n)$ of polynomials in $x := (x_1, \ldots, x_n)$ we consider a one parameter family of systems $f(x,t)= (f_1(x,t), \ldots, f_n(x,t))$ of polynomials that $f(x,0) = f$. If for generic $t$ the intersection multiplicity at the origin of $f_1(x,t), \ldots, f_n(x,t)$ is smaller than that of $f_1, \ldots, f_n$, then there is a curve $C$ consisting of non-zero roots of $f_1(x,t), \ldots, f_n(x,t)$ for $t \neq 0$ which `approaches the origin' as `$t$ approaches zero'. Similarly, if for generic $t$ the number of isolated roots of $f_1(x,t), \ldots, f_n(x,t)$ on $\ensuremath{\mathbb{K}}^n$ is greater than that of $f_1, \ldots, f_n$, then there is a curve $C$ consisting of isolated roots of $f_1(x,t), \ldots, f_n(x,t)$ for $t \neq 0$ such that as `$t$ approaches zero', $C$ either `goes to infinity', or `approaches' a non-isolated root of $f_1, \ldots, f_n$. If $K$ is the smallest coordinate subspace of $\ensuremath{\mathbb{K}}^n$ containing $C$ and $\nu$ is an appropriate weighted order such that $\nu(x_i)$ is proportional to the order of vanishing of $x_i|_C$ at the limiting point on $C$ as $t$ approaches $0$, then the non-degeneracy criteria are violated for $K$ and $\nu$. \subsubsection{} \label{necessary-sketch} Similarly, the basic idea of our proof of the {\bf necessity} of the non-degeneracy criteria is as follows: given a degenerate system $f = (f_1, \ldots, f_n)$ of polynomials and a coordinate subspace $K$ of $\ensuremath{\mathbb{K}}^n$ such that \eqref{bkk-bar-0} fails for $f_1|_K, \ldots, f_n|_K$, we try to construct following Bernstein \cite{bern} an explicit deformation of $f(x,t)$ of $f$ such that for generic $t$ there is a non-zero root of $f_1(x,t), \ldots, f_n(x,t)$ on $K$ which approaches the origin as $t$ approaches zero; if this is possible, then it follows that the intersection multiplicity at the origin of a degenerate system is higher than the minimum possible value. Similarly, if the non-degeneracy criterion for the extended BKK bound fails for $f_1|_K, \ldots, f_n|_K$, we try to construct a deformation $f(x,t)$ of $f$ such that for generic $t$ there is an isolated root of $f_1(x,t), \ldots, f_n(x,t)$ on $K$ which approaches either infinity or a non-isolated root of $f$ as $t$ approaches zero. \subsubsection{} \label{pathologies} The proof of sufficiency of \eqref{bkk-bar-0} follows easily from the arguments outlined in \cref{sufficient-sketch}. However, the proof of necessity of \eqref{bkk-bar-0} runs into a problem in the case that it fails for $f_1|_K, \ldots, f_n|_K$ for some coordinate subspace $K$ such that the dimension of $K$ is {\em less than} the number $N_K$ of $f_j$'s which do not identically vanish on $K$. In that case generic systems have {\em no} solution on the `torus' (see \cref{torus-footnote}) of $K$, so that Bernstein's deformation trick outlined in \cref{necessary-sketch} can not be executed. However, we show in \cref{reduction-lemma} that this scenario can be ignored: if $f_1|_K, \ldots, f_n|_K$ is degenerate for some $K$ such that $\dim (K) < N_K$, then there is a coordinate subspace $K' \supset K$ such that $f_1|_{K'}, \ldots, f_n|_{K'}$ is also degenerate and $\dim (K') = N_{K'}$. It follows that the deformation trick can be performed on $K'$. The necessity of \eqref{bkk-bar-0} follows then in a straightforward way via the arguments sketched in \cref{necessary-sketch}. Regarding the non-degeneracy criterion for the extended BKK bound, however, one has to be more careful and show that for generic $t$ the extra root of $f(x,t)$ is in fact isolated. \subsubsection{} \label{pathologies2} While the non-degeneracy criterion \eqref{bkk-bar-0} for intersection multiplicity is simple and very close to the non-degeneracy criterion \eqref{bkk} for the BKK bound, the non-degeneracy criterion in \cref{generic-bkk-thm-0} for the extended BKK bound is evidently more complicated. One of the reasons for this is the presence of non-isolated zeroes: indeed, if a system of $n$ polynomials has a non-isolated root on $(\kk^*)^n$, then it follows from Bernstein's theorem that the number of isolated roots on $(\kk^*)^n$ is {\em less} than that of a generic system\footnote{provided a generic system has at least one isolated root on $(\kk^*)^n$.}. However, for the case of $\ensuremath{\mathbb{K}}^n$, it is possible to have two systems of polynomials with identical Newton polytopes such that both have the maximum possible number of isolated roots, but one has non-isolated roots on $\ensuremath{\mathbb{K}}^n$ while the other has only isolated roots (\cref{b-example}). More precisely, one has the following: \begin{prop}[A special case of \cref{non-isolated-cor}] \label{non-isolated-lemma-0} Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$. Then each coordinate subspace $K$ of $\ensuremath{\mathbb{K}}^n$ has one (and only one) of the following properties: \begin{enumerate} \item \label{non-isolated-<-0} a generic system supported at $\mathscr{P}$ has finitely many (possibly zero) roots on the `torus' (see \cref{torus-footnote} for the definition) $K^*$ of $K$; existence of non-isolated roots on $K^*$ implies that the number of isolated roots on $\ensuremath{\mathbb{K}}^n$ is less than that of a generic system. \item \label{non-isolated-not-<-0} no system supported at $\mathscr{P}$ has any isolated root on $K^*$; existence of non-isolated roots on $K^*$ does not necessarily imply that the number of isolated roots on $\ensuremath{\mathbb{K}}^n$ is less than that of a generic system. \end{enumerate} \end{prop} It follows that while trying to detect degeneracy of $f_1, \ldots, f_n$ for the extended BKK bound, for every coordinate subspace $K$ of $\ensuremath{\mathbb{K}}^n$, \begin{itemize} \item non-isolated roots of $f_1|_K, \ldots, f_n|_K$ should be detected only if $K$ is as in assertion \eqref{non-isolated-<-0} of \cref{non-isolated-lemma-0} (this is partly why condition \ref{TPrime-property} of \cref{generic-bkk-thm-0} is applied only to coordinate subspaces from ${\tilde \scrT'}_\mscrP$), and \item a criterion should be formulated to detect if any isolated root of $f_1|_K, \ldots, f_n|_K$ is {\em non-isolated} in the set of zeroes of $f_1, \ldots, f_n$ in $\ensuremath{\mathbb{K}}^n$ (\cref{c-example}); this is partly condition \ref{TP-property} of \cref{generic-bkk-thm-0}. \end{itemize} \subsection{Idea of proof of the formulae for intersection multiplicity and the extended BKK bound} \label{formula-idea-section} Our formulae for intersection multiplicity (\cref{multiplicity-thm}) and extended BKK bound (\cref{bkk-thm}) are quite different from the formulae arrived in \cite{hurmfels-bern} and \cite[Corollary 1]{rojas-toric} via Huber and Sturmfels' polynomial homotopy arguments, and we do not know of any direct proof of their equivalence. % Huber and Sturmfels' formulae originate from taking points in a generic fiber of the map determined by a non-degenerate system of polynomials, and tracking how many of these approach the origin (for the intersection multiplicity) or the pre-image of the origin (for the extended BKK bound). On the other hand, our formulae originate from the same approach as of Bernstein's proof of the BKK bound: given a non-degenerate system of polynomials $f_1, \ldots, f_n$, we consider the curves in the zero set of $f_2, \ldots, f_n$, and sum up the order at the origin (for the intersection multiplicity) or at points at infinity (for the extended BKK bound) of the restriction of $f_1$ along these curves. On the level of convex geometry, our formulae generalize the following (well-known) property of the mixed volume of $n$ convex bodies $\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n$: \begin{align} \mv(\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n) &= \sum_\omega \omega(\ensuremath{\mathcal{P}}_1)\mv(\ld_\omega(\ensuremath{\mathcal{P}}_2), \ldots, \ld_\omega(\ensuremath{\mathcal{P}}_n)) \tag{$*$} \label{mixed-expression} \end{align} where $\mv$ denotes the mixed volume, the sum on the right hand side is over all `directions' $\omega$, $\omega(\ensuremath{\mathcal{P}}_1)$ is the maximum of $\langle \omega, v \rangle$ over all $v \in \ensuremath{\mathcal{P}}_1$, and $\ld_\omega(\ensuremath{\mathcal{P}}_j)$ is the `leading face' of $\ensuremath{\mathcal{P}}_j$ in the direction of $\omega$. In particular, as in the case of \eqref{mixed-expression}, the expressions on the right hand sides of our formulae from \cref{multiplicity-thm,bkk-thm} are {\em not} symmetric in their arguments. It will be nice to have a convex geometric argument for their symmetric dependence on the input polytopes. \begin{rem} \label{convenient-remark} Combining Kushnirenko's \cite{kush-poly-milnor} results with ours we see that if the Newton diagram of a polynomial $f$ is convenient, then the Milnor number of the singularity at the origin of $f = 0$ is given by Kushnirenko's formula from \cite{kush-poly-milnor} iff $f$ is Milnor non-degenerate\footnote{provided there exist Newton non-degenerate or inner non-degenerate polynomials with the same diagram - which depends on the characteristic of the field.}. In particular, in this case Kushnirenko's and our formulae compute the same number. However, the latter expresses the number as a sum of positive numbers, whereas the former is an alternating sum - it should be interesting to find a direct proof for the equivalence of these formulae. \end{rem} \subsection{Organization} In \cref{example-section} we give some examples of new phenomena that one encounters when extending BKK bound from the torus to the affine space. \Cref{notection} introduces some notations and conventions to be used throughout this article. In \cref{mult-section} we state the solution to the intersection multiplicity problem, which we apply in \cref{milnor-section} to the problem of generic Milnor number. \Cref{milnor-section} also contains some open problems (\cref{open-problem}) on Milnor non-degeneracy. In \cref{affine-section} we state the results on extended BKK bound. \Crefrange{generic-proof-section}{bkk-proof-section} are devoted to the proofs of results from \cref{mult-section,affine-section}. \Cref{bkk-section} contains some technical results on existence and deformations of systems of polynomials which satisfy certain stronger non-degeneracy conditions; we suspect these are well known to experts, but we could not find any reference in the literature. \Cref{technical-section} contains a technical lemma on the algebraic definition of intersection multiplicity which we use in the proofs of computations of intersection multiplicity (\cref{multiplicity-thm}) and the extended BKK bound (\cref{bkk-thm}). In \cref{counter-section} we give some counter examples to \cite[Theorem I]{rojas-wang} and \cite[Affine Point Theorem II]{rojas-toric}. \subsection{Acknowledgments} I thank Pierre Milman and Dmitry Kerner for reading some of the earlier drafts of this article and providing inputs which made it substantially richer and more readable.\\ In a sequel we apply the results of this article to the study of the general `affine B\'ezout problem' of counting number of isolated solutions of an arbitrary (i.e.\ possibly degenerate) system of polynomials: we start from the extended BKK bound as the first approximation, and correct it in a finite number of steps to find the exact number of solutions. Explicit non-degeneracy criteria (such as the ones from this article) are required at each step to detect if the correct bound has been reached. {\small \tableofcontents} \addtocontents{toc}{\protect\setcounter{tocdepth}{2}} \section{Examples of oddities regarding extension of BKK bound to $\ensuremath{\mathbb{K}}^n$}\label{example-section} For polynomials $f_1, \ldots f_n \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ and a subset $W$ of $\ensuremath{\mathbb{K}}^n$, below we write $\multiso{f_1}{f_n}_W$ for the number (counted with multiplicity) of isolated zeroes of $f_1, \ldots, f_n$ which appear on $W$. \begin{example}\label{b-example} Let $g_1, g_2, g_3 \in \ensuremath{\mathbb{K}}[x,y,z]$ be polynomials supported at the tetrahedron $\ensuremath{\mathcal{T}}$ with vertices $(0,0,0), (1,1,0), (0, 1, 1), (1,0,1)$, i.e.\ $g_j = a_j +b_jxy + c_jyz + d_jzx$, $a_j, b_j, c_j, d_j \in \ensuremath{\mathbb{K}}$, $j = 1, 2, 3$. Let \begin{align*} f_j &:= \begin{cases} g_j & \text{if}\ j=1,2,\\ zg_j & \text{if}\ j=3. \end{cases} \end{align*} Then \begin{defnlist} \item \label{b-example-generic-0} If all the coefficients are generic, then the set $V(f_1, f_2, f_3)$ of zeroes of $f_1, f_2, f_3$ on $\ensuremath{\mathbb{K}}^3$ is isolated. Moreover, $V(f_1, f_2, f_3) \subset (\ensuremath{\mathbb{K}}^*)^3$ and Bernstein's theorem implies that \begin{align*} [f_1,f_2,f_3]^{iso}_{\ensuremath{\mathbb{K}}^3} = [g_1,g_2,g_3]^{iso}_{(\ensuremath{\mathbb{K}}^*)^3} = 6\vol(\ensuremath{\mathcal{T}}) = 2. \end{align*} \item \label{b-example-generic-1} On the other hand, if $a_1 = a_2$, $b_1 = b_2$, and the rest of the coefficients are generic, then $\{z = a_1 + b_1xy = 0\}$ is a positive dimensional component of $V(f_1, f_2, f_3)$. However, all the zeroes of $f_1, \ldots, f_3$ on $(\ensuremath{\mathbb{K}}^*)^3$ are still isolated, and Bernstein's theorem guarantees that $[f_1,f_2,f_3]^{iso}_{\ensuremath{\mathbb{K}}^3} = 2$. \item \label{b-example-non-generic} Finally, if $a_1 = a_2 = a_3$, $b_1 = b_2 = b_3$, and the rest of the coefficients are generic, then again $\{z = a_1 + b_1xy = 0\}$ is the positive dimensional component of $V(f_1, f_2, f_3)$. However, \eqref{bkk} fails for the weighted order $\nu$ with weights $(-1,1,2)$ for $(x,y,z)$, and Bernstein's theorem implies that $[f_1,f_2,f_3]^{iso}_{\ensuremath{\mathbb{K}}^3} < 2$. \end{defnlist} \end{example} \begin{example}\label{c-example} Take $f_1, f_2 \in \ensuremath{\mathbb{K}}[x_1, x_2]$ such that the {\em Newton polytope} (see \cref{Newton-definition}) $\ensuremath{\mathcal{P}}_i$ of each $f_i$ contains the origin, the mixed volume $\mv(\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2)$ of $\ensuremath{\mathcal{P}}_1$ and $\ensuremath{\mathcal{P}}_2$ is non-zero, and $f_1, f_2$ are BKK non-degenerate (i.e.\ they satisfy \eqref{bkk}). Let $f_{3,1}, f_{3,2}, f_{4,1}, f_{4,2}$ be arbitrary polynomials in $\ensuremath{\mathbb{K}}[x_1, x_2]$, and set \begin{align*} f_j &:= x_3f_{j,1}(x_1, x_2) + x_4f_{j,2}(x_1,x_2) \in \ensuremath{\mathbb{K}}[x_1, x_2, x_3, x_4],\ j = 3, 4. \end{align*} Then \begin{defnlist} \item If the coefficients of $f_{3,1}, f_{3,2}, f_{4,1}, f_{4,2}$ are generic, then Bernstein's theorem implies that \begin{align*} [f_1,f_2,f_3, f_4]^{iso}_{\ensuremath{\mathbb{K}}^4} = [f_1,f_2,x_3, x_4]^{iso}_{\ensuremath{\mathbb{K}}^4} = \mv(\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2). \end{align*} \item \label{c-example-non-generic} Now fix a common zero $z$ of $f_1,f_2$ in $(\ensuremath{\mathbb{K}}^*)^2$. If $f_{3,1}(z) = f_{4,1}(z) = 0$, then $\{(z_1,z_2, t, 0): t \in \ensuremath{\mathbb{K}}\} \subseteq V(f_1, f_2, f_3, f_4)$, so that $(z_1, z_2, 0,0)$ is no longer an isolated zero of $f_1, f_2, f_3, f_4$. It follows that $[f_1,f_2,f_3, f_4]^{iso}_{\ensuremath{\mathbb{K}}^4} < \mv(\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2)$. \end{defnlist} \end{example} \section{Notations, conventions and remarks} \label{notection} Throughout this article, $\ensuremath{\mathbb{K}}$ is an algebraically closed field of arbitrary characteristic and $n$ is a {\em positive} integer. \begin{defn} \label{Newton-definition} Let $h := \sum_\alpha c_\alpha x^{\alpha} \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. \begin{itemize} \item The {\em support} of $h$ is $\supp(h) := \{\alpha: c_\alpha \neq 0\} \subseteq \ensuremath{\mathbb{Z}}^n$. \item The {\em Newton polytope} $\np(h)$ of $h$ is the convex hull $\conv(\supp(h))$ of $\supp(h)$ in $\ensuremath{\mathbb{R}}^n$. \item The {\em Newton diagram} $\nd(h)$ of $h$ is the union of the compact faces of $(\ensuremath{\mathbb{R}}_{\geq 0})^n + \np(h)$. \item A {\em diagram} in $\ensuremath{\mathbb{R}}^n$ is the Newton diagram of some polynomial in $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. \end{itemize} \end{defn} \begin{notation} We write $[n] := \{1, \ldots, n\}$. For $I \subseteq [n]$ and $k = \ensuremath{\mathbb{K}}$ or $\ensuremath{\mathbb{R}}$, we write \begin{align*} k^I & := \{(x_1, \ldots, x_n) \in k^n: x_i = 0\ \text{if}\ i \not\in I\} \cong k^{|I|} \end{align*} and denote by $\pi_I: k^n \to k^I$ the projection in the coordinates indexed by $I$, i.e.\ \begin{align} \text{the $j$-th coordinate of $\pi_I(x)$}&:= \begin{cases} x_j &\text{if}\ j \in I\\ 0 &\text{if}\ j \not\in I. \end{cases} \label{pi_I} \end{align} For $\Gamma \subseteq \ensuremath{\mathbb{R}}^n$, we write $$\Gamma^I := \Gamma \cap \rr^I$$ We denote by $\ensuremath{\mathbb{K}}^*$ the {\em algebraic torus} $\ensuremath{\mathbb{K}}\setminus \{0\}$, and write \begin{align*} (\kk^*)^I &:= \{\prod_{i\in I} x_i \neq 0\} \cap \kk^I \cong (\ensuremath{\mathbb{K}}^*)^{|I|} \end{align*} Note that $\Kii{\emptyset} = \Kstarii{\emptyset} = \{0\}$. We write $A^I$ for the coordinate ring $\ensuremath{\mathbb{K}}[\kk^I] = \ensuremath{\mathbb{K}}[x_i: i \in I]$ of $\kk^I$. \end{notation} \begin{defn} \label{val-definition} Let $I := \{i_1, \ldots, i_m\} \subseteq [n]$, where $m := |I|$. A {\em monomial valuation} or a {\em weighted order} on $A^I$ is a map $\nu: A^I\setminus \{0\} \to \ensuremath{\mathbb{Z}}$ defined by \begin{align*} \nu(\sum_\alpha c_\alpha x^\alpha) := \min\{v \cdot \alpha: c_\alpha \neq 0\} \end{align*} where $v:= (v_{i_1}, \ldots, v_{i_m}) \in \ensuremath{\mathbb{Z}}^m$ is such that \begin{itemize} \item either $v = (0, \ldots, 0)$ (the trivial valuation), or \item $\gcd(v_{i_1}, \ldots, v_{i_m}) = 1$ (see \cref{gcd-convention} below). \end{itemize} We will often abuse the notation and identify $\nu$ with $v$. Given $I^* \subseteq I$, we say that the {\em center} of $\nu$ is $\Kstarii{I^*}$ if \begin{itemize} \item $\nu_{i_j} \geq 0$ for all $j$, $1 \leq j \leq m$, and \item $I^* = \{i_j: \nu_{i_j} = 0\}$. \end{itemize} In particular, we say that $\nu$ is {\em centered at the origin} if $\nu_{i_j} > 0$ for each $j$, $1 \leq j \leq m$. If there exists $j$ such that $\nu_{i_j} < 0$, then we say that the $\nu$ is {\em centered at infinity}. We write $\scrV^I$ (resp.\ $\scrV^I_0$, $\scrV^I_\infty$) for the space of all monomial valuations (resp.\ monomial valuations centered at the origin, monomial valuations centered at infinity) on $A^I$. We also write $\scrV := \Vii{[n]}$, $\scrV_0 := \Vzeroii{[n]}$, $\scrV_\infty := \Vii{[n]}_\infty$. \end{defn} \begin{defn} A {\em weighted degree} $\omega$ on $A^I$ is the negative of a monomial valuation on $A^I$. We say that $\omega$ is {\em centered at infinity} on $A^I$ if $\omega(x_i) > 0$ for at least one $i \in I$, or equivalently, if $-\omega \in \scrV^I_\infty$. We denote by $\Omega^{I}_\infty$ the space of all weighted degrees centered at infinity on $A^I$, and write $\Omega_\infty := \wtsii{[n]}$. \end{defn} \begin{rem} We want to emphasize that by our definition the greatest common divisor of weights of each non-trivial monomial valuation or weighted degree is $1$. \end{rem} \begin{defn} \label{val-diagram-defn} Let $I \subseteq [n]$, $\nu \in \scrV^I$, $\omega \in \Omega^{I}_\infty$, $\Gamma \subseteq \rr^I$, $f = \sum_\alpha c_\alpha x^\alpha \in A^I$. Define \begin{align*} \nu(\Gamma) &:= \inf\{\nu \cdot \alpha: \alpha \in \Gamma\} \\ \omega(\Gamma) &:= \sup\{\omega \cdot \alpha: \alpha \in \Gamma\} \end{align*} The {\em initial face} (resp.\ {\em leading face}) of $\Gamma$ and the {\em initial form} (resp.\ {\em leading form}) of $f$ corresponding to $\nu$ (resp.\ $\omega$) are defined as \begin{align*} \In_\nu(\Gamma) &:= \{\alpha \in \Gamma: \nu \cdot \alpha = \nu(\Gamma)\} & \ld_\omega(\Gamma) &:= \{\alpha \in \Gamma: \omega \cdot \alpha = \omega(\Gamma)\} \\ \In_\nu(f) &:= \sum_{\alpha\in \In_\nu(\supp(f))} c_\alpha x^\alpha & \ld_\omega(f) &:= \sum_{\alpha\in \ld_\omega(\supp(f))} c_\alpha x^\alpha \end{align*} \end{defn} \begin{rem} Note that if $\Gamma = \emptyset$ (resp.\ if $f = 0$) then $\nu(\Gamma) = \infty$, $\omega(\Gamma) = -\infty$ (resp.\ $\nu(f) = \infty$, $\omega(f) = -\infty$). Also, if $\nu$ is the trivial valuation, then $\In_\nu(\Gamma) = \Gamma$ and $\In_\nu(f) = f$ for all $\Gamma$ and $f$. \end{rem} \begin{convention} \label{gcd-convention} Given $k_1, \ldots, k_l \in \ensuremath{\mathbb{Z}}$, $d := \gcd(k_1, \ldots, k_l)$ is defined as follows: \begin{itemize} \item if each $k_j = 0$, then $d = 0$. \item otherwise $d$ is the {\em positive} greatest common divisor of the non-zero elements in $\{k_1, \ldots, k_l\}$. \end{itemize} \end{convention} \begin{convention} \label{coconvention} An assertion of the form ``codimension of $X$ in $Y$ is $m$'', where $X$ is a subvariety of $Y$ and $\dim(Y) < m$, means that $X = \emptyset$. \end{convention} \begin{defn} \label{mixed-defn} Let $\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n$ be convex polytopes in $\ensuremath{\mathbb{R}}^n$. The $n$-dimensional {\em mixed volume} $\mv(\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ of $\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n$, we denote the coefficient of $\lambda_1 \cdots \lambda_n$ in the homogeneous polynomial $\vol(\lambda_1\ensuremath{\mathcal{P}}_1 + \cdots + \lambda_n\ensuremath{\mathcal{P}}_n)$. If $\ensuremath{\mathcal{P}}_1 + \cdots + \ensuremath{\mathcal{P}}_k$ has dimension $\leq k$, $1 \leq k \leq n$, then we write $MV(\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_k)$ to denote the $k$-dimensional mixed volume of $\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_k$. \end{defn} \begin{rem} Note that according to our definition $\mv(\ensuremath{\mathcal{P}}, \ldots, \ensuremath{\mathcal{P}}) = n!\vol(\ensuremath{\mathcal{P}})$. There is another convention present in the literature in which mixed volume is defined to be $1/n!$ times the mixed volume from our definition. We chose to follow the convention that makes our formulae simpler. \end{rem} \section{Intersection multiplicity} \label{mult-section} In this section we state our results on intersection multiplicity at the origin of generic systems of $n$ polynomials or power series in $n$ variables. \Cref{mult-support-solution} characterizes collections of diagrams which admit polynomials with finite intersection multiplicity at the origin. \Cref{multiplicity-thm} gives a formula for the minimum intersection multiplicity and \cref{multiplicity-explanation} describes the meaning of the terms that appear in the formula. The necessary and sufficient conditions for the attainment of the minimum intersection multiplicity are described in \cref{generic-thm}. \Cref{weak-lemma} gives an equivalent criterion with fewer conditions. \begin{defn} \label{Isolated-defn} Let $\mathscr{G} := (\Gamma_1, \ldots, \Gamma_m)$ be an ordered collection of diagrams in $\ensuremath{\mathbb{R}}^n$ and $I$ be a non-empty subset of $[n]$. Define \begin{align} \scrN^{I}_{\mscrG} &:= \{j \in [m]: \Gamma^I_j \neq \emptyset\} \label{NiG} \end{align} We say that \begin{itemize} \item $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ are {\em $\mathscr{G}$-admissible} iff $\nd(f_j) = \Gamma_j$, $1 \leq j \leq m$. \item $\mathscr{G}$ is {\em $I$-isolated at the origin} iff $\left|\scrN^{I}_{\mscrG}\right| \geq |I|$. \end{itemize} \end{defn} The choice of the term {\em isolated} in \cref{Isolated-defn} is motivated by the following lemma: \begin{lemma} \label{isolated-lemma} Let $\mathscr{G} := (\Gamma_1, \ldots, \Gamma_m)$ be a collection of diagrams in $\ensuremath{\mathbb{R}}^n$ and $I \subseteq [n]$. Assume $0 \not\in \Gamma_i$ for each $j$, $1 \leq j \leq m$. \begin{enumerate} \item If $\mathscr{G}$ is not $I$-isolated at the origin, then for all $\mathscr{G}$-admissible $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, $V(f_1, \ldots, f_m) \cap \kk^I$ is non-empty with positive dimensional components. \item If $\mathscr{G}$ is $I$-isolated at the origin, then $V(f_1, \ldots, f_m) \cap (\kk^*)^I$ is isolated for generic $\mathscr{G}$-admissible $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. \end{enumerate} \end{lemma} \begin{proof} Note that $ \{j : \Gamma^I_J \neq \emptyset\} = \{j: f_j|_{\kk^I} \not \equiv 0\}$. If $\mathscr{G}$ is not $I$-isolated at the origin, then it follows that $V(f_1, \ldots, f_m) \cap \kk^I$ is defined by less than $|I|$ elements, which implies the first assertion (note that the assumption $0 \not\in \bigcup_j \Gamma_j$ is necessary). The second assertion follows from Bernstein's theorem. \end{proof} \begin{defn} \label{intersection-multiplicity} The {\em intersection multiplicity} $\multzero{f_1}{f_n}$ at the origin of $f_1, \ldots, f_n \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ is the dimension (as a vector space over $\ensuremath{\mathbb{K}}$) of the quotient ring of $\ensuremath{\mathbb{K}}[[x_1, \ldots, x_n]]$ modulo the ideal generated by $f_1, \ldots, f_n$. Let $\mathscr{G} := (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$. Define \begin{align*} \multzero{\Gamma_1}{\Gamma_n} := \min\{\multzero{f_1}{f_n}: f_1, \ldots, f_n\ \text{are $\mathscr{G}$-admissible polynomials in $\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$}\} \end{align*} \end{defn} Note that $0 \leq \multzero{\Gamma_1}{\Gamma_n} \leq \infty$. \Cref{isolated-lemma} immediately gives a combinatorial way to determine if $\multzero{\Gamma_1}{\Gamma_n}$ is zero or infinity: \begin{cor}[cf.\ {\cite[Lemma 2]{rojas-toric}, \cite[Proposition 5]{herrero}}] \label{mult-support-solution} Let $\Gamma_1, \ldots, \Gamma_n$ be diagrams in $\ensuremath{\mathbb{R}}^n$. \begin{enumerate} \item $\multzero{\Gamma_1}{\Gamma_n} = 0$ iff $0 \in \Gamma_i$ for some $i$, $1 \leq i \leq n$. \item Assume $\multzero{\Gamma_1}{\Gamma_n} \neq 0$. Then $\multzero{\Gamma_1}{\Gamma_n}< \infty$ iff $(\Gamma_1, \ldots, \Gamma_n)$ is $I$-isolated at the origin for all $I \subseteq [n]$. \qed \end{enumerate} \end{cor} \Cref{multiplicity-thm} below describes the formula for the minimum intersection multiplicity; it uses \cref{mult-star-defn} which in turn uses notions from \cref{val-definition,val-diagram-defn}. \begin{defn} \label{mult-star-defn} Given polyhedra $\Gamma_1, \ldots, \Gamma_n$ in $\ensuremath{\mathbb{R}}^n$, define \begin{align} \multstar{\Gamma_1}{\Gamma_n} &:= \sum_{\nu \in \scrV_0} \nu(\Gamma_1) ~ \mv(\In_\nu(\Gamma_2), \ldots, \In_\nu(\Gamma_n)) \label{mult-star} \end{align} \end{defn} \begin{thm}[Formula for generic intersection multiplicity]\label{multiplicity-thm} Let $\Gamma_1, \ldots, \Gamma_n$ be diagrams in $\ensuremath{\mathbb{R}}^n$. Let \begin{align} \ensuremath{\mathcal{I}}_{\mathscr{G},1} := \{I \subseteq [n]: I \neq \emptyset,\ |\scrN^{I}_{\mscrG}| = |I|,\ 1 \in \scrN^{I}_{\mscrG}\} \label{I1-list} \end{align} where $\scrN^{I}_{\mscrG}$ is as in \eqref{NiG}. Assume $0 < \multzero{\Gamma_1}{\Gamma_n} < \infty$. Then with notations as in \cref{mult-star-defn}, \begin{align} \multzero{\Gamma_1}{\Gamma_n} &= \sum_{I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}} \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \times \multstar{\Gamma^{I}_1, \Gamma^{I}_{j'_1}}{\Gamma^{I}_{j'_{k-1}}} \label{multiplicity-formula} \end{align} where for each $I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}$, \begin{itemize} \item $I' := [n]\setminus I$, $k := |I|$, \item $j_1, \ldots, j_{n-k}$ are elements of $[n]\setminus \scrN^{I}_{\mscrG}$, \item $j'_1, \ldots, j'_{k-1}$ are elements of $\scrN^{I}_{\mscrG}\setminus \{1\}$. \end{itemize} \end{thm} \begin{proof} See \cref{multiplicity-proof-section}. \end{proof} \begin{convention} \label{formula-convention} In this article, in particular in \eqref{mult-star} and \eqref{multiplicity-formula}, we use the following conventions: \begin{defnlist} \item $\infty$ times $0$ is interpreted as $0$. \item Empty intersection products and mixed volumes are defined as $1$. In particular, the term $\mv(\In_\nu(\Gamma_2), \ldots, \In_\nu(\Gamma_n))$ from \eqref{mult-star} in $n = 1$ case, and the term $ \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})}$ from \eqref{multiplicity-formula} in $I = [n]$ case are both defined as $1$. \end{defnlist} \end{convention} \begin{example} \label{ex0} Let $f_1 = a_1x + b_1y + c_1z$, $f_2 = a_2x^3 + b_2xz^2 + c_2y^3 + d_2yz^2$, $f_3 = a_3x^2 + b_3xz^2 + c_3y^2 + d_3yz^2$, where $a_i, b_i, c_i, d_i$'s are generic elements of $\ensuremath{\mathbb{K}}$. We use \eqref{multiplicity-formula} to compute $[f_1,f_2,f_3]_0$. It is straightforward to check that $\ensuremath{\mathcal{I}}_{\mathscr{G},1} = \{\{1,2,3\},\{3\}\}$. From \eqref{multiplicity-formula} we see that \begin{align*} [f_1,f_2,f_3]_0 &= [\Gamma_1, \Gamma_2, \Gamma_3]^*_0 + [\pi_{\{1,2\}}(\Gamma_2), \pi_{\{1,2\}}(\Gamma_3)]_0 ~ [\Gamma^{\{3\}}_1]^*_0 \end{align*} Note that $\pi_{\{1,2\}}(\Gamma_j)$, for each $j \in \{2,3\}$, is the diagram of a linear polynomial with no constant terms. Therefore $[\pi_{\{1,2\}}(\Gamma_2), \pi_{\{1,2\}}(\Gamma_3)]_0 = 1$. It is straightforward to see that $\Gamma_2 + \Gamma_3$ has two facets with inner normals in $(\ensuremath{\mathbb{R}}_{> 0})^3$, and these inner normals are $\nu_1 := (1,1,1)$ and $\nu_2 := (2,2,1)$. Then \begin{align*} [f_1,f_2,f_3]_0 &= \nu_1(\Gamma_1)\mv(\In_{\nu_1}(\Gamma_2), \In_{\nu_1}(\Gamma_3)) + \nu_2(\Gamma_1)\mv(\In_{\nu_2}(\Gamma_2), \In_{\nu_2}(\Gamma_3)) + 1 \cdot \ord_z(f_1|_{x=y=0}) \\ &= 1 \cdot 4 + 1 \cdot 1 + 1 = 6 \end{align*} \end{example} \begin{rem} [`Explanation' of the right hand side of \eqref{multiplicity-formula}] \label{multiplicity-explanation} Use the notations from \cref{multiplicity-thm}. Let $f_1, \ldots, f_n$ be generic $(\Gamma_1, \ldots, \Gamma_n)$-admissible polynomials. \begin{defnlist} \item For each $I \subseteq [n]$, note that $j \in \scrN^{I}_{\mscrG}$ iff $\kk^I \not\subseteq V(f_j)$. Since $0 < \multzero{\Gamma_1}{\Gamma_n} < \infty$, it follows that \begin{itemize} \item $|\scrN^{I}_{\mscrG}| \geq |I|$ for each $I \subseteq [n]$. \item If $|\scrN^{I}_{\mscrG}| = |I|$ then there are $n-k$ polynomials (where $k = |I|$) $f_{j_1}, \ldots, f_{j_{n-k}}$ which identically vanish on $\kk^I$. \end{itemize} \item $ \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})}$ is precisely the intersection multiplicity of the hypersurfaces defined by $f_{j_1}, \ldots, f_{j_{n-k}}$ along $\kk^I$. \item $\multstar{\Gamma^{I}_1, \Gamma^{I}_{j'_1}}{\Gamma^{I}_{j'_{k-1}}} $ is the sum (counted with multiplicity) of order of vanishing at the origin of restrictions of $f_1$ at the branches of the curves on $(\kk^*)^I$ defined by $f_{j'_1}, \ldots, f_{j'_{k-1}}$. \end{defnlist} \end{rem} \begin{rem} \label{mult-monotone} It is not too hard to show that under \cref{formula-convention} identity \eqref{multiplicity-formula} is equivalent to the following identity: \begin{align} \multzero{\Gamma_1}{\Gamma_n} &= \sum_{\scriptsize \begin{array}{c} I \in [n] \\ I \neq \emptyset \end{array}} \sum_{\scriptsize \begin{array}{c} 1 \not\in J \subseteq [n]\\ |J| = n - |I| \end{array}} \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \times \multstar{\Gamma^{I}_1, \Gamma^{I}_{j'_1}}{\Gamma^{I}_{j'_{k-1}}} \label{multiplicity-formula'} \end{align} Given a diagram $\Gamma$ and a point $\alpha$ in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$, we say that $\alpha$ is {\em below} $\Gamma$ iff there exists $\nu \in \scrV_0$ such that $\nu(\alpha) < \nu(\Gamma)$. Identity \eqref{multiplicity-formula'} provides an easy way to see the monotonicity of intersection multiplicity, i.e.\ the following fact (for which we do not know of any simple algebraic proof): \begin{align*} \parbox{.9\textwidth}{\em If no point of $\Gamma'_j$ lies below $\Gamma_j$ for all $j$, $1 \leq j \leq n$, then $\multzero{\Gamma'_1}{\Gamma'_n} \geq \multzero{\Gamma_1}{\Gamma_n}$.} \end{align*} Indeed, it suffices to prove the assertion for the case that $\Gamma'_j = \Gamma_j$ for $j = 2, \ldots, n$. Expanding the second term in the sum on the right hand side of \eqref{multiplicity-formula'} gives: \begin{align*} \multzero{\Gamma_1}{\Gamma_n} &= \sum_{\scriptsize \begin{array}{c} I \in [n] \\ I \neq \emptyset \end{array}} \sum_{\scriptsize \begin{array}{c} 1 \not\in J \subseteq [n]\\ |J| = n - |I| \end{array}} \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \sum_{\nu \in \scrV^I_0} \nu(\Gamma^I_1) ~ \mv(\In_\nu(\Gamma^I_{j'_1}), \ldots, \In_\nu(\Gamma^I_{j'_{k-1}})) \end{align*} Since $\nu(\Gamma'^I_1) \geq \nu(\Gamma^I_1)$ for all $\nu \in \scrV^I_0$ and all $I$, the assertion follows. \end{rem} We now describe the non-degeneracy condition for the intersection multiplicity at the origin. \begin{defn} \label{generic-defn} \mbox{} \begin{itemize} \item We say that polynomials $f_1, \ldots, f_m$, $m \geq n$, in $(x_1, \ldots, x_n)$ are {\em BKK non-degenerate at the origin} iff $\In_\nu(f_1), \ldots, \In_\nu(f_m)$ have no common zero in $(\kk^*)^n$ for all $\nu \in \scrV_0$. \item Let $\mathscr{G}:= (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$ such that $\multzero{\Gamma_1}{\Gamma_n} < \infty$. We say that polynomials $f_1, \ldots, f_n$ are {\em $\mathscr{G}$-non-degenerate} iff they are $\mathscr{G}$-admissible and $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK non-degenerate at the origin for every non-empty subset $I$ of $[n]$. \end{itemize} \end{defn} \begin{thm} [Non-degeneracy for intersection multiplicity] \label{generic-thm} Let $\mathscr{G}:=(\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$ such that $\multzero{\Gamma_1}{\Gamma_n} < \infty$. Let $f_1, \ldots, f_n$ be $\mathscr{G}$-admissible polynomials. Then the following are equivalent: \begin{enumerate} \item $\multzero{f_1}{f_n} = \multzero{\Gamma_1}{\Gamma_n}$. \item $f_1, \ldots, f_n$ are $\mathscr{G}$-non-degenerate. \end{enumerate} \end{thm} \begin{proof} See \cref{generic-proof}. \end{proof} \begin{rem} Note that we define $\mathscr{G}$-non-degeneracy only in the case that $\multzero{\Gamma_1}{\Gamma_n} < \infty$. In fact if $\multzero{\Gamma_1}{\Gamma_n} = \infty$, then every system of $\mathscr{G}$-admissible polynomials violates the condition from \cref{generic-defn}. In view of the non-degeneracy conditions for the extended BKK bound in \cref{affine-section}, it seems that for the correct notion of non-degeneracy which would extend to $\multzero{\Gamma_1}{\Gamma_n} = \infty$ case, one has to \begin{itemize} \item replace ``$\In_\nu(f_1), \ldots, \In_\nu(f_m)$ have no common zero in $(\kk^*)^n$'' by ``the subvariety of $(\kk^*)^n$ defined by $\In_\nu(f_i)$'s has the {\em smallest possible} dimension,'' \item and add a set of conditions analogous to property \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy (\cref{generic-bkk-defn}) to ensure `correct infinitesimal intersections' of the subvarieties in the torus determined by restrictions of the polynomials to different coordinate subspaces. \end{itemize} \end{rem} In the proof of \cref{generic-thm} in \cref{generic-proof-section} we use the following result which shows that while testing for $\mathscr{G}$-non-degeneracy one can omit certain coordinate subspaces of $\ensuremath{\mathbb{K}}^n$. \begin{lemma} \label{weak-lemma} Let $\mathscr{G}:= (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$. Define \begin{align} \scrN_{\mscrG} &:= \{I \subseteq [n]:\ I \neq \emptyset,\ |\scrN^{I}_{\mscrG}| = |I|\}, \label{NG} \end{align} where $\scrN^{I}_{\mscrG}$ is as in \eqref{NiG}. Let $f_1, \ldots, f_n$ be $\mathscr{G}$-admissible polynomials in $(x_1, \ldots, x_n)$. Assume $\multzero{\Gamma_1}{\Gamma_n} < \infty$. Then the following are equivalent: \begin{enumerate} \item $f_i$'s are $\mathscr{G}$-non-degenerate. \item $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK non-degenerate at the origin for every $I \in \scrN_{\mscrG}$. \end{enumerate} \end{lemma} \begin{proof} This is precisely \cref{weak-lemma2}. \end{proof} \section{Milnor number} \label{milnor-section} In this section we apply the results from \cref{mult-section} to study Milnor number at the origin of generic polynomials or power series. We define the notion of (partial) Milnor non-degeneracy (\cref{partial-defn}) and show in \cref{milnor-thm} that it is the correct notion of non-degeneracy in this context. As an immediate corollary we give an effective solution to \cref{kush-problem} (\cref{kush-solution}). We also recall (\cref{inndefinition}) the classical notions of non-degeneracy, give examples to illustrate that they are not necessary for the minimality of Milnor number (\cref{Milnor-not-Newton,Milnor-not-inner}), and give direct arguments (\cref{newton-wall-lemma}) to show that they are special cases of partial Milnor non-degeneracy (at least when the origin is an isolated singularity). We end this section with statements of some open problems. \subsection{(Partial) Milnor non-degeneracy} \begin{defn} \label{partial-defn} For a diagram $\Gamma \subseteq \ensuremath{\mathbb{R}}^n$, let $\dGammad{i}$ be the Newton diagram of $\dgd{i}$ for a generic polynomial $g \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ with Newton diagram $\Gamma$. Let $f \in\ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ be a polynomial with Newton diagram $\Gamma$ and $\Gamma_j$ be the Newton diagram of $\dfd{j}$, $1 \leq j \leq n$. We say that \begin{itemize} \item $f$ is {\em partially Milnor non-degenerate} iff $\multzero{\Gamma_1}{\Gamma_n} < \infty$ and $\dfd{1}, \ldots, \dfd{n}$ are $(\Gamma_1, \ldots, \Gamma_n)$-non-degenerate (see \cref{generic-defn}). \item $f$ is {\em Milnor non-degenerate} iff $f$ is partially Milnor non-degenerate and $\multzero{\Gamma_1}{\Gamma_n} = \multzero{\dGammad{1}}{\dGammad{n}}$. \end{itemize} \end{defn} \begin{rem} \label{milnor-remark} \mbox{} \begin{defnlist} [ ref= \therem(\alph{defnlisti})] \item $\dGammad{i}$ depends on the characteristic of $\ensuremath{\mathbb{K}}$. \item In positive characteristic, it is possible for polynomials with the same Newton diagram to have distinct Newton diagrams for a partial derivative. In particular, it is possible that $\nd(f) = \Gamma$, but $\nd(\dfd{i}) \neq \dGammad{i}$ for some $i$, $1 \leq i \leq n$. \item In zero characteristic partial Milnor non-degeneracy is equivalent to Milnor non-degeneracy. \end{defnlist} \end{rem} Let $f, \Gamma_1, \ldots, \Gamma_n$ be as in \cref{partial-defn}. Let $\ensuremath{\mathcal{A}}$ be the space of polynomials $g$ such that $\supp(g) \subseteq \supp(f)$. Recall the definition of $\mu(f)$ from \cref{milnor-footnote}. \begin{thm}[Non-degeneracy for Milnor number at the origin] \label {milnor-thm} \mbox{} \begin{enumerate} \item \label{milnor-bound} $\mu(f) \geq \multzero{\Gamma_1}{\Gamma_n} \geq \multzero{\dGammad{1}}{\dGammad{n}}$. \item \label{milnor-existence} The set of polynomials $g \in \ensuremath{\mathcal{A}}$ which satisfy $\mu(g) = \multzero{\Gamma_1}{\Gamma_n}$ contains a non-empty open subset of $\ensuremath{\mathcal{A}}$. \item \label{milnor-partial-genericness} The following are equivalent: \begin{enumerate} \item $f$ is partially Milnor non-degenerate. \item $\mu(f) = \multzero{\Gamma_1}{\Gamma_n} < \infty$. \end{enumerate} \item \label{milnor-genericness} The following are equivalent: \begin{enumerate} \item $f$ is Milnor non-degenerate. \item $\mu(f) = \multzero{\dGammad{1}}{\dGammad{n}} < \infty$. \end{enumerate} \end{enumerate} \end{thm} \begin{proof} Assertions \eqref{milnor-bound}, \eqref{milnor-partial-genericness} and \eqref{milnor-genericness} follow from \cref{mult-support-solution,multiplicity-thm,generic-thm,mult-monotone}. We now prove assertion \eqref{milnor-existence}. Without any loss of generality we may assume that $\multzero{\Gamma_1}{\Gamma_n} < \infty$. Let $I \subseteq [n]$ and $\nu \in \scrV^I_0$. It suffices to show that there is an open subset $\ensuremath{\mathcal{U}}$ of $\ensuremath{\mathcal{A}}$ such that $V(\In_\nu(\dgd{1}|_{\kk^I}), \ldots, \In_\nu(\dgd{n}|_{\kk^I})) \cap (\kk^*)^n= \emptyset$ for all $g \in \ensuremath{\mathcal{U}}$.\\ Assume without loss of generality that \ $I = \{1, \ldots, k\}$. Take a generic $g \in \ensuremath{\mathcal{A}}$. Due to \cref{weak-lemma} we may also assume that $\dgd{j}|_{\kk^I} \not\equiv 0$ for precisely $k$ values of $j$; denote them by $1 \leq j_1 < \cdots < j_k \leq n$. Write \begin{align*} g &= g_0(x_1, \ldots, x_k) + \sum_{i = k+1}^n x_ig_i(x_1, \ldots, x_k) \\ &\qquad\quad + ~ \text{terms with quadratic or higher order in $(x_{k+1}, \ldots, x_n)$} \end{align*} If $g_0 \equiv 0$, then each $j_i > k$, and \[ V(\In_\nu(\dgd{1}|_{\kk^I}),\ldots, \In_\nu(\dgd{n}|_{\kk^I}))= V(\In_\nu(g_{j_1}), \ldots, \In_\nu(g_{j_k})) \] Since $g_{j_l}$'s are independent of each other, and since dimension of the (convex hull of the) sum of supports of $\In_\nu(g_{j_l})$'s is smaller than $k$, it follows that for generic $g \in \ensuremath{\mathcal{A}}$, $V(\In_\nu(g_{j_1}), \ldots, \In_\nu(g_{j_k})) \cap (\kk^*)^n = \emptyset$ (\cref{herrero}), as required. So assume $g_0 \not\equiv 0$. Then for each $i$, $1 \leq i \leq k$, if $\dx{\In_\nu(g_0)}{x_i} \not\equiv 0$, then $\In_\nu(\dgd{i}) = \dx{\In_\nu(g_0)}{x_i} $. If $z \in V(\In_\nu(\dgd{1}|_{\kk^I}),\ldots, \In_\nu(\dgd{n}|_{\kk^I})) \cap (\kk^*)^n$, then it follows that \begin{prooflist} \item \label{zero-partial-x} $(\dx{\In_\nu(g_0)}{x_i})(z) = 0$, $1 \leq i \leq k$. \end{prooflist} Let $\Lambda$ be the `smallest' linear subspace of $\ensuremath{\mathbb{R}}^n$ such that the support of $\In_\nu(g_0)$ is contained in a translation of $\Lambda$; let $l := \dim(\Lambda)$. Choose a basis $\beta_1, \ldots, \beta_k$ of $\ensuremath{\mathbb{Z}}^k$ such that \begin{prooflist}[resume] \item $\beta_1, \ldots, \beta_l$ is a basis of $\Lambda$. \item $\beta_1, \ldots, \beta_{k-1}$ is a basis of $\nu^\perp := \{\beta \in \ensuremath{\mathbb{Z}}^k: \nu \cdot \beta = 0\}$. \end{prooflist} Let $y_i := x^{\beta_i}$, $1 \leq i \leq k$. Then $(y_1, \ldots, y_k)$ is a system of coordinates on $(\kk^*)^I$. Observation \ref{zero-partial-x} implies that \begin{prooflist}[resume] \item \label{zero-partial-y} $(\dx{\In_\nu(g_0)}{y_i})(z) = 0$, $1 \leq i \leq k$. \end{prooflist} In $(y_1, \ldots, y_k)$-coordinates, $\In_\nu(g_0)$ is of the form $y_k^dg^*_0(y_1, \ldots, y_l)$, where $d := \nu(g_0) > 0$. Observation \ref{zero-partial-y} implies that \begin{prooflist}[resume] \item $g^*_0(z) = (\dx{g^*_0}{y_1})(z) = \cdots = (\dx{g^*_0}{y_l})(z) = 0$. \end{prooflist} An appropriate Bertini-type theorem in arbitrary characteristics, e.g.\ \cite[Theorem 3.1]{jelonek-bertini} implies that for generic $g \in \ensuremath{\mathcal{A}}$, $V(g^*_0, \dx{g^*_0}{y_1},\ldots, \dx{g^*_0}{y_l}) \cap (\kk^*)^n = \emptyset $. This completes the proof of assertion \eqref{milnor-existence}. \end{proof} Let $\Sigma \subseteq (\ensuremath{\mathbb{Z}}_{\geq 0})^n$. For each $i = 1, \ldots, n$, define $\Gamma_i$ to be the Newton diagram of $\dfd{i}$ for any $f$ with $\supp(f) = \Sigma$. The following result, which is immediate from \cref{milnor-thm}, combined with \cref{mult-support-solution,multiplicity-thm}, gives an effective solution to \cref{kush-problem}. \begin{cor}[Solution to \cref{kush-problem}] \label{kush-solution} The following are equivalent: \begin{enumerate} \item there exists a power series $f$ supported at $\Sigma$ such that $\mu(f) < \infty$, \item $\multzero{\Gamma_1}{\Gamma_n} < \infty$. \end{enumerate} In the case that either of these conditions holds, the minimum value of all $\mu(f)$ such that $\supp(f) = \Sigma$ is precisely $\multzero{\Gamma_1}{\Gamma_n}$. \qed \end{cor} \subsection{Relation between Milnor non-degeneracy and the classical non-degeneracy conditions} \begin{defn}[(Inner) Newton non-degeneracy {\cite[Section 3]{boubakri-greuel-markwig}}] \label{inndefinition} Let $f = \sum_\alpha c_\alpha x^\alpha \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. We write $\ensuremath{\mathfrak{j}}(f)$ for the ideal generated by the partial derivatives of $f$. A {\em $C$-polytope} is a diagram containing a point with positive coordinates on every coordinate axis. If no point of $\supp(f)$ lies `below'\footnote{See the sentence immediately following identity \eqref{multiplicity-formula'}} $\ensuremath{\mathcal{P}}$, then for a face $\Delta$ of $\ensuremath{\mathcal{P}}$, we write $\In_\Delta(f) := \sum_{\alpha \in \Delta} c_\alpha x^\alpha$. An {\em inner face} of a polytope $\ensuremath{\mathcal{P}}$ is a face not contained in any coordinate subspace. We say that \begin{itemize} \item $f$ is {\em Newton non-degenerate} iff $\ensuremath{\mathfrak{j}}(\In_\nu(f))$ has no zero in $(\kk^*)^n$ for each $\nu \in \scrV_0$. \item $\nd(f)$ is {\em convenient} if it is a $C$-polytope. \item $f$ is {\em inner Newton non-degenerate} with respect to a $C$-polytope $\ensuremath{\mathcal{P}}$ iff \begin{itemize} \item no point in $\supp(f)$ lies `below' $\ensuremath{\mathcal{P}}$. \item for every inner face $\Delta$ of $\ensuremath{\mathcal{P}}$ and for every non-empty subset $I$ of $[n]$, \begin{align*} \Delta^I \neq \emptyset \ensuremath{\Rightarrow} V(\ensuremath{\mathfrak{j}}(\In_\Delta(f))) \cap (\kk^*)^I = \emptyset \end{align*} \end{itemize} \item $f$ is {\em inner Newton non-degenerate} iff it is inner Newton non-degenerate with respect to a $C$-polytope $\ensuremath{\mathcal{P}}$. \end{itemize} \end{defn} \begin{example}[Newton non-degeneracy is not necessary for Milnor non-degeneracy] \label{Milnor-not-Newton} Let $f := x_1 + (x_2 + x_3)^q$, where $q \geq 2$ is not divisible by $\charac(\ensuremath{\mathbb{K}})$. Then $\np(f)$ is convenient and $f$ is not Newton non-degenerate. However, $f$ is Milnor non-degenerate with $\mu(f) = 1$. \end{example} \begin{example}[Inner Newton non-degeneracy is not necessary for Milnor non-degeneracy] \label{Milnor-not-inner} Let $f := x_3^3 + x_1(x_1 + x_2)^2 + x_2^6 + x_2^4x_3$. Note that $\nd(f)$ is convenient. Let $\Delta$ be the facet of $\nd(f)$ with inner normal $(1,1,1)$. Then $\Delta$ is an inner face of $\nd(f)$ and $\In_\Delta(f) = x_3^3 + x_1(x_1 + x_2)^2$. it is straightforward to check that for all but finitely many characteristics (including zero), \begin{itemize} \item The origin is an isolated singular point of $V(f)$. \item If $I = \{1,2\}$, then $\Delta^I \neq \emptyset$ and $ V(\ensuremath{\mathfrak{j}}(\In_\Delta(f))) \cap (\kk^*)^I \neq \emptyset$. It follows that $f$ is not inner Newton non-degenerate. \item $f$ is Milnor non-degenerate. \end{itemize} \end{example} \begin{example} [Diagram which admits no (inner) Newton non-degenerate polynomial] \label{no-(I)NND} If $\charac(\ensuremath{\mathbb{K}}) > 0$ and $\Gamma$ is any diagram which has a vertex all of whose coordinates are divisible by $\charac(\ensuremath{\mathbb{K}})$, then every polynomial with Newton diagram $\Gamma$ is (inner) Newton degenerate. \end{example} \begin{defn} \label{compatible-defn} Let $I \subseteq [n]$, $\nu \in \scrV^I_0$, $\nu' \in \scrV_0$. We say that $\nu'$ is {\em compatible with} $\nu$ iff $\nu'|_{A^I}$ is proportional to $\nu$. \end{defn} We now give direct proofs that the classical non-degeneracy conditions are special cases of partial Milnor non-degeneracy, at least when $\mu(f) < \infty$. We use the following lemma, which is straightforward to verify. \begin{lemma} \label{non-empty-lemma} Let $\Gamma_1, \ldots, \Gamma_m$ be diagrams in $\ensuremath{\mathbb{R}}^n$, $I \subseteq [n]$ and $\nu \in \scrV^I_0$. Let $\nu' \in \scrV_0$ be such that \begin{enumerate} \item $\nu'$ is compatible with $\nu$ (\cref{compatible-defn}), \item $\nu'(x_{i'})/\max\{\nu(x_i): i \in I\} \gg 1$ for all $i' \not \in I$. \end{enumerate} Then $\In_{\nu'}(\Gamma_j) = \In_{\nu}(\Gamma^I_j)$ for all $j \in [m]$ such that $\Gamma^I_j \neq \emptyset$ .\qed \end{lemma} \begin{prop} \label{newton-wall-lemma} In each of the following cases $f$ is partially Milnor non-degenerate: \begin{enumerate} \item \label{newton-wall-1} $\nd(f)$ is convenient and $f$ is Newton non-degenerate. \item \label{newton-wall-2} $f$ is Newton non-degenerate and $\mu(f) < \infty$. \item \label{newton-wall-3} $f$ is inner Newton non-degenerate. \end{enumerate} \end{prop} \begin{proof} Note that the first assertion is a special case of the second one. Nonetheless, we start with a direct proof of the first assertion since it is easier to see. Let $\Gamma := \nd(f)$ and $\Gamma_i := \nd(\dfd{i})$, $1 \leq i \leq n$. Pick a non-empty subset $I$ of $[n]$ and $\nu \in \scrV^I_0$. Assume $\nd(f)$ is convenient and $f$ is Newton non-degenerate. Since $\nd(f)$ is convenient, $\Gamma^I \neq \emptyset$. Therefore we can find $\nu' \in \scrV_0$ such that $\nu'$ is compatible with $\nu$ and $\In_{\nu'}(\Gamma) \subset \rr^I$. Then $\In_{\nu'}(f)$ depends only on $(x_i: i \in I)$. Since $f$ is Newton non-degenerate, it follows that $\partial(\In_{\nu'}(f))/\partial x_i$, $i \in I$, do not have any common zero in $(\kk^*)^n$. Now note that for each $i \in I$ such that $\partial(\In_{\nu'}(f))/\partial x_i$ is not identically zero, \[ \partial(\In_{\nu'}(f))/\partial x_i = \partial(\In_{\nu}(f|_{\kk^I}))/\partial x_i = \In_{\nu}((\dfd{i})|_{\kk^I}) \] This implies that $\In_{\nu}((\dfd{i})|_{\kk^I})$, $i \in I$, do not have any common zero in $(\kk^*)^I$, as required for partial Milnor non-degeneracy of $f$. \\ Now we prove the second assertion. Assume $\mu(f) < \infty$ and $f$ is {\em not} partially Milnor non-degenerate. It suffices to show that $f$ is not Newton non-degenerate. Pick $I \subseteq [n]$ and $\nu \in \scrV^I_0$ such that $\In_\nu(\dfd{1}|_{\kk^I}), \ldots, \In_\nu(\dfd{n}|_{\kk^I})$ have a common zero in $(\kk^*)^n$. We may assume that $I = \{1, \ldots, k\}$, $1 \leq k \leq n$. At first consider the case that $f|_{\kk^I} \not\equiv 0$. Pick $\nu' \in \scrV_0$ such that \begin{itemize} \item $\nu'$ satisfies \cref{non-empty-lemma} with $m = n$ and $\Gamma_i = \nd(\dfd{i})$, $i = 1, \ldots, n$, \item $\In_{\nu'}(f) \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_k]$. \end{itemize} Then $\ensuremath{\mathfrak{j}}(\In_{\nu'}(f))$ have a common zero in $(\kk^*)^n$, so that $f$ is not Newton non-degenerate, as required. Now assume that $f|_{\kk^I} \equiv 0$. Without any loss of generality we may assume that $\dfd{i}|_{\kk^I} \not\equiv 0$ iff $i = k+1, \ldots, l$, $k < l \leq n$. Then we can write $$f = \sum_{i= k+1}^{l} x_if_i + \sum_{i > j >k} x_ix_jf_{i,j}$$ where $f_{k+1}, \ldots, f_{n'} \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_k]$ and $f_{i,j} \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. It is straightforward to check that there are positive integers $m_{k+1}, \ldots, m_n$ such that if $\nu'$ is the monomial valuation corresponding to weights $(\nu_1, \ldots, \nu_k, m_{k+1}, \ldots, m_n)$, then \begin{align*} \nu'(f) = m_{k+1} + \nu(f_{k+1}) = \cdots = m_l + \nu(f_l) < \nu'(x_ix_jf_{i,j})\ \text{for each}\ i\geq j > k. \end{align*} But then $\In_{\nu'}(f) = \sum_{i=k+1}^l x_i \In_{\nu}(f_i)$ and therefore $\partial (\In_{\nu'}(f))/\partial x_i = \In_\nu(\dfd{i}|_{\kk^I})$, $i = k+1, \ldots, n$. It follows that $\ensuremath{\mathfrak{j}}(\In_{\nu'}(f))$ have a common zero in $(\kk^*)^n$, and therefore $f$ is not Newton non-degenerate, as required.\\ It remains to prove the third assertion. Assume $f$ is inner Newton non-degenerate with respect to some $C$-polytope $\ensuremath{\mathcal{P}}$. For each $i' \not\in I$, let $M_{i'}$ be the set of {\em integral} elements in $\ensuremath{\mathcal{P}}$ of the form $\alpha + e_{i'}$ with $\alpha \in \rr^I$ (where $e_{i'}$ is the $i'$-th unit vector), and define \begin{align*} m_{i'} &:= \min\{\nu(\alpha'-e_{i'}): \alpha' \in M_{i'}\} \\ \Delta_{i'} &:= \{\alpha' \in M_{i'}: \nu(\alpha'-e_{i'}) = m_{i'}\} \end{align*} Let $\Delta_0 := \In_{\nu}(\ensuremath{\mathcal{P}}^I)$. Given $\eta \in \scrV_0$, we write $\ensuremath{\mathcal{P}}_\eta$ for the face $\In_\eta(\ensuremath{\mathcal{P}})$ of $\ensuremath{\mathcal{P}}$. Let $\nu'_0 \in \scrV_0$ be a weighted order such that \begin{align*} \nu'(x_i) &= \begin{cases} \nu(x_i) & \text{if}\ i \in I, \\ \gg 1 &\text{otherwise.} \end{cases} \end{align*} Then $\ensuremath{\mathcal{P}}_{\nu'_0} = \Delta_0$. In particular, if $I $ is a proper subset of $[n]$, then $\ensuremath{\mathcal{P}}_{\nu'_0}$ is not an inner face. Assume without loss of generality that $I = \{1, \ldots, k\}$. We now modify $\nu'_0$ as follows: decrease the ratio of weights of $x_{k+1}$ and $x_1$ while keeping the ratios of weights of $x_j$ and $x_1$ fixed for every other $x_j$ - until the corresponding face contains a point with positive $x_{k+1}$-coordinate. Then repeat the same process for $x_{k+2}, \ldots, x_n$. Let $\nu'_1$ be the resulting weighted order. Note that \begin{prooflist} \item \label{inner'-0} $\ensuremath{\mathcal{P}}_{\nu'_1}^I = \Delta_0$. \item \label{inner'-j} For each $j > k$, either $M_j \cap \ensuremath{\mathcal{P}}_{\nu'_1} = \emptyset$ or $M_j \cap \ensuremath{\mathcal{P}}_{\nu'_1} = \Delta_j$. \end{prooflist} Let $\tilde I := \{i: 1 \leq i \leq k,$ $\ensuremath{\mathcal{P}}_{\nu'_1}$ does not contain an element with positive $x_i$-coordinate$\}$. Then $\ensuremath{\mathcal{P}}_{\nu'_1}$ is not an inner face iff $\tilde I$ is non-empty. Assume without loss of generality that $\tilde I = \{l+1, \ldots, k\}$, $1 \leq l \leq k\}$. Define $\nu'_2$ by modifying $\nu'_1$ in the same way as the construction of $\nu'_1$, by decreasing relative weights of $x_i$'s for $i \in \tilde I$. Then $\ensuremath{\mathcal{P}}_{\nu'_2}$ is an inner face containing $\ensuremath{\mathcal{P}}_{\nu'_1}$. In particular, $\ensuremath{\mathcal{P}}_{\nu'_2}^{I\setminus \tilde I} \neq \emptyset$. By inner non-degeneracy of $f$ it follows that $\ensuremath{\mathfrak{j}}(\In_{\ensuremath{\mathcal{P}}_{\nu'_2}}(f))$ does not have any zero in $\Kstarii{I \setminus \tilde I}$. It is straightforward to check using \ref{inner'-0} and \ref{inner'-j} that the partial derivatives of $\In_{\ensuremath{\mathcal{P}}_{\nu'_2}}(f)$ that do not identically vanish on $\Kii{I \setminus \tilde I}$ belong to the set of $\In_{\nu}((\dfd{j})|_{\kk^I})$, $1\leq j \leq n$. Since for each $i \in \tilde I$, $x_i$ does not appear in the expression of $\In_{\nu}((\dfd{j})|_{\kk^I})$ for any $j$, $1 \leq j \leq n$, it follows that $\In_{\nu}((\dfd{j})|_{\kk^I})$, $1 \leq j \leq n$, do not have any common zero in $(\kk^*)^I$, as required. \end{proof} \begin{problem} \label{open-problem} \mbox{} \begin{enumerate} \item Find a more `natural' definition of Milnor non-degeneracy. In particular, is it possible to replace the artificial looking condition ``$\multzero{\Gamma_1}{\Gamma_n} = \multzero{\dGammad{1}}{\dGammad{n}}$'' by ``$\Gamma_j = \dGammad{j},\ 1 \leq j \leq n$''? \item In the statement of \cref{newton-wall-lemma}, can ``partially Milnor non-degenerate'' be replaced by ``Milnor non-degenerate''? \item Find an expression for $\multzero{\dGammad{1}}{\dGammad{n}}$ which is `more closely' related to $\Gamma$ than our formula \eqref{multiplicity-formula}. In particular, find a direct proof that if $\Gamma$ is convenient and no coordinate of any of its vertices is divisible by $\charac(\ensuremath{\mathbb{K}})$, then $\multzero{\dGammad{1}}{\dGammad{n}}$ equals an alternating sum of volumes as in Kushnirenko's original formula. \end{enumerate} \end{problem} \section{Extending BKK bound to complements of unions of coordinate subspaces} \label{affine-section} In this section we extend Bernstein's theorem to the complement of an arbitrary union of coordinate subspaces in $\ensuremath{\mathbb{K}}^n$. The main results are \cref{bkk-thm}, which gives a formula for the maximum number of isolated solutions of polynomials with given Newton polytopes, and \cref{generic-bkk-thm}, which gives an explicit characterization of those systems of polynomials which attain this maximum. \Cref{weak-bkkriterion} gives an equivalent characterization involving fewer conditions. \subsection{Extended BKK bound} Throughout this section $\mathscr{P}:=(\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_m)$ is an ordered collection of convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$. \begin{defn} We say that $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ are {\em $\mathscr{P}$-admissible} iff $\np(f_j) = \ensuremath{\mathcal{P}}_j$, $1 \leq j \leq m$. Let $I \subseteq [n]$. We say that \begin{itemize} \item $\mathscr{P}$ is {\em $(\kk^*)^I$-isolated} if $V(f_1, \ldots, f_n) \cap (\kk^*)^I$ is isolated for generic $\ensuremath{\mathcal{P}}$-admissible $f_1, \ldots, f_n$. \item $\mathscr{P}$ is {\em $(\kk^*)^I$-trivial} if $V(f_1, \ldots, f_n) \cap (\kk^*)^I$ is empty for generic $\ensuremath{\mathcal{P}}$-admissible $f_1, \ldots, f_n$. \end{itemize} \end{defn} The following result gives combinatorial characterizations of $I \subseteq [n]$ such that $\mathscr{P}$ is $(\kk^*)^I$-isolated or $(\kk^*)^I$-trivial. It also shows that if $\mathscr{P}$ is {\em not} $(\kk^*)^I$-isolated, then the zeroes of {\em every} collection of $\mathscr{P}$-admissible system of polynomials on $(\kk^*)^I$ is non-isolated. \begin{prop} \label{herrero} Let $I \subseteq [n]$. \begin{enumerate} \item The following are equivalent: \begin{enumerate} \item $\mathscr{P}$ is $(\kk^*)^I$-trivial. \item There exists $J \subseteq [m]$ such that $|\{j \in J: \ensuremath{\mathcal{P}}^I_j \neq \emptyset\}| > \dim(\sum_{j \in J} \ensuremath{\mathcal{P}}^I_j)$. \end{enumerate} \item The following are equivalent: \begin{enumerate} \item $\mathscr{P}$ is $(\kk^*)^I$-isolated. \item Either $\mathscr{P}$ is $(\kk^*)^I$-trivial or $|\{j: 1 \leq j \leq m,\ \ensuremath{\mathcal{P}}^I_j \neq \emptyset\}| = |I|$. \end{enumerate} \item If $\mathscr{P}$ is not $(\kk^*)^I$-isolated, then for all $\mathscr{P}$-admissible $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, \begin{enumerate} \item $V(f_1, \ldots, f_m) \cap (\kk^*)^I$ is non-empty, and \item every component of $V(f_1, \ldots, f_m) \cap \kk^I$ is positive dimensional. \end{enumerate} \end{enumerate} \end{prop} \begin{proof} Follows from \cite[Propositions 5 and 6]{herrero}. \end{proof} We now isolate a collection of coordinate subspaces of $\ensuremath{\mathbb{K}}^n$ such that all the isolated zeroes on $\ensuremath{\mathbb{K}}^n$ of a generic $\mathscr{P}$-admissible system of polynomials lie on the complement of their union. Define \begin{align} \scrN^{I}_{\mscrP} &:= \{j: \ensuremath{\mathcal{P}}^I_j \neq \emptyset\},\ I \subseteq [n] \label{NiP} \\ \scrE_\mscrP &:= \{I \subseteq [n]: \text{there is $\tilde I \supseteq I$ such that $|\NiiP{\tilde I}| < |\tilde I|$}\} \label{SP} \end{align} The following is an immediate corollary of \cref{herrero}. \begin{cor} \label{Pisolated-cor} Let $f_1, \ldots, f_m$ be $\ensuremath{\mathcal{P}}$-admissible polynomials. Then \begin{enumerate} \item If $S \in \scrE_\mscrP$, then each point of $V(f_1, \ldots, f_m) \cap \Kii{S}$ is non-isolated in $V(f_1, \ldots, f_m) \subseteq \ensuremath{\mathbb{K}}^n$. \item If $f_i$'s are generic, then each point of $V(f_1, \ldots, f_m)\setminus \bigcup_{S \in \scrE_\mscrP} \Kii{S}$ is isolated. \qed \end{enumerate} \end{cor} Now we introduce some notations which will be convenient in dealing with complements of coordinate subspaces of $\ensuremath{\mathbb{K}}^n$. \begin{defn} \label{VIS} Let $\mathscr{S}$ be a collection of subsets of $[n]$ and $I = \{i_1, \ldots, i_k\}$ be a non-empty subset of $[n]$. We denote by $\kk^I_{\mathscr{S}}$ the set $\kk^I \setminus \bigcup_{S \in \mathscr{S}} \Kstarii{S}$, by $\mathscr{S}^I$ the following collection of subsets of $I$: \begin{align} \mathscr{S}^I := \{S \in \mathscr{S}: S \subseteq I\} \label{SI} \end{align} and by $\overbar \mathscr{S}$ the closure of $\mathscr{S}$ under inclusion, i.e.\ \begin{align} \overbar \mathscr{S} := \{J \in [n]: J \subseteq S\ \text{for some}\ S \in \mathscr{S}\}. \label{barS} \end{align} Note that $$\kk^I \setminus \kk^I_{\mathscr{S}} = \bigcup_{S \in \mathscr{S}^I} \Kstarii{S},\ \text{and}\ \kk^I \setminus \kk^I_{\overbar \mathscr{S}} = \bigcup_{S \in {\overbar \mathscr{S}}^I} \Kii{S}.$$ We denote by $\scrV^I_\mscrS$ the set of monomial valuations $\nu$ on $\kk^I$ which are centered at $\kk^I \setminus \kk^I_{\mathscr{S}}$, i.e.\ for which there exists $S \in \mathscr{S}^I$ such that \begin{itemize} \item $\nu(x_s)= 0$ for all $s \in S$. \item $\nu(x_s)> 0$ for all $s \in I \setminus S$. \end{itemize} Finally, for $I = [n]$, we write $\ensuremath{\mathbb{K}}^n_\mathscr{S}$ for $\Kii{[n]}_\mathscr{S}$ and $\scrV_\mscrS$ for $\VSii{[n]}$. \end{defn} \begin{example} \label{KnS-remark} \mbox{} \begin{itemize} \item $\kk^I_{\emptyset} = \kk^I$ and $\VSiii{I}{\emptyset} = \emptyset$; in particular $\ensuremath{\mathbb{K}}^n_\emptyset= \ensuremath{\mathbb{K}}^n$. \item $\kk^I_{\{\emptyset\}} = \kk^I \setminus \{0\}$ and $\VSiii{I}{\{\emptyset\}} = \scrV^I_0$. \item $\kk^I_{2^I\setminus\{I\}} = (\kk^*)^I$. \end{itemize} \end{example} \begin{defn} \label{generic-intersection-mult-defn} Let $f_1, \ldots f_n \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. For each $z \in \ensuremath{\mathbb{K}}^n$, we write $\mult{f_1}{f_n}_z$ for the intersection multiplicity of $f_i$'s at $z$, and for $W \subseteq \ensuremath{\mathbb{K}}^n$, we define \begin{align*} \mult{f_1}{f_n}_W &:= \sum_{z \in W}\mult{f_1}{f_n}_z \end{align*} In other words, $\mult{f_1}{f_n}_W$ is the number (counted with multiplicity) of the zeroes of $f_1, \ldots, f_n$ on $W$. We also write $\multiso{f_1}{f_n}_W$ for the number (counted with multiplicity) of the zeroes of $f_1, \ldots, f_n$ on $W$ which are isolated in the (possibly larger) set of zeroes of $f_1, \ldots, f_n$ on $\ensuremath{\mathbb{K}}^n$. Note that $\multiso{f_1}{f_n}_W$ is always finite, whereas $\mult{f_1}{f_n}_W$ is either infinite or equal to $\multiso{f_1}{f_n}_W$. Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$. Define $\mult{\scrP_1}{\scrP_n}_W$ (resp.\ $\multiso{\scrP_1}{\scrP_n}_W$) to be $\mult{f_1}{f_n}_W$ (resp.\ $\multiso{f_1}{f_n}_W$) for generic $\mathscr{P}$-admissible $f_1, \ldots, f_n$. \end{defn} The following corollary is simply a reformulation of \cref{Pisolated-cor} in terms of the notations introduced in \cref{VIS,generic-intersection-mult-defn}. \begin{cor} \label{Pisolated-cor2} Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$. Let $\scrE_\mscrP$ be as in \eqref{SP}. Then \begin{enumerate} \item For each collection $\mathscr{S}$ of subsets of $[n]$, $\multiso{\scrP_1}{\scrP_n}_{\kk^n_{\overbar \mscrS}} = \mult{\scrP_1}{\scrP_n}_{\ensuremath{\mathbb{K}}^n_{\overbar \mathscr{S} \cup \scrE_\mscrP}} < \infty$. \item In particular, $\multiso{\scrP_1}{\scrP_n}_{\ensuremath{\mathbb{K}}^n} = \mult{\scrP_1}{\scrP_n}_{\ensuremath{\mathbb{K}}^n_{\scrE_\mscrP}} < \infty$. \qed \end{enumerate} \end{cor} Given a collection $\mathscr{S}$ of subsets of $[n]$, \cref{bkk-thm} below gives the formula for the number of isolated zeroes of a generic system of polynomials on $\kk^n_\mscrS$. The formula uses \cref{multstarinfty-defn}; it is instructive to compare formula \eqref{bkk-formula} with formula \eqref{multiplicity-formula} from \cref{multiplicity-thm}. \begin{defn} \label{multstarinfty-defn} Given a collection $\mathscr{S}$ of subsets of $[n]$ and a collection $\mathscr{P} := (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$, define \begin{align*} \multstarinftyy{\scrP_1}{\scrP_n} &:= \sum_{\omega \in \Omega_\infty} \omega(\ensuremath{\mathcal{P}}_1) \mv(\ld_\omega(\ensuremath{\mathcal{P}}_2), \ldots, \ld_\omega(\ensuremath{\mathcal{P}}_n))\\ \multstarSS{\scrP_1}{\scrP_n} &:= \sum_{\nu \in \scrV^I_\mscrS} \nu(\ensuremath{\mathcal{P}}_1) \mv(\In_\nu(\ensuremath{\mathcal{P}}_2), \ldots, \In_\nu(\ensuremath{\mathcal{P}}_n)) \\ \multstarinftyS{\scrP_1}{\scrP_n} &:= \multstarinftyy{\scrP_1}{\scrP_n} - \multstarSS{\scrP_1}{\scrP_n} \end{align*} \end{defn} \begin{thm}[Extended BKK bound \label{bkk-thm} With the notation from \cref{multstarinfty-defn}, define \begin{align} \begin{split} \scrI_{\mscrS,\mscrP,1} &:= \{I \subseteq [n]: I \not\in \overbar \mathscr{S} \cup \scrE_\mscrP \cup \{\emptyset\},\ |\scrN^{I}_{\mscrP}| = |I|,\ \text{$\mathscr{P}$ is $(\kk^*)^I$-non-trivial},\ 1 \in \scrN^{I}_{\mscrP} \} \end{split} \label{I1-list-P} \end{align} where $\scrN^{I}_{\mscrP}$ is as in \eqref{NiP}. Let $\Gamma_j$ be the Newton diagram of generic polynomials supported at $\ensuremath{\mathcal{P}}_j$, $1 \leq j \leq n$. Then \begin{align} \multiso{\scrP_1}{\scrP_n}_{\kk^n_{\overbar \mscrS}} &= \sum_{I \in \scrI_{\mscrS,\mscrP,1}} \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \times \multstarinftySS{\ensuremath{\mathcal{P}}^{I}_1, \ensuremath{\mathcal{P}}^{I}_{j'_1}}{\ensuremath{\mathcal{P}}^{I}_{j'_{k-1}}}{(\overbar \mathscr{S} \cup \scrE_\mscrP)^I} \label{bkk-formula} \end{align} where for each $I \in \scrI_{\mscrS,\mscrP,1}$, \begin{itemize} \item $I' := [n]\setminus I$, $k := |I|$, \item $j_1, \ldots, j_{n-k}$ are elements of $[n]\setminus \scrN^{I}_{\mscrP}$, \item $j'_1, \ldots, j'_{k-1}$ are elements of $\scrN^{I}_{\mscrP} \setminus \{1\}$, \item $(\overbar\mathscr{S} \cup \scrE_\mscrP)^I = \{S \in \overbar \mathscr{S} \cup \scrE_\mscrP: S \subseteq I\} $ (as in \eqref{SI}). \end{itemize} \end{thm} \begin{proof} See \cref{bkk-proof-section}. \end{proof} \begin{example} \label{ex0-bkk} Consider $f_1,f_2, f_3$ from \cref{ex0}. We compute $[f_1,f_2,f_3]_{\ensuremath{\mathbb{K}}^3}$ by applying \cref{bkk-thm} with $\mathscr{S} = \emptyset$ (recall from \cref{KnS-remark} that $\ensuremath{\mathbb{K}}^3 = \ensuremath{\mathbb{K}}^3_\emptyset$). Let $\ensuremath{\mathcal{P}}_i$ be the Newton polytope of $f_i$ for each $i$, and set $\mathscr{P} := (\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3)$. It is straightforward to see that $\scrE_\mscrP = \emptyset$, $\scrI_{\mscrS,\mscrP,1} = \{\{1,2,3\},\{3\}\}$. Moreover, $\ensuremath{\mathcal{P}}_2 + \ensuremath{\mathcal{P}}_3$ has precisely two facets with outer normals in $\ensuremath{\mathbb{R}}^3\setminus (\ensuremath{\mathbb{R}}_{\leq 0})^3$, and the outer normals are $\omega_1 := (1,1,1)$ and $\omega_2 := (2,2,1)$. It follows that \begin{align*} [f_1,f_2,f_3]_{\ensuremath{\mathbb{K}}^3} &= [\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3]^*_\infty + [\pi_{\{1,2\}}(\Gamma_2), \pi_{\{1,2\}}(\Gamma_3)]_0 ~ [\ensuremath{\mathcal{P}}^{\{3\}}_1]^*_\infty \\ &= \omega_1(\ensuremath{\mathcal{P}}_1)\mv(\ld_{\omega_1}(\ensuremath{\mathcal{P}}_2), \ld_{\omega_1}(\ensuremath{\mathcal{P}}_3)) + \omega_2(\ensuremath{\mathcal{P}}_1)\mv(\ld_{\omega_2}(\ensuremath{\mathcal{P}}_2), \ld_{\omega_2}(\ensuremath{\mathcal{P}}_3)) + 1 \cdot \deg_z(f_1|_{x=y=0}) \\ &= 1 \cdot 2 + 2 \cdot 3 + 1 = 9 \end{align*} \end{example} \subsection{Non-degeneracy condition for the extended BKK bound} In general it is possible that a given polynomial system has an isolated zero on a coordinate subspace $K$ of $\ensuremath{\mathbb{K}}^n$, even though generic systems with the same Newton polytopes (as those of the given system) have no zeroes on $K$; for example, the system $(x + y - 1, 2x - y - 2)$ (over a field of characteristic not equal to two) has an isolated zero on the coordinate subspace $y = 0$. \Cref{absolute-defn} introduces a class of coordinate subspaces for which this is not possible. \begin{defn} \label{absolute-defn} Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$. For $I \subseteq [n]$, define $\scrN^{I}_{\mscrP}$ as in \eqref{NiP}. We say that $\mathscr{P}$ is {\em $(\kk^*)^I$-exotrivial} if \begin{defnlist} \item \label{absolute--less} there is $\tilde I \supseteq I$ such that $|\NiiP{\tilde I}| < |\tilde I|$, i.e.\ $I \in \scrE_\mscrP$, or \item \label{absolute-equal} $\mathscr{P}$ is $(\kk^*)^I$-trivial, and there is $\tilde I \supseteq I$ such that \begin{defnlist} \item $\mathscr{P}$ is $\Kstarii{\tilde I}$-trivial, \item $|\NiiP{\tilde I}| = |\tilde I|$, \item $|\NiiP{I^*}| > |I^*|$ for each $I^*$ such that $I \subseteq I^* \subsetneqq \tilde I$. \end{defnlist} \end{defnlist} \end{defn} \begin{prop} \label{absolute-prop} Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$ and $I \subseteq [n]$. If $\mathscr{P}$ is $(\kk^*)^I$-exotrivial, then \begin{align*} \multiso{f_1}{f_n}_{(\kk^*)^I} = 0 \end{align*} for every $\mathscr{P}$-admissible system of polynomials $f_1, \ldots, f_n$. \end{prop} \begin{proof} It is proven in \cref{absolute-cor}. \end{proof} \begin{rem} The prefix `exo' in {\em exotrivial} is to convey that $I$ satisfies the conclusion of \cref{absolute-prop} due to an `external' reason, namely the presence of some $\tilde I \supseteq I$ as in \cref{absolute-defn}. The other reason for $I$ satisfying the conclusion of \cref{absolute-prop} is `internal': the presence of some $\tilde I \subseteq I$ satisfying certain properties; we omit the discussion here since it is not relevant for the purpose of this article. \end{rem} \Cref{generic-bkk-defn} introduces $\mathscr{P}$-non-degeneracy, and \cref{generic-bkk-thm}, which is the main result of this section, shows that this is the correct non-degeneracy criterion for the extended BKK bound. \begin{defn} \label{generic-bkk-defn} Let $f_1, \ldots, f_m$, $m \geq n$, be polynomials in $(x_1, \ldots, x_n)$ and $\mathscr{S}$ be a collection of subsets of $[n]$. \begin{itemize} \item We say that \begin{itemize} \item $f_i$'s are {\em BKK non-degenerate at infinity} if $\ld_\omega(f_1), \ldots, \ld_\omega(f_m)$ have no common zero in $(\kk^*)^n$ for all $\omega \in \Omega_\infty$. \item $f_i$'s are {\em BKK non-degenerate at $\mathscr{S}$} if $\In_\nu(f_1), \ldots, \In_\nu(f_m)$ have no common zero in $(\kk^*)^n$ for all $\nu \in \scrV_\mscrS$. \end{itemize} \item Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$. Let $\mathscr{S}$ be a collection of subsets of $[n]$. Define \begin{align} \scrT_{\mscrS,\mscrP} &:= \{I \subseteq [n]: I \not \in \overbar \mathscr{S} \cup \scrE_\mscrP,\ \mathscr{P}\ \text{is $(\kk^*)^I$-exotrivial}\} \label{TSP}\\ \scrT'_{\mscrS,\mscrP} &:= \{I \subseteq [n]: I \not\in \overbar \mathscr{S} \cup \scrE_\mscrP,\ \mathscr{P}\ \text{is {\em not} $(\kk^*)^I$-exotrivial}\} \label{TSPrime} \end{align} \item We say that polynomials $f_1, \ldots, f_n$ are {\em $\mathscr{P}$-non-degenerate on $\kk^n_{\overbar \mscrS}$} iff \begin{defnlist} \item \label{Padmissible} They are $\mathscr{P}$-admissible. \item \label{TSPrime-property} For all $I \in \scrT'_{\mscrS,\mscrP}$, $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK non-degenerate both at infinity and at $(\overbar \mathscr{S} \cup \scrT_{\mscrS,\mscrP})^I = \{S \in \overbar \mathscr{S} \cup \scrT_{\mscrS,\mscrP}: S \subseteq I\}$. \item \label{NTSP-property} For all $I \in \scrT_{\mscrS,\mscrP}$, $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK non-degenerate at ${\scrT'}^I_{\mscrS,\mscrP} = \{S \in \scrT'_{\mscrS,\mscrP}: S \subseteq I\}$. \end{defnlist} \end{itemize} \end{defn} \begin{thm}[Non-degeneracy condition for the extended BKK bound] \label{generic-bkk-thm} Let $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$ and $f_1, \ldots, f_n$ be $\mathscr{P}$-admissible polynomials in $(x_1, \ldots, x_n)$. Let $\scrT_{\mscrS,\mscrP}$ be as in \eqref{TSP}. Then \begin{enumerate} \item \label{generic-assertion-0} $\multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} = \multiso{f_1}{f_n}_{\KnSS{\overbar \mathscr{S} \cup \scrT_{\mscrS,\mscrP}}} \leq \mult{\scrP_1}{\scrP_n}_{\KnSS{\overbar \mathscr{S} \cup \scrT_{\mscrS,\mscrP}}} < \infty$. \item \label{generic-assertion} The following are equivalent: \begin{enumerate} \item $f_i$'s are $\mathscr{P}$-non-degenerate on $\kk^n_{\overbar \mscrS}$ \item $\multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} = \mult{\scrP_1}{\scrP_n}_{\KnSS{\overbar\mathscr{S} \cup \scrT_{\mscrS,\mscrP}}} $ \end{enumerate} \item \label{isolated-assertion} If $f_i$'s are $\mathscr{P}$-non-degenerate on $\kk^n_{\overbar \mscrS}$, then \begin{align*} \multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} = \mult{f_1}{f_n}_{\KnSS{\overbar\mathscr{S} \cup \scrT_{\mscrS,\mscrP}}} \end{align*} In particular, in this case all zeroes of $f_1, \ldots, f_n$ on $\KnSS{\overbar\mathscr{S} \cup \scrT_{\mscrS,\mscrP}}$ are isolated. \end{enumerate} \end{thm} \begin{proof} See \cref{generic-bkk-proof}. \end{proof} \begin{example}[\Cref{b-example} continued]\label{b-example-contd} Let $\mathscr{P} := (\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3)$, where $\ensuremath{\mathcal{P}}_i$ are the Newton polytopes of $f_i$ from \cref{b-example}. We computed in \cref{b-example} that $[\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3]_{\ensuremath{\mathbb{K}}^3} = 2$. Now we examine when $f_1,f_2,f_3$ are $\mathscr{P}$-non-degenerate on $\ensuremath{\mathbb{K}}^3$, so take $\mathscr{S} = \emptyset$. It is straightforward to check that $\scrE_\mscrP = \emptyset$, $\scrT_{\mscrS,\mscrP}=\{ \{1,2\}\}$ and $\scrT'_{\mscrS,\mscrP}$ is the collection of all non-empty subsets of $\{1,2,3\}$ excluding $\{1,2\}$. In cases \ref{b-example-generic-0} and \ref{b-example-generic-1} of \cref{b-example}, it is straightforward to check that $f_1, f_2, f_3$ are $\mathscr{P}$-non-degenerate, so that \cref{generic-bkk-thm} implies that the extended BKK bound is attained in these cases, as we saw in \cref{b-example}. On the other hand, in case \ref{b-example-non-generic} the discussion from \cref{b-example} shows that for $I = \{1, 2, 3\}$ the BKK non-degeneracy at infinity fails with $\omega = (1,-1,-2)$; this violates condition \ref{TSPrime-property} of $\mathscr{P}$-non-degeneracy and therefore \cref{generic-bkk-thm} implies that the number of isolated zeroes of $f_1, f_2, f_3$ is less than $[\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3]_{\ensuremath{\mathbb{K}}^3} = 2$, as observed in \cref{b-example}. \end{example} \begin{example}[\Cref{c-example} continued]\label{c-example-contd} Let $\mathscr{P} := (\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3, \ensuremath{\mathcal{P}}_4)$, where $\ensuremath{\mathcal{P}}_i$ are the Newton polytopes of $f_i$ from \cref{c-example}, and let $\mathscr{S} := \emptyset$. It is straightforward to check that $\scrE_\mscrP = \emptyset$, $\scrT_{\mscrS,\mscrP}=\{ \{1,2,3,4\}\}$ and $\scrT'_{\mscrS,\mscrP}$ is the collection of all proper non-empty subsets of $\{1,2,3,4\}$. It follows that in case \ref{c-example-non-generic} of \cref{c-example}, condition \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy fails with $I = \{1,2,3,4\}$ and the monomial valuation $\nu$ corresponding to weights $(0,0,1,2)$ for $(x_1, \ldots, x_4)$. \Cref{generic-bkk-thm} therefore implies that in this case the number of isolated zeroes of $f_1, f_2, f_3, f_4$ is less than $[\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3, \ensuremath{\mathcal{P}}_4]_{\ensuremath{\mathbb{K}}^4}$, as we found out in \cref{c-example}. \end{example} \Cref{generic-bkk-thm} implies a dichotomy among coordinate subspaces of $\ensuremath{\mathbb{K}}^n$: \begin{cor} \label{non-isolated-cor} Let $\mathscr{P} = (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$ and $\mathscr{S}$ be a collection of subsets of $[n]$. Assume $\mult{\scrP_1}{\scrP_n}_{\kk^n_{\overbar \mscrS}} > 0$. Then for every $\mathscr{P}$-admissible system $f = (f_1, \ldots, f_n)$ of polynomials and every subset $I$ of $[n]$ such that $I \not \in \overbar{\mathscr{S}}$ (so that $(\kk^*)^I \subseteq \kk^n_{\overbar \mscrS}$), \begin{enumerate} \item \label{non-isolated-not-<} If $I \in \scrT_{\mscrS,\mscrP}$, then the existence of non-isolated roots of $f$ on $(\kk^*)^I$ does not necessarily imply that $\multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} < \multiso{\scrP_1}{\scrP_n}_{\kk^n_{\overbar \mscrS}}$. \item \label{non-isolated-<} If $I \in \scrT'_{\mscrS,\mscrP}$, then the existence of non-isolated roots of $f$ on $(\kk^*)^I$ implies that $\multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} < \multiso{\scrP_1}{\scrP_n}_{\kk^n_{\overbar \mscrS}}$. \end{enumerate} \end{cor} \begin{proof} This is proven in \cref{non-isolated-section}. \end{proof} The following result describes a criterion which is equivalent to $\mathscr{P}$-non-degeneracy, but requires checking fewer conditions. \begin{lemma} \label{weak-bkkriterion} Let $\mathscr{P} := (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ be a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$ and $\mathscr{S}$ be a collection of subsets of $[n]$. Define \begin{align} \scrT^*_{\mscrS,\mscrP} &:= \{I\in \scrT_{\mscrS,\mscrP}:\ |\scrN^{I}_{\mscrP}| = |I|\}, \label{TSPstar}\\ \scrN_{\mscrS,\mscrP} &:= \{I \in \scrT'_{\mscrS,\mscrP}: \mathscr{P}\ \text{is $(\kk^*)^I$-non-trivial}\}, \label{NSP} \end{align} where $\scrN^{I}_{\mscrP}, \scrT_{\mscrS,\mscrP}, \scrT'_{\mscrS,\mscrP}$ are respectively as in \eqref{NiP}, \eqref{TSP} and \eqref{TSPrime}. Let $f_1, \ldots, f_n \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. Then the following are equivalent: \begin{enumerate} \item $f_i$'s are $\mathscr{P}$-non-degenerate on $\kk^n_{\overbar \mscrS}$. \item \begin{enumerate}[label=(\roman{enumii})] \item property \ref{Padmissible} of $\mathscr{P}$-non-degeneracy holds, \item property \ref{TSPrime-property} of $\mathscr{P}$-non-degeneracy holds with $\scrT'_{\mscrS,\mscrP}$ replaced by $\scrN_{\mscrS,\mscrP}$, \item property \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy holds with $\scrT_{\mscrS,\mscrP}$ replaced by $\scrT^*_{\mscrS,\mscrP}$. \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} This is \cref{weak-bkk-lemma}. \end{proof} \begin{remexample}[Warning] In property \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy $\scrT'_{\mscrS,\mscrP}$ can {\em not} be replaced by $\scrN_{\mscrS,\mscrP}$. Indeed, let \begin{align*} f_1 &= 1 + x+y + z \\ f_2 &= 1 + x + 2y + 3z \\ f_3 &= y(ax + by + cz)+ z(a'x+b'y+c'z) \\ f_4 &= w(1+x) \end{align*} where $a,b,c,a',b',c'$ are generic elements in $\ensuremath{\mathbb{K}}^*$. Consider the ordering $(w,x,y,z)$ of the variables. If $\mathscr{P}$ is the collection of newton polytopes of $f_1, \ldots, f_4$, then $\scrT_{\mscrS,\mscrP}$ (resp.\ $\scrT'_{\mscrS,\mscrP}$) is the collection of all subsets of $\{1,2,3,4\}$ containing (resp.\ not containing) $1$. It is straightforward to check that \ref{NTSP-property} of of $\mathscr{P}$-non-degeneracy on $\ensuremath{\mathbb{K}}^4$ fails with $I = \{1, 2,3,4\}$ and $\nu = (1,0,1,1)$ and therefore the the number of isolated solutions of $f_1, \ldots, f_4$ on $\ensuremath{\mathbb{K}}^4$ is less than $[\ensuremath{\mathcal{P}}_1, \ensuremath{\mathcal{P}}_2, \ensuremath{\mathcal{P}}_3]^{iso}_{\ensuremath{\mathbb{K}}^4}$. However, if $\scrT'_{\mscrS,\mscrP}$ is replaced by $\scrN_{\mscrS,\mscrP}$ in \ref{NTSP-property}, then this system would be $\mathscr{P}$-non-degenerate on $\ensuremath{\mathbb{K}}^4$. \end{remexample} \section{Proof of the non-degeneracy condition for intersection multiplicity} \label{generic-proof-section} In this section we prove \cref{generic-thm} following the outline in \cref{idea-section}. We start with clarifying what we mean by a `branch' of a curve and the corresponding `initial coefficients' of the (restrictions of) coordinate functions. The initial coefficients of a branch corresponds in an obvious way to a common zero of the corresponding initial forms of polynomials defining the curve (\cref{branch-lemma}), and this is in a sense the basic observation behind Bernstein's non-degeneracy conditions. In \cref{blow-up-intersection,blow-up-coordinates} we compile some (straightforward) facts about {\em weighted blow-ups} at the origin. The main technical result of \cref{pre-multiplicity-section} is \cref{reduction-lemma} which is our key to circumvent the problem outlined in \cref{pathologies}. In \cref{generic-proof} we use these results to prove \cref{generic-thm}. \subsection{Preparatory results} \label{pre-generic-section} \begin{defn} \label{branchnition} Let $C$ be a (possibly reducible) curve on a variety $X$. Let $\pi_{C'}: C' \to C$ be a fixed desingularization of $C$ and $\bar C'$ be a fixed non-singular compactification of $C'$. A {\em branch} of $C$ is the germ of a point $z \in C'$, i.e.\ it is an equivalence class of the equivalence relation $\sim$ on pairs $B := (Z,z)$ such that \begin{itemize} \item $Z$ is an open subset of an irreducible component of $\bar C'$ and $z \in Z$. \item $(Z, z) \sim (Z',z')$ iff $z = z'$ and $Z \cap Z'$ is open in both $Z$ and $Z'$. \end{itemize} In the case that $z \in \pi_{C'}^{-1}(O)$ for some $O \in C$, we say that $B$ is a {\em branch at $O$}. If $z \in \bar C' \setminus C$, we say that $B$ is a {\em branch at infinity} (with respect to $X$). \end{defn} \begin{defn} \label{branch-remark} Assume $X \subseteq \ensuremath{\mathbb{K}}^n$. Let $B := (Z,z)$ be a branch of a curve $C \subseteq X$. Identify $Z^* := Z \setminus z$ with its image on $C$ and let $I_B := \{i: x_i|_{Z^*} \not\equiv 0\}$. Note that $\Kii{I_B}$ is the smallest coordinate subspace of $\ensuremath{\mathbb{K}}^n$ which contains $Z^*$. For $f \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, we write $f|_B$ for $f|_{Z^*}$ and $\nu_B(f)$ for $\ord_z(f|_B)$. Let $I := I_B$ and $d := \gcd(\nu_B(x_i): i \in I_B)$. We denote by $\nu^I_B$ the monomial valuation in $\scrV^I$ corresponding to weights $\nu_B(x_i)/d$ for $x_i$ for each $i \in I$. Fix an arbitrary element $\phi_B$ in $\ensuremath{\mathbb{K}}(X)$ such that $\phi_B|_B$ is well defined and $\ord_z(\phi|_B) = 1$. Define \begin{align*} \In_B(x_j) &:= \begin{cases} 0 &\text{if}\ j \not\in I \\ \left. \frac{x_j}{(\phi_B)^{\nu_B(x_i)}} \right|_z &\text{if}\ j \in I. \end{cases}\\ \In(B) &:= (\In_B(x_1), \ldots, \In_B(x_n)) \in (\kk^*)^I \end{align*} Note that $\In(B)$ depends on the choice of $\phi_B$. In all cases below, whenever a branch $B$ is considered, the corresponding $\phi_B$ is assumed to be fixed in the beginning; in other words, branches are to be understood as {\em triplets} $(Z,z, \phi)$. \end{defn} \begin{rem} In general $\nu^I_B$ is {\em not} the restriction of $\nu_B$ to the coordinate ring $A^I$ of $\kk^I$. Indeed, $\nu^I_B$ is a monomial valuation, but unless $B$ is a coordinate axis, $\nu_B|_{A^I}$ is not even a discrete valuation, e.g.\ if $f$ is a non-zero polynomial in $A^I$ that vanishes on $B$, then $\nu_B(f) = \infty$. \end{rem} If $\charac(\ensuremath{\mathbb{K}}) = 0$, then the weights of $\nu^I_B$ are proportional to the exponents of initial terms of Puiseux series expansions of $x_j$ determined by $B$ and $\In(B)$ is (up to multiplication by a non-zero constant) the vector of coefficients of these initial terms. The following lemma states the fact, which is straightforward to verify, that $\In(B)$ is a common zero of the initial forms (corresponding to $\nu^I_B$) of all polynomials which vanish on $B$. \begin{lemma} \label{branch-lemma} Let $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ and $B$ be a branch of a curve contained in $V(f_1, \ldots, f_m)$. Let $I :=I_B$ and $\nu := \nu^I_B$. Then $\In(B) \in V(\In_\nu(f_1|_{\kk^I}), \ldots, \In_\nu(f_m|_{\kk^I})) \cap (\kk^*)^I$. \qed \end{lemma} Now we interpret common zeroes of initial forms of $f_j$'s in terms of {\em weighted blow-ups} of $\ensuremath{\mathbb{K}}^n$ at the origin. \begin{defn} \label{blow-up-defn} Let $\nu \in \scrV_0$. The {\em $\nu$-weighted blow up $X_\nu$ at the origin} with respect to $(x_1, \ldots, x_n)$-coordinates is the blow up of $\ensuremath{\mathbb{K}}^n$ at the ideal $\ensuremath{\mathfrak{q}}_N$ generated by all monomials $x^\alpha$ in $(x_1, \ldots, x_n)$ such that $\nu (x^\alpha) = Np$, where $p := \lcm(\nu_1, \ldots, \nu_n)$ and $N$ is a sufficiently large integer. The exceptional divisor $E_\nu$ has a natural structure of the weighted projective space $\ensuremath{\mathbb{P}}(\nu_1, \ldots, \nu_n)$ with (weighted) homogeneous coordinates $[x_1: \cdots : x_n]$. We call $E^*_\nu := E_\nu \setminus V(x_1\cdots x_n)$ the set of {\em interior points} of $E_\nu$. Note that $E^*_\nu \cong (\ensuremath{\mathbb{K}}^*)^{n-1}$ and $X_\nu$ is non-singular at every point of $E^*_\nu$. \end{defn} The next two results follow from standard facts about toric varieties: \begin{lemma} \label{blow-up-intersection} Let $\nu, X_\nu, E^*_\nu$ be as in \cref{blow-up-defn}. Let $z \in (\kk^*)^n$ and let $[z] \in E^*_\nu$ be the point whose homogeneous coordinates are the same as the coordinates of $z$. Let $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$. The following are equivalent: \begin{enumerate} \item $z$ is a common zero of $\In_\nu(f_j)$, $1 \leq j \leq m$. \item $[z]$ is in the intersection on $X_\nu$ of the strict transforms of $\{f_j = 0\}$. \qed \end{enumerate} \end{lemma} \begin{lemma} \label{blow-up-coordinates} Let $I \subseteq [n]$, $\nu \in \scrV^I_0$, $\nu' \in \scrV_0$ be compatible with $\nu$. Let $X_\nu$ (resp.\ $X'_{\nu'}$) be the $\nu$-weighted (resp.\ $\nu'$-weighted) blow up of $\kk^I$ (resp.\ $\ensuremath{\mathbb{K}}^n$). \begin{enumerate} \item $X_\nu$ can be identified with the strict transform of $\kk^I$ in $X'_{\nu'}$. \item If $f \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ is such that $\In_{\nu'}(f) = \In_\nu(f|_{\kk^I})$, then under the above identification the strict transform of $\{f|_{\kk^I} = 0\}$ in $X_\nu$ corresponds to the intersection of the strict transforms of $\kk^I$ and $\{f = 0\}$ on $X'_{\nu'}$. \item \label{coordinate-assertion} Assume $\gcd(\nu'(x_i): i \in I) = 1$. Let $k := |I|$. Then under the above identification $E^*_\nu$ is contained in an non-singular open subset $W$ of $X'_{\nu'}$ isomorphic to $ (\kk^*)^{k-1} \times \ensuremath{\mathbb{K}}^{n-k+1}$ with respect to coordinates $(z_1, \ldots, z_n)$ such that \begin{enumerate} \item $z_1, \ldots, z_k$ are monomials in $(x_i: i \in I)$. \item $\nu'(z_1) = \nu(z_1) = 1$, $\nu'(z_i) = 0$, $2 \leq i \leq n$. \item for all $i' \not\in I$, $z_{i'} = x_{i'}/z_1^{m_{i'}}$ for some positive integer $m_{i'}$. \item $E_{\nu'} \cap W = V(z_1)$. \item $X_\nu \cap W = V(z_{k+1}, \ldots, z_n)$. \item $E_\nu \cap W = E^*_\nu \cap W = V(z_1, z_{k+1}, \ldots, z_n)$. \qed \end{enumerate} \end{enumerate} \end{lemma} Our next result (\cref{reduction-lemma}) plays a crucial role in this article; it is the key to \cref{weak-lemma,weak-bkkriterion} which state that when testing for our non-degeneracy conditions, we may omit the coordinate subspaces which are problematic for the reason described in \cref{pathologies}. \begin{lemma} \label{reduction-lemma} Let $I$ be a non-empty subset of $[n]$ and $f_1, \ldots, f_m \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n][x_i^{-1}: i \in I]$. Let $\scrT^{I}:= \{j \in [m]: f_j|_{(\kk^*)^I} \equiv 0\}$. Assume \begin{enumerate} \item \label{less-than-assumtion} $|\scrT^{I}| < n - |I|$. \item \label{initial-assumption} there exists $\nu \in \scrV^I$ such that $\In_\nu(f_1|_{(\kk^*)^I}), \ldots, \In_\nu(f_n|_{(\kk^*)^I})$ have a common zero $u \in (\kk^*)^n$. \end{enumerate} Then there exists $\tilde I \supsetneqq I$ and $\tilde \nu \in \Vii{\tilde I}$ such that \begin{enumerate} \setcounter{enumi}{2} \item \label{compatible-reduction} $\tilde \nu$ is compatible with $\nu$. \item \label{positive-reduction} $\tilde \nu(x_j) > 0$ for each $j \in \tilde I\setminus I$. \item \label{zero-reduction} $\In_{\tilde \nu}(f_1|_{\Kstarii{\tilde I}}), \ldots, \In_{\tilde \nu}(f_n|_{\Kstarii{\tilde I}})$ have a common zero in $\tilde u \in (\kk^*)^n$ such that $\pi_I(\tilde u) = \pi_I(u)$. \end{enumerate} \end{lemma} \begin{proof} We may assume without any loss of generality that $I = \{1, \ldots, k\}$ and $\scrT^{I} = \{1, \ldots, l\}$, $1 \leq k < n$, $l < n-k$. Let $a := \pi_I(u)$. Then $a =(a_1, \ldots, a_k, 0, \ldots, 0)$ for some $a_1, \ldots, a_k \in \ensuremath{\mathbb{K}}^*$. At first consider the case that $\nu$ is the trivial valuation. Assumption \eqref{initial-assumption} then says that $a$ is a common zero of $f_1, \ldots, f_m$ in $(\kk^*)^I$. Let $(z_1, \ldots, z_n)$ be the translation of $(x_1, \ldots, x_n)$ by $a$, i.e.\ \begin{align*} z_j &:= \begin{cases} x_j - a_j &\text{if}\ 1 \leq j \leq k\\ x_j &\text{if}\ k < j \leq n. \end{cases} \end{align*} Let $X_{\nu^*}$ be the weighted blow up of $\ensuremath{\mathbb{K}}^n$ at $a$ in $(z_1, \ldots, z_n)$ coordinates corresponding to a monomial valuation $\nu^*$ such that \begin{itemize} \item $\nu^*(z_j) > 0$, $1 \leq j \leq k-1$, \item $\nu^*(z_k) = 1$, \item $\nu^*(z_j) \gg 1$ for $k < j \leq n$. \end{itemize} Let $E_{\nu^*}$ be the exceptional divisor of the blow up $X_{\nu^*} \to \ensuremath{\mathbb{K}}^n$. Pick a point $a'$ on the intersection of $E_{\nu^*}$ with the strict transform $Y'$ of $(\kk^*)^I$. Since $l < n-k$, there is an irreducible component $V$ of $V(f_1, \ldots, f_l)$ properly containing $(\kk^*)^I$. Then the strict transform $V'$ of $V$ properly contains $Y'$. Choose a curve $C' \subseteq V'$ such that $a' \in C' \not\subseteq Y' \cup E_{\nu^*}$. Let $B = (Z',z')$ be a branch of $C'$ such that $z'$ is in the pre-image (in the desingularization of $C'$) of $a'$. Let $\tilde I := \{i : 1 \leq i \leq n, x_i|_B \not\equiv 0\}$, $\tilde \nu := \nuBii{\tilde I}$ and $\In(B) := (\In_B(x_1), \ldots, \In_B(x_n)) \in \Kstarii{\tilde I}$ (\cref{branch-remark}). Then \begin{itemize} \item For $1 \leq j \leq k$, $\nu_B(z_j) > 0$, so that $\nu_B(x_j) = 0$. In particular $\In_B(x_j) = a_j$, $1 \leq j \leq k$, and $I \subseteq \tilde I$. \item Since $C' \not\subseteq Y'$, it follows that $I \subsetneqq \tilde I$. \item $\nu_B(x_{i'}) \gg 1$ for each $i' \in \tilde I\setminus I$. It follows that $\In_{\tilde \nu}(f_j) = \In_\nu(f_j)$ for each $j$, $l+1 \leq j \leq m$. (\cref{non-empty-lemma}). This implies that $\In(B)$ is a common zero of $\In_{\tilde \nu}(f_j|_{\Kii{\tilde I}})$, $l+1 \leq j \leq m$. \item Since $C \subset V(f_j)$ for each $j =1, \ldots, l$, it follows that $\In(B)$ is a common zero of $\In_{\tilde \nu}(f_j|_{\Kii{\tilde I}})$, $1 \leq j \leq l$ (\cref{branch-lemma}). \end{itemize} This proves the lemma for the case that $\nu$ is the trivial valuation. \\ Now assume $\nu$ is not the trivial valuation. Note that the hypotheses and conclusions do not change if we multiply any of the $f_j$'s by a monomials in $(x_i: i \in I)$. It follows that after a monomial change of coordinates on $(\kk^*)^I$ we may assume that \begin{prooflist} \item $f_j \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, $ 1 \leq j \leq m$, \item \label{nu-standard} $\nu(x_j) = 0 $, $1 \leq j < k$, and $\nu(x_k) = 1$. \end{prooflist} Assumption \eqref{initial-assumption} then implies that $f_1, \ldots, f_m$ have a common zero $a = (a_1, \ldots, a_{k-1}, 0, \ldots, 0)$ with $a_1, \ldots, a_{k-1} \in \ensuremath{\mathbb{K}}^*$. Now consider the translation $(z_1, \ldots, z_n)$ of $(x_1, \ldots, x_n)$ by $a$, and define $X_{\nu^*}$, $E_{\nu^*}$ and $Y'$ as in the proof of the trivial case. Then there is an affine open subset $U$ of $X_{\nu^*}$ with coordinates $(z_1/(z_k)^{\lambda_1}, \ldots, z_n/(z_k)^{\lambda_n})$, where $\lambda_j = \nu^*(z_j)$, $1 \leq j \leq n$, such that $E_{\nu^*} \cap Y' \cap U \neq \emptyset$. Pick $a' \in E_{\nu^*} \cap Y' \cap U$. Now the same arguments as in the proof of the trivial case complete the proof. \end{proof} Let $\mathscr{G}:= (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$. Define $\scrN_{\mscrG}$ and $\scrN^{I}_{\mscrG}$ as in \eqref{NG} and \eqref{NiG}. \begin{cor}[\Cref{weak-lemma}] \label{weak-lemma2} Let $f_1, \ldots, f_n$ be $\mathscr{G}$-admissible polynomials in $(x_1, \ldots, x_n)$. Assume $\multzero{\Gamma_1}{\Gamma_n} < \infty$. Then the following are equivalent: \begin{enumerate} \item \label{all-non-degnerate} $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK non-degenerate at the origin for every $I \subseteq [n]$. \item \label{IG-non-degenerate} $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK non-degenerate at the origin for every $I \in \scrN_{\mscrG}$. \end{enumerate} \end{cor} \begin{proof} We may assume $\multzero{\Gamma_1}{\Gamma_n} \neq 0$ (since otherwise both assertions are always true). We only have to prove \eqref{IG-non-degenerate} $\ensuremath{\Rightarrow}$ \eqref{all-non-degnerate}. It suffices to prove the following claim: ``if there is $I \subseteq [n]$ such that $f_1|_{\kk^I}, \ldots, f_m|_{\kk^I}$ are BKK degenerate at the origin, then there is $\tilde I \in \scrN_{\mscrG}$ such that $f_1|_{\Kii{\tilde I}}, \ldots, f_m|_{\Kii{\tilde I}}$ are BKK degenerate at the origin.''\\ We prove the claim by induction on $n-|I|$. If $|I| = n$, then $I \in \scrN_{\mscrG}$, so there is nothing to prove. Now assume that the claim is true whenever $|I| > k$, $1 \leq k \leq n$. Pick $I \subseteq [n]$ such that $|I| = k$ and $f_1|_{\kk^I}, \ldots, f_m|_{\kk^I}$ are BKK degenerate at the origin. Since $0 < \multzero{\Gamma_1}{\Gamma_n} < \infty$, \cref{mult-support-solution} implies that $|\scrN^{I}_{\mscrG}| \geq k$. If $|\scrN^{I}_{\mscrG}| = k$, then $I \in \scrN_{\mscrG}$, so assume that $|\scrN^{I}_{\mscrG}| > k$. Then \cref{reduction-lemma} implies that there exists $\tilde I \supsetneqq I$ such that $f_1|_{\Kii{\tilde I}}, \ldots, f_n|_{\Kii{\tilde I}}$ are BKK degenerate at the origin. Since $|\tilde I| > k$, we are done by induction. \end{proof} \subsection{Proof of \cref{generic-thm}} \label{generic-proof} In view of \cref{mult-support-solution} it suffices to treat the case that $0 < \multzero{\Gamma_1}{\Gamma_n} < \infty$. At first we establish that $\mathscr{G}$-non-degeneracy is sufficient for the intersection multiplicity to be generic. \begin{convention} Below sometimes we work with $\ensuremath{\mathbb{K}}^{n+1}$ with coordinates $(x_1, \ldots, x_n, t)$. In those cases we usually denote the coordinates of elements of $\ensuremath{\mathbb{K}}^{n+1}$ as pairs, with the last component of the pair denoting the $t$-coordinate. In particular, the origin of $\ensuremath{\mathbb{K}}^{n+1}$ is denoted as $(0,0)$. \end{convention} \begin{claim} \label{sufficient-lemma} Let $f_1, \ldots, f_n$ and $g_1, \ldots, g_n$ be two systems of $\mathscr{G}$-admissible polynomials such that $\multzero{g_1}{g_n} < \multzero{f_1}{f_n}$. Then $f_i$'s are $\mathscr{G}$-degenerate. \end{claim} \begin{proof If $\multzero{f_1}{f_n} = \infty$, \cref{branch-lemma} implies that $f_i$'s are $\mathscr{G}$-degenerate. So we may assume that $\multzero{f_1}{f_n} < \infty$. Let $Z$ be the subscheme of $\ensuremath{\mathbb{K}}^{n+1}$ defined by polynomials $h_i := tg_i + (1-t)f_i \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n,t]$, $1 \leq i \leq n$. \Cref{isolated-family} implies that there exists an open subset $\tilde U$ of $\ensuremath{\mathbb{K}}^{n+1}$ containing $(0,0)$ and $(0,1)$ such that \begin{prooflist}[series=sufficient-list] \item \label{U-tilde-1} $\tilde Z := Z \cap \tilde U$ has dimension $1$. \item \label{U-tilde-1.5} No irreducible component of $\tilde Z$ is `vertical', i.e.\ contained in a hyperplane of the form $\{t = \epsilon\}$ for some $\epsilon \in \ensuremath{\mathbb{K}}$. \item \label{U-tilde-2} For generic $\epsilon \in \ensuremath{\mathbb{K}}$, the points on $\tilde Z \cap \{t = \epsilon\}$ are isolated zeroes of $h_i|_{t = \epsilon}$, $1 \leq i \leq n$. \end{prooflist} Since $\tilde Z$ is a complete intersection, we can write $\tilde Z = \sum m_i \tilde Z_i$, where $\tilde Z_i$'s are irreducible compoents of $\tilde Z$ and $m_i$'s are intersection multiplicities of $h_1, \ldots, h_n$ along $\tilde Z_i$'s. We may assume that $\tilde Z_1$ is the line $\{0\} \times \ensuremath{\mathbb{K}}$. For each $\epsilon \in \ensuremath{\mathbb{K}}$, let $H_\epsilon := V(t - \epsilon) \subseteq \ensuremath{\mathbb{K}}^{n+1}$. Then \begin{align*} m_1 &= H_1 \cdot m_1\tilde Z_1 = H_1 \cdot \tilde Z = \multzero{h_1|_{t=1}}{h_n|_{t=1}} = \multzero{g_1}{g_n} \end{align*} It follows that \begin{align*} H_0 \cdot \tilde Z = \multzero{h_1|_{t=0}}{h_n|_{t=0}} = \multzero{f_1}{f_n} > m_1 \end{align*} Since $H_0$ intersects $\tilde Z$ properly, it follows that there is another irreducible component $\tilde Z_2$ of $\tilde Z$ such that $(0, 0) \in \tilde Z_2$. Pick a branch $B$ of $\tilde Z_2$ at $(0,0)$. Let $\tilde I := I_B \subseteq [n+1]$ and $\tilde \nu := \nuBii{\tilde I}$ (\cref{branch-remark}). Note that $\{n+1\} \subsetneqq \tilde I$. Let $I := \tilde I \cap [n]$ and $\nu := \tilde \nu|_{A^I}$. Since $g_j$ and $f_j$ have the same Newton diagram for each $j$, it follows that $\tilde \nu(f_j) = \tilde \nu(g_j) = \nu(f_j)$. Since $\tilde \nu(t) > 0$, it follows that $\In_{\tilde \nu}(h_j|_{\Kii{\tilde I}}) = \In_\nu(f_j|_{\kk^I})$, $1 \leq j \leq n$. \Cref{branch-lemma} then implies that $\In_\nu(f_1|_{\kk^I}), \ldots, \In_\nu(f_n|_{\kk^I})$ have a common zero in $(\kk^*)^I$. Since $\nu$ is a (positive multiple of some) valuation in $\scrV^I_0$, it follows that $f_i$'s are $\ensuremath{\mathcal{G}}$- degenerate, as required. \end{proof} It remains to prove that $\mathscr{G}$-non-degeneracy is necessary for the intersection multiplicity to be generic. Take $f_1, \ldots, f_n$ which are not $\mathscr{G}$-non-degenerate. We have to show that $\multzero{f_1}{f_n} > \multzero{\Gamma_1}{\Gamma_n}$. Since $\multzero{\Gamma_1}{\Gamma_n} < \infty$, it suffices to consider the case that $0 < \multzero{f_1}{f_n}< \infty$. \\ Let $\scrN_{\mscrG}$ and $\scrN^{I}_{\mscrG}$ be as in \eqref{NG} and \eqref{NiG}. \Cref{weak-lemma2} implies that there is $I \in \scrN_{\mscrG}$ and $\nu \in \scrV^I_0$ such that $\In_\nu(f_j|_{\kk^I})$, $ 1 \leq j \leq n$, have a common zero $z \in (\kk^*)^I$. We may assume that $I = \scrN^{I}_{\mscrG} = \{1, \ldots, k\}$. Choose $y = (y_1, \ldots, y_k, 0, \ldots, 0) \in (\kk^*)^I$ such that $f_j(y) \neq 0$ for all $j$, $1 \leq j \leq k$. Let $C$ be the rational curve on $\kk^I$ parametrized by $c(t):= (c_1(t), \ldots, c_n(t)) :\ensuremath{\mathbb{K}} \to \kk^I$ given by \begin{align} c_j(t) := \begin{cases} z_jt^{\nu_j} + (y_j - z_j)t^{\nu'_j} &\text{if}\ 1 \leq j \leq k, \\ 0 &\text{if}\ k < j \leq n. \end{cases} \label{c_j} \end{align} where $\nu_j := \nu(x_j)$ and $\nu'_j$ is an integer greater than $\nu_j$, $1 \leq j \leq k$. Note that $c(0) = 0$ and $c(1) = y$. Let $g_1, \ldots, g_k$ be $(\Gamma_1, \ldots, \Gamma_k)$-admissible polynomials such that $\In_\nu(g_j|_{\kk^I})(z) \neq 0$ for each $j$, $1 \leq j \leq k$. Then for each $j$, $1 \leq j \leq k$, \begin{align} \ord_t(g_j|_C) = \nu(\Gamma^I_j) < \ord_t(f_j|_C) \label{g_j<f_j} \end{align} Let $\mu_j := \nu(\Gamma^I_j)$, $1 \leq j \leq k$. Then for each $j \leq k$, $t^{-\mu_j}(f_j(c(t))$ and $t^{-\mu_j}(g_j(c(t))$ are polynomials in $t$. Let $U$ be a neighborhood of the origin on $\ensuremath{\mathbb{K}}^n$ such that the origin is the only point in $V(f_1, \ldots, f_n) \cap U$. Then $T := c^{-1}(U) \subseteq \ensuremath{\mathbb{K}}$ is an open neighborhood of $0$ in $\ensuremath{\mathbb{K}}$. Define \begin{align} h_j &:= \begin{cases} (t^{-\mu_j}f_j(c(t)))g_j - (t^{-\mu_j}g_j(c(t)))f_j & \text{if}\ 1 \leq j \leq k,\\ f_j & \text{if}\ k < j \leq n. \end{cases} \label{h_j} \end{align} Identity \eqref{g_j<f_j} implies that \begin{prooflist} \setcounter{prooflisti}{3} \item \label{h_j-at-0} $h_j(x,0)$ is a non-zero constant times $f_j$ for each $j$, $1 \leq j \leq k$. \end{prooflist} Let $Z$ be the subscheme of $U \times T$ defined by $h_j(x,t)$, $1 \leq j \leq n$. Let $\tilde Z$ be the union of the irreducible components of $Z$ containing $(0,0)$. Since $0$ is an isolated point on $V(f_1|_{\kk^I}, \ldots, f_k|_{\kk^I})$, it follows that $(0,0)$ is an isolated point of $Z \cap \{t=0\}$. This implies that $\tilde Z $ has pure dimension one. In particular, \begin{prooflist}[resume] \item the `$t$-axis' $Z_t := \{0\} \times T$ and $\tilde C :=\{ (c(t),t): t \in T\}$ are irreducible components of $\tilde Z$. \end{prooflist} Choose an open subset $U^*$ of $U \times T$ such that $Z \cap U^* = Z_t \cap U^*$. Note that $Z_t \cap U^*= \{0\} \times T^*$ for some open subset $T^*$ of $T$. Then \begin{prooflist}[resume] \item for each $\epsilon \in T^*$, the origin is an isolated zero of $h_1|_{t= \epsilon}, \ldots, h_n|_{t = \epsilon}$. \end{prooflist} Since $\tilde Z$ is a complete intersection, we can write $\tilde Z = \sum m_i \tilde Z_i$, where $\tilde Z_i$'s are irreducible components of $\tilde Z$ and $m_i$'s are intersection multiplicities of $h_1, \ldots, h_n$ along $\tilde Z_i$'s. We may assume that $\tilde Z_1 = Z_t$ and $\tilde Z_2 = \tilde C$. For each $\epsilon \in T$, let $H_\epsilon$ be the hypersurface $U \times \{\epsilon\}$ in $U \times T$. Then for all $\epsilon \in T^*$, \begin{align*} m_1 &= H_\epsilon \cdot m_1Z_t = H_\epsilon \cdot (Z \cap U^*) = \multzero{h_1|_{t=\epsilon}}{h_n|_{t=\epsilon}} \end{align*} where $\multzero{h_1|_{t=\epsilon}}{h_n|_{t=\epsilon}}$ is the intersection multiplicity at the origin of $h_1|_{t=\epsilon}, \ldots, h_n|_{t=\epsilon}$. Now choose a neighborhood $\tilde U$ of $(0,0)$ in $U \times T$ such that $Z \cap \tilde U \cap H_0$ consists of only $\{(0,0)\}$. Since $H_0$ intersects each $\tilde Z_i$ properly, it follows by \ref{h_j-at-0} and definition of intersection multiplicity that \begin{align*} \multzero{f_1}{f_n} &= (Z \cap \tilde U) \cdot H_0 = (m_1 Z_t + m_2 \tilde C + \cdots) \cdot H_0 > m_1 Z_t \cdot H_0 = m_1 \end{align*} Since $h_1|_{t=\epsilon}, \ldots, h_n|_{t=\epsilon}$ are $\mathscr{G}$-admissible for generic $\epsilon \in T^*$, this implies that $\multzero{f_1}{f_n} > \multzero{\Gamma_1}{\Gamma_n}$, as required. \qed \section{Proof of non-degeneracy conditions for the extended BKK bound} \label{generic-bkk-section} In this section we prove \cref{generic-bkk-thm} following the approach outlined in \cref{idea-section}. In \cref{pre-bkk-section} we study properties of exotrivial coordinate subspaces and prove \cref{weak-bkkriterion}. The main technical result of \cref{pre-bkk-section} is \cref{trivial-lemma} which is the basis for property \ref{NTSP-property} of $\mathscr{P}$- non-degeneracy. In \cref{generic-bkk-proof} we use these results to establish \cref{generic-bkk-thm} \subsection{Some properties of exotrivial coordinate subspaces} \label{pre-bkk-section} Throughout this section $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ is a collection of $n$ convex integral polytopes in $(\ensuremath{\mathbb{R}}_{\geq 0})^n$, $f_1, \ldots, f_n$ are $\mathscr{P}$-admissible polynomials in $(x_1, \ldots, x_n)$, and $\mathscr{S}$ is a collection of subsets of $[n]$. In the proof of the result below we use the notion of {\em complete BKK non-degeneracy} (\cref{bkk-defn}). \begin{lemma}[\Cref{absolute-prop}]\label{absolute-cor} Let $I \subseteq [n]$. If $\mathscr{P}$ is $(\kk^*)^I$-exotrivial, then the zero set of $f_1, \ldots, f_n$ on $\ensuremath{\mathbb{K}}^n$ has no isolated point on $(\kk^*)^I$. \end{lemma} \begin{proof} If there exists $\tilde I \supseteq I$ such that $|\NiiP{\tilde I}| < |\tilde I|$, then the statement is clear; so assume $\mathscr{P}$ is $(\kk^*)^I$-trivial and there is $\tilde I \supseteq I$ as in property \ref{absolute-equal} of \cref{absolute-defn}. By restricting all $f_j$'s to $\Kii{\tilde I}$, we may assume that $\tilde I = [n]$. Let $V_f$ be the set of isolated zeroes of $f_1, \ldots, f_n$ in $\ensuremath{\mathbb{K}}^n$. It suffices to consider the case that $V_f \neq \emptyset$.\\ Let $g_1, \ldots, g_n$ be $\mathscr{P}$-admissible polynomials which are {\em completely BKK non-degenerate} on $\ensuremath{\mathbb{K}}^n$ (\cref{bkk-existence}). Let $h_i := tg_i + (1-t)f_i \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n,t]$, $1 \leq i \leq n$. \Cref{isolated-family} implies that there exists an open subset $\tilde U$ of $\ensuremath{\mathbb{K}}^{n+1}$ containing $V_f \times \{0\}$ which satisfies properties \ref{U-tilde-1}--\ref{U-tilde-2} from the proof of \cref{sufficient-lemma}. Let $\tilde Z := V(h_1, \ldots, h_n) \cap \tilde U$ and $H_\epsilon := V(t-\epsilon) \subseteq \ensuremath{\mathbb{K}}^{n+1}$, $\epsilon \in \ensuremath{\mathbb{K}}$. Let $\bar Z, \bar H_\epsilon$ be the closures respectively of $\tilde Z, H_\epsilon$ in $\ensuremath{\mathbb{P}}^n \times \ensuremath{\mathbb{P}}^1$. Pick $z \in V_f$ and an irreducible component $Z^*$ of $\bar Z$ containing $(z,0)$. Since $Z^*$ intersects $\bar H_0$ properly, and since $\bar H_0$ is linearly equivalent to $\bar H_1$, it follows that $Z^*$ intersects $\bar H_1$ as well. Pick a branch $B$ of $Z^*$ at $(z^*,1) \in Z^* \cap \bar H_1$. Let $\tilde I := I_B \subseteq [n+1]$, $\tilde \nu := \nuBii{\tilde I}$ (\cref{branch-remark}), $I^* := \tilde I \cap [n]$ and $\nu^* := \tilde \nu|_{\Aii{I^*}}$. Since $g_j$ and $f_j$ have the same Newton polytope for each $j$, it follows that $\tilde \nu(f_j) = \tilde \nu(g_j) = \nu^*(g_j)$. Since $\tilde \nu(t-1) > 0$, it follows that $\tilde \nu(t) = 0$ and $\In_{\tilde \nu}(h_j|_{\Kii{\tilde I}}) = t\In_{\nu^*}(g_j|_{\Kii{I^*}})$, $1 \leq j \leq n$. \Cref{branch-lemma} then implies that $\In_{\nu^*}(g_1|_{\Kii{I^*}}), \ldots, \In_{\nu^*}(g_n|_{\Kii{I^*}})$ have a common zero in $\Kstarii{I^*}$. But then $\mathscr{P}$ is {\em not} $\Kstarii{I^*}$-trivial (\cref{bkk-lemma}). By our assumption on $I$ it follows that $I^* \subsetneqq I$. Since $z$ is in the closure of $\Kstarii{I^*}$ in $\ensuremath{\mathbb{K}}^n$, it follows that $z \not\in (\kk^*)^I$, as required. \end{proof} \begin{lemma}[\Cref{weak-bkkriterion}] \label{weak-bkk-lemma} Let $f_1, \ldots, f_n \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$ and the notations be as in \cref{weak-bkkriterion}. Then \begin{enumerate} \item \label{NSP-non-degenerate} property \ref{TSPrime-property} of $\mathscr{P}$-non-degeneracy holds iff it holds with $\scrT'_{\mscrS,\mscrP}$ replaced by $\scrN_{\mscrS,\mscrP}$, \item \label{TSPstar-non-degenerate} property \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy holds iff it holds with $\scrT_{\mscrS,\mscrP}$ replaced by $\scrT^*_{\mscrS,\mscrP}$. \end{enumerate} \end{lemma} \begin{proof} Since $\scrN_{\mscrS,\mscrP} \subseteq \scrT_{\mscrS,\mscrP}'$, for assertion \eqref{NSP-non-degenerate} we only have to show the $(\Leftarrow)$ implication. It follows by \cref{absolute-defn} that for every $I \in \scrT_{\mscrS,\mscrP}' \setminus \scrN_{\mscrS,\mscrP}$, there exists $\tilde I \in \scrN_{\mscrS,\mscrP}$ such that $I \subsetneqq \tilde I$ and $|\NiiP{I^*}| > |I^*|$ for each $I^*$ such that $I \subseteqq I^* \subsetneqq \tilde I$. \Cref{herrero} implies that $|\NiiP{\tilde I}| = |\tilde I|$. Since restricting all $f_j$'s to $\Kii{\tilde I}$ yields a system with the same number of non-zero polynomials as the number of variables, it suffices to prove the following claim: ``if there is $I \subseteq [n]$ such that $|\scrN^{I}| > |I|$ and $f_1|_{\kk^I}, \ldots, f_n|_{\kk^I}$ are BKK degenerate at infinity (resp.\ at $(\overbar \mathscr{S} \cup \scrT_{\mscrS,\mscrP})^I$), then there is $\tilde I \supsetneqq I$ such that $f_1|_{\Kii{\tilde I}}, \ldots, f_n|_{\Kii{\tilde I}}$ are BKK degenerate at infinity (resp.\ at $(\overbar \mathscr{S} \cup \scrT_{\mscrS,\mscrP})^{\tilde I}$).'' This follows by an immediate application of \cref{reduction-lemma}, completing the proof of assertion \eqref{NSP-non-degenerate}. Assertion \eqref{TSPstar-non-degenerate} follows from a similar application of \cref{reduction-lemma}. \end{proof} \Cref{trivial-lemma,trivial-cor} below are key to property \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy. \begin{lemma} \label{trivial-lemma} Assume $\mathscr{P}$ is $(\kk^*)^n$-trivial. Let $\nu$ be a monomial valuation in $\scrV$ centered at $(\kk^*)^I$, $I \subseteq [n]$. Assume $\In_\nu(f_1), \ldots,\In_\nu(f_n)$ have a common zero $a \in (\kk^*)^n$. Then $\pi_{I}(a) \in (\kk^*)^I$ is a non-isolated point of the zero-set $V(f_1, \ldots, f_n)$ of $f_1, \ldots, f_n$ on $\ensuremath{\mathbb{K}}^n$. \end{lemma} \begin{proof} It is straightforward to check that $\pi_{I}(a)$ is in $V(f_1, \ldots, f_n)$; we only have to show that it is non-isolated in there. \Cref{herrero} implies that there is $J \subseteq [n]$ such that $p:= \dim(\sum_{j \in J} \ensuremath{\mathcal{P}}_j) <|J|$. Let $\Pi$ be the (unique) $p$-dimensional linear subspace of $\ensuremath{\mathbb{R}}^n$ such that $\sum_{j \in J} \ensuremath{\mathcal{P}}_j$ is contained in a translate of $\Pi$. Let $\Pi_0 := \{\alpha \in \Pi: \nu \cdot \alpha = 0\}$ and $r := \dim(\Pi_0)$. Then either $p=r$, or $p = r+1$. We consider these cases separately:\\ {\bf Case 1: $\*{p = r}$.} Let $\alpha_1, \ldots, \alpha_r$ be a basis of $\Pi_0$. In this case $\Pi = \Pi_0$, so that for each $j \in J$, $f_j$ is $\nu$-homogeneous and is a linear combination of monomials in $ x^{\alpha_1}, \ldots, x^{\alpha_r}$. For each $i$, $1 \leq i \leq n$, let $c_i := a^{\alpha_i}$. Let $\tilde Y$ be the sub-variety of $(\kk^*)^n$ determined by $x^{\alpha_i} - c_i$, $1 \leq i \leq r$ and $Y$ be the closure of $\tilde Y$ in $\ensuremath{\mathbb{K}}^n$. Then \begin{prooflist} \item $Y$ is irreducible of codimension $r$ in $\ensuremath{\mathbb{K}}^n$. \item $Y \subseteq V(f_j :j \in J)$, since $\tilde Y \subseteq V(f_j :j \in J)$. \item \label{C} $\pi_I(a) \in Y$. Indeed, it is clear if $\nu$ is the trivial valuation. Otherwise let $C$ be the curve parametrized by $x_i := a_it^{\nu_i}$, $t \in \ensuremath{\mathbb{K}}$. Then $C \cap (\kk^*)^n \subseteq \tilde Y$, so that $C \subseteq Y$. Now note that $\pi_I(a) \in C$. \end{prooflist} It follows that one of the irreducible components of $V(f_j :j \in J)$ containing $\pi_I(a)$ has codimension smaller than $|J|$. This implies the lemma in this case.\\ {\bf Case 2: $\*{p = r+1}$.} Choose a basis $\alpha_1, \ldots, \alpha_n$ of $\ensuremath{\mathbb{Z}}^n$ such that \begin{prooflist}[resume] \item $\alpha_1, \ldots, \alpha_r$ is a basis of $\Pi_0$. \item $\alpha_1, \ldots, \alpha_{n-1}$ is a basis of $\nu^\perp := \{\alpha \in \ensuremath{\mathbb{Z}}^n: \nu \cdot \alpha = 0\}$. \item $\alpha_1, \ldots, \alpha_{r}, \alpha_n$ is a basis of $\Pi$. \item $\nu \cdot \alpha_n = 1$. \end{prooflist} Let $\tilde Y$ and $Y$ be as in the proof of case 1. Let $y_i := x^{\alpha_i}$, $1 \leq i \leq n$. Then $y_i$'s form a system of coordinates on $(\kk^*)^n$ and the projection onto $(y_{r+1}, \ldots, y_n)$ restricts to an isomorphism $\tilde Y \cong (\ensuremath{\mathbb{K}}^*)^{n-r}$. Let $\beta_1, \ldots, \beta_n \in \ensuremath{\mathbb{Z}}^n$ be such that $x_j = \prod_{i=1}^n y_i^{\beta_{j,i}}$, $1 \leq j \leq n$. Then $Y$ is the closure of the image of the map \begin{align*} &(\ensuremath{\mathbb{K}}^*)^{n-r} \ni (y_{r+1}, \ldots, y_n) \mapsto (\tilde c_1 \prod_{i=r+1}^n y_i^{\beta_{1,i}}, \ldots, \tilde c_n \prod_{i=r+1}^n y_i^{\beta_{n,i}}) \in \ensuremath{\mathbb{K}}^n, \\ & \text{where}\ \tilde c_j = \prod_{i=1}^r c_i^{\beta_{j,i}},\ 1 \leq j \leq n. \end{align*} Let $Y'$ be the closure of the image of the map \begin{align*} &(\ensuremath{\mathbb{K}}^*)^{n-r} \ni (y_{r+1}, \ldots, y_n) \mapsto (\prod_{i=r+1}^n y_i^{\beta_{1,i}}, \ldots, \prod_{i=r+1}^n y_i^{\beta_{n,i}}) \in \ensuremath{\mathbb{K}}^n \end{align*} Then $Y$ is isomorphic to $Y'$ via the map $(x_1, \ldots, x_n) \mapsto (\tilde c_1^{-1}x_1, \ldots, \tilde c_n^{-1}x_n)$. Note that $Y'$ is an affine toric variety corresponding to the semigroup generated by $\beta'_j := (\beta_{j,r+1}, \ldots, \beta_{j,n})$, $1 \leq j \leq n$. For each $j$, $1 \leq j \leq n$, \begin{align} \nu(x_j) = \sum_{i=1}^n \beta_{j,i}\nu(y_i) = \beta_{j,n} \label{beta-j-n} \end{align} Let $\Theta$ be the coordinate hyperplane of $\ensuremath{\mathbb{Z}}^{n-r}$ generated by the first $n-r-1$ axes. \eqref{beta-j-n} implies that $\beta'_j$'s lie on $\Theta$ for $j \in I$ and `above' $\Theta$ otherwise. Let $\Sigma$ be the cone in $\ensuremath{\mathbb{R}}^{n-r}$ generated by $\beta'_1, \ldots, \beta'_n$. Note that $(0, \ldots, 0, 1)$ is in the dual $\check{\Sigma}$ of $\Sigma$. Choose an edge $E$ of $\check{\Sigma}$ such that \begin{prooflist}[resume] \item \label{E-1} $E$ is an edge of the face of $\check \Sigma$ whose relative interior contains $(0, \ldots, 0, 1)$. \item the last coordinate of each non-zero element on $E$ is positive. \end{prooflist} Let $Z'$ be the torus invariant divisor of $Y'$ corresponding to $E$ and $Z$ be the isomorphic image of $Z'$ in $Y$. Let $\eta = (\eta_{r+1}, \ldots, \eta_n)$ be the smallest non-zero element on $E$ with integer coordinates. For each $b := (b_{r+1}, \ldots, b_n) \in (\ensuremath{\mathbb{K}}^*)^{n-r}$, consider the curve $\gamma_b$ parametrized by $\ensuremath{\mathbb{K}}^* \ni t \mapsto (b_{r+1}t^{\eta_{r+1}}, \ldots, b_nt^{\eta_n}) \in (\ensuremath{\mathbb{K}}^*)^{n-r}$. Then \begin{prooflist}[resume] \item \label{limit-points} the `limit at zero' of (the image of) $\gamma_b$ is a point on $Z$; we denote it by $\bar \gamma_b(0)$. (The closure of the image of) $\gamma_b$ is transversal to $Z$ at $\bar \gamma_b(0)$. The set of all $\bar \gamma_b(0)$ as $b$ varies over $(\ensuremath{\mathbb{K}}^*)^{n-r}$ is open in $Z$. \item It follows that $\ord_Z(y_n|_Y) = \ord_t(y_n|_{\gamma_b}) = \ord_t(b_nt^{\eta_n}) = \eta_n > 0$. \end{prooflist} Fix $i$, $1 \leq i \leq s$. Write $f_i$ as $f_i = f_{i,0} + f_{i,1} + \cdots $, where $f_{i,j}$'s are $\nu$-homogeneous with $\nu(f_{i,0}) < \nu(f_{i,1}) < \cdots$. Let $\alpha \in \supp(f_{i,0})$, then for each $j$, $f_{i,j}/x^\alpha = \sum_\beta \lambda_{i,j,\beta} y_1^{\beta_1} \cdots y_r^{\beta_r}y_n^{\nu(f_{i,j})}$ with $\lambda_{i,j,\beta} \in \ensuremath{\mathbb{K}}$. Pick $b \in (\ensuremath{\mathbb{K}}^*)^{n-r}$. Since $f_{i,0}(a) = 0$, it follows that \begin{align*} (f_i/x^\alpha)|_{\gamma_b} = \sum_{j,\beta} \lambda_{i,j,\beta} c_1^{\beta_1} \cdots c_r^{\beta_r}(b_nt^{\eta_n})^{\nu(f_{i,j})} = \sum_{j > 0} \sum_\beta \lambda_{i,j,\beta} c_1^{\beta_1} \cdots c_r^{\beta_r}b_n^{\nu(f_{i,j})} t^{\nu(f_{i,j})\eta_n} \end{align*} Since $\bar \gamma_b(0) \in Z \subseteq \ensuremath{\mathbb{K}}^n$, it follows that $\ord_t(x^\alpha) \geq 0$, so that $\ord_t(f_i|_{\gamma_b}) > 0$. Observation \ref{limit-points} then implies that $Z \subseteq V(f_i)$. It follows that $Z \subseteq V(f_1, \ldots, f_n)$. On the other hand, \ref{C} and \ref{E-1} imply that $\pi_I(a) \in Z$. Since $\dim(Z) = n- r - 1 > n-|J|$, this proves the lemma. \end{proof} \begin{cor} \label{trivial-cor} Let $I \subseteq J \subseteq [n]$ and $\nu \in \Vii{J}$ with center in $(\kk^*)^I$. Assume $\mathscr{P}$ is $\Kstarii{J}$-exotrivial and $\In_\nu(f_1|_{J}) \ldots,\In_\nu(f_n|_{J})$ have a common zero $a \in (\kk^*)^n$. Then $\pi_{I}(a) \in (\kk^*)^I$ is a non-isolated point of the zero-set $V(f_1, \ldots, f_n)$ of $f_1, \ldots, f_n$ on $\ensuremath{\mathbb{K}}^n$. \end{cor} \begin{proof} Note that $\pi_{I}(a) \in V(f_1, \ldots, f_n)$. It suffices to consider the case that $J$ satisfies property \ref{absolute-equal} of \cref{absolute-defn} (since in the other case the statement is clear). Then there is $\tilde I \supseteq J$ such that $\mathscr{P}$ is $\Kstarii{\tilde I}$-trivial, $|\NiiP{\tilde I}| = |\tilde I|$, and \begin{itemize} \item either $\tilde I = J$, \item or $|\NiiP{I^*}| > |I^*|$ for each $I^*$ such that $J \subseteq I^* \subsetneqq \tilde I$. \end{itemize} Replacing $f_j$'s by $f_j|_{\Kii{\tilde I}}$ and then applying \cref{reduction-lemma} we see that the hypothesis of the corollary remains true if we replace $J$ by $\tilde I$ and therefore the corollary reduces to the case that $J = [n]$. Then it follows from \cref{trivial-lemma}. \end{proof} \subsection{Proof of \cref{generic-bkk-thm}} \label{generic-bkk-proof} Assertion \eqref{generic-assertion-0} follows from a straightforward application of \cref{absolute-prop}. Assertion \eqref{isolated-assertion} follows from \cref{absolute-prop} and \cref{curve-claim}. \begin{claim}\label{curve-claim} Let $\mathscr{S}' := \overbar\mathscr{S} \cup \scrT_{\mscrS,\mscrP}$. Assume the (Krull) dimension of $V(f_1, \ldots, f_n) \cap \kk^n_{\mscrS'}$ is non-zero. Then $f_1, \ldots, f_n$ are $\mathscr{P}$-degenerate on $\kk^n_{\overbar \mscrS}$. \end{claim} \begin{proof} Take an irreducible curve $C \subseteq V(f_1, \ldots, f_n) \cap \kk^n_{\mscrS'}$ and let $\kk^I$ be the smallest coordinate subspace of $\ensuremath{\mathbb{K}}^n$ which contains $C$. Since $I \in \scrT'_{\mscrS,\mscrP}$, applying \cref{branch-lemma} with a branch $B$ of $C$ at infinity shows that property \ref{TSPrime-property} of $\mathscr{P}$-non-degeneracy fails. \end{proof} It remains to prove the second assertion. The $(\Rightarrow)$ implication follows from \cref{sufficiently-bkk} below. \begin{claim} \label{sufficiently-bkk} If $f_1, \ldots, f_n$ and $g_1, \ldots, g_n$ are two systems of $\mathscr{P}$-admissible polynomials such that $\multiso{f_1}{f_n}_{\kk^n_{\mscrS'}} < \multiso{g_1}{g_n}_{\kk^n_{\mscrS'}}$, then $f_i$'s are $\mathscr{P}$-degenerate on $\kk^n_{\overbar \mscrS}$. \end{claim} \begin{proof} Let $h_i := tg_i + (1-t)f_i \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n,t]$, $1 \leq i \leq n$. Let $V_f$ (resp.\ $V_g$) be the set of isolated zeroes of $f_1, \ldots, f_n$ (resp.\ $g_1, \ldots, g_n$) on $\kk^n_{\mscrS'} $. \Cref{isolated-family} implies that there exists an open subset $\tilde U$ of $\KnSS{\mathscr{S}'} \times \ensuremath{\mathbb{K}}$ which contains $V_f \times \{0\}$ and $V_g \times \{1\}$, and satisfies properties \ref{U-tilde-1} --\ref{U-tilde-2} from the proof of \cref{sufficient-lemma}. Without any loss of generality we may also assume that \begin{prooflist}[resume=sufficient-list] \item \label{U-tilde-4} $\tilde U \cap \{t = 1\}$ (resp.\ $\tilde U \cap \{t = 0\}$) does not intersect any non-isolated zeroes of $g_1, \ldots, g_n$ (resp.\ $f_1, \ldots,f_n$) in $\KnSS{\mathscr{S}'}$ . \end{prooflist} Define $\tilde Z, \bar Z, H_\epsilon, \bar H_\epsilon$ as in the proof of \cref{absolute-cor}. Property \ref{U-tilde-4} implies that $$H_1 \cdot \tilde Z = \multiso{g_1}{g_n}_{\KnSS{\mathscr{S}'}} = \multiso{g_1}{g_n}_{\kk^n_\mscrS} > \multiso{f_1}{f_n}_{\kk^n_\mscrS} = H_0 \cdot \tilde Z$$ Since $\bar H_1 \cdot \bar Z = \bar H_0 \cdot \bar Z$, it follows that there exists $(z,0) \in \bar Z$ such that one of the following holds: \begin{enumerate}[label= (\alph{enumi})] \item \label{z-S} $z \in \Kstarii{S}$ for some $S \in \mathscr{S}'$. \item \label{z-infty} $z \in \ensuremath{\mathbb{P}}^n \setminus \ensuremath{\mathbb{K}}^n$. \item \label{z-non-isolated} $z$ is a non-isolated point of $V(f_1, \ldots, f_n) \cap \kk^n_{\mscrS'}$. \end{enumerate} Pick $z$ satisfying one of the preceding conditions. Let $Z^*$ be an irreducible component of $\tilde Z$ containing $(z,0)$ and let $B$ be a branch of $Z^*$ at $(z,0)$. Define $\tilde I, \tilde \nu, I^* $ and $\nu^*$ as in the proof of \cref{absolute-cor}. Note that \begin{prooflist}[resume] \item \label{I^*-1} $I^* \not\in \mathscr{S}'$ (since $Z^* \cap (\Kstarii{I^*} \times \ensuremath{\mathbb{K}})$ non-empty and open in $Z^*$). \end{prooflist} Property \ref{U-tilde-2} from the proof of \cref{sufficient-lemma} implies that there are $\mathscr{P}$-admissible systems of polynomials which have isolated zero(es) on $\Kstarii{I^*}$. \Cref{absolute-cor} then implies that $\mathscr{P}$ is not $(\kk^*)^I$-exotrivial. Combining this with observation \ref{I^*-1}, we have \begin{prooflist}[resume] \item \label{I^*-2} $I^* \in \scrT'_{\mscrS,\mscrP}$. \end{prooflist} Since $g_j$ and $f_j$ have the same Newton polytope for each $j$, it follows that $\tilde \nu(f_j) = \tilde \nu(g_j) = \nu^*(f_j)$. Since $\tilde \nu(t) > 0$, it follows that $\In_{\tilde \nu}(h_j|_{\Kii{\tilde I}}) = \In_{\nu^*}(f_j|_{\Kii{I^*}})$, $1 \leq j \leq n$. \Cref{branch-lemma} then implies that \begin{prooflist}[resume] \item \label{common-observation} $\In_{\nu^*}(f_1|_{\Kii{I^*}}), \ldots, \In_{\nu^*}(f_n|_{\Kii{I^*}})$ have a common zero in $\Kstarii{I^*}$. \end{prooflist} Now note that \begin{prooflist}[resume] \item \label{KnS} If \ref{z-S} holds, then $\nu^* \in \VSii{I^*}$, so that \ref{common-observation} implies that $f_1|_{\Kii{I^*}}, \ldots, f_n|_{\Kii{I^*}}$ are BKK degenerate at $\mathscr{S}'^I$. \item \label{infinity} If \ref{z-infty} holds, then $-\nu^* \in \wtsii{I^*}$, so that \ref{common-observation} implies that $f_1|_{\Kii{I^*}}, \ldots, f_n|_{\Kii{I^*}}$ are BKK degenerate at infinity. \item \label{z-non-isolated-1} If \ref{z-non-isolated} holds, then \cref{curve-claim} implies that $f_1, \ldots, f_n$ are $\mathscr{P}$-degenerate on $\kk^n_{\overbar \mscrS}$. \end{prooflist} Observations \ref{KnS}--\ref{z-non-isolated-1} imply that in every scenario $f_1, \ldots, f_n$ are $\mathscr{P}$-degenerate on $\kk^n_{\overbar \mscrS}$. \end{proof} It remains to prove that $\mathscr{P}$-non-degeneracy on $\kk^n_{\overbar \mscrS}$ is necessary for the number (counted with multiplicity) of isolated solutions in $\kk^n_{\mscrS'}$ to be maximal. So take $\mathscr{P}$-admissible $f_1, \ldots, f_n$ which are not $\mathscr{P}$-non-degenerate on $\kk^n_{\overbar \mscrS}$. We will show that $\multiso{f_1}{f_n}_{\kk^n_{\mscrS'}} < \mult{\scrP_1}{\scrP_n}_{\kk^n_{\mscrS'}}$. \begin{claim} \label{nu-claim} There exists $I \in \scrN_{\mscrS,\mscrP}$ and $\nu \in \scrV^I$ such that $\In_\nu(f_1|_{\kk^I}), \ldots, \In_\nu(f_n|_{\kk^I})$ have a common zero $z$ in $(\kk^*)^I$, and one of the following holds: \begin{enumerate}[label= (\alph{enumi}$'$)] \item \label{nu-S} $\nu \in \scrV^I_{\mscrS'}$, \item\label{nu-infinity} $-\nu \in \Omega^{I}_\infty$, or \item \label{nu-non-isolated} $\nu$ is centered at $\Kstarii{I'}$ for some $I' \in \scrT'_{\mscrS,\mscrP}$ and $\pi_{I'}(z)$ is non-isolated in $V(f_1, \ldots, f_n)$. \end{enumerate} \end{claim} \begin{proof} If property \ref{TSPrime-property} of $\mathscr{P}$-non-degeneracy fails, then \cref{weak-bkk-lemma} implies that the claim is true and either \ref{nu-S} or \ref{nu-infinity} holds. On the other hand, if property \ref{NTSP-property} of $\mathscr{P}$-non-degeneracy fails, then there exist $I' \subseteq J \subseteq [n]$ such that $J \in \scrT_{\mscrS,\mscrP}$, $I ' \in \scrT'_{\mscrS,\mscrP}$, and $\In_{\eta}(f_1|_{J}) \ldots,\In_{\eta}(f_n|_{J})$ have a common zero $a \in (\kk^*)^n$ for some $\eta \in \Vii{J}$ with center in $\Kstarii{J'}$. \Cref{trivial-cor} implies that $\pi_{I'}(a)$ is a non-isolated point of $V(f_1, \ldots, f_n)$. Now pick the smallest $I \supseteq I'$ such that $|\scrN^{I}_{\mscrP}| = |I|$ (where $\scrN^{I}_{\mscrP}$ is as in \eqref{NiP}). Since $J$ is not exotrivial, it follows that $I \in \scrN_{\mscrS,\mscrP}$. An application of \cref{reduction-lemma} (with $I = I'$ and $\nu = $ the trivial valuation) implies that there exists $\nu \in \scrV^I$ centered at $\Kstarii{I'}$ and a common zero $z$ of $\In_\nu(f_1|_{\kk^I}), \ldots, \In_\nu(f_n|_{\kk^I})$ in $(\kk^*)^n$ such that $\pi_{I'}(z) = \pi_{I'}(a)$. Therefore \ref{nu-non-isolated} holds, and it completes the proof of the claim. \end{proof} Let $I,\nu, z$ be as in \cref{nu-claim}. Since $I \in \scrN_{\mscrS,\mscrP}$, it follows that $|I| = |\scrN^{I}_{\mscrP}|$ . We may assume that $I = \scrN^{I}_{\mscrP} = \{1, \ldots, k\}$. Choose $(\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_k)$-admissible polynomials $g_1, \ldots, g_k$ such that \begin{defnlist}[resume] \item $g_1|_{\kk^I}, \ldots, g_k|_{\kk^I}$ are completely BKK non-degenerate (\cref{bkk-defn}). \item $\In_\nu(g_j|_{\kk^I})(z) \neq 0$ for each $j$, $1 \leq j \leq k$. \end{defnlist} \begin{claim} \label{isolated-claim} For each $y \in (\kk^*)^I$, let $h_{y,i}(x) := g_i(y)f_i(x) - f_i(y)g_i(x)$, $1 \leq i \leq k$. \begin{enumerate} \item \label{dominant-claim} The rational map $\phi: (\ensuremath{\mathbb{K}}^*)^k \dashrightarrow (\ensuremath{\mathbb{K}}^*)^k$ induced by $ (g_1/f_1, \ldots, g_k/f_k)$ is dominant. \item There is a non-empty open subset $W$ of $(\kk^*)^I$ such that every $y \in W$ is an isolated zero of $h_{y,1}, \ldots, h_{y,k}$. \end{enumerate} \end{claim} \begin{proof} \Cref{bkk-deformation} implies that for generic $c := (c_1, \ldots, c_k) \in (\ensuremath{\mathbb{K}}^*)^k$, $g_1 - c_1f_1, \ldots, g_k - c_kf_k$ are BKK-non-degenerate, so that $\phi^{-1}(c)$ is finite (\cref{bkk-lemma}). On the other hand, since $\mathscr{P}$ is $(\kk^*)^I$-non-trivial, it follows that $\multiso{\ensuremath{\mathcal{P}}_1}{\ensuremath{\mathcal{P}}_k}_{(\kk^*)^I} > 0$, so that \cref{sufficiently-bkk,bkk-lemma} imply that $\phi^{-1}(c)$ is non-empty for generic $c \in(\ensuremath{\mathbb{K}}^*)^k$. This implies the first assertion. For the second assertion, take $W$ to be an open subset of $(\kk^*)^I \setminus V(f_1\cdots f_k)$ such that $\phi^{-1}(\phi(w))$ is finite for all $w \in U$. \end{proof} Let $W$ be as in \cref{isolated-claim}. Choose $y = (y_1, \ldots, y_k, 0, \ldots, 0) \in W\setminus V(f_1\cdots f_k)$. Let $C$ be the rational curve on $\kk^I$ parametrized by $c(t) = (c_1(t), \ldots, c_n(t)) :\ensuremath{\mathbb{K}}^* \to \kk^I$ defined as in \eqref{c_j}. Then $c(1) = y$. For $c(0)$ there are three possible options: \begin{defnlist}[resume] \item \label{nu-S-1} If \ref{nu-S} holds, then $c(0) \in \Kstarii{S}$ for some $S \in \mathscr{S}'$. \item \label{nu-infinity-1} If \ref{nu-infinity} holds, then at least one $c_j(t)$ has a pole at $0$; in other words, $c(0)$ is a point at infinity. \item \label{nu-non-isolated-1} If \ref{nu-non-isolated} holds, then $c(0)$ is a non-isolated zero of $f_1, \ldots, f_n$ in $\ensuremath{\mathbb{K}}^n$. \end{defnlist} Let $V_f$ be the set of isolated zeroes of $f_1, \ldots, f_n$ in $\kk^n_{\mscrS'}$. Define $h_1, \ldots, h_n$ as in \eqref{h_j}. \Cref{isolated-family} implies that there is an open subset $U^*$ of $\ensuremath{\mathbb{K}}^{n+1}$ containing $V_f \times \{0\}$ such that $Z^* := V(h_1, \ldots, h_n) \cap U^*$ has only one dimensional `non-vertical' components. Let $\tilde Z$ be the union of irreducible components of $Z^*$ containing points from $V_f \times \{0\}$. For each $\epsilon \in \ensuremath{\mathbb{K}}$, let $H_\epsilon$ be the hyperplane $\{t = \epsilon\}$ in $\ensuremath{\mathbb{K}}^{n+1}$. Then for generic $\epsilon \in \ensuremath{\mathbb{K}}$, \begin{align*} \multiso{h_1|_{t=\epsilon}}{h_n|_{t=\epsilon}}_{\kk^n_{\mscrS'}} \geq H_\epsilon \cdot \tilde Z = H_0 \cdot \tilde Z = \multiso{f_1}{f_n}_{\kk^n_{\mscrS'}} \end{align*} On the other hand, observations \ref{nu-S-1}--\ref{nu-non-isolated-1} imply that $\{(t,c(t)): t \in \ensuremath{\mathbb{K}}^*\}$ does not belong to any component of $\tilde Z$. Since for generic $\epsilon \in \ensuremath{\mathbb{K}}^*$, $c(\epsilon)$ is an isolated zero of $h_1|_{t=\epsilon}, \ldots, h_n|_{t=\epsilon}$ (\cref{isolated-claim}), it follows that $\multiso{h_1|_{t=\epsilon}}{h_n|_{t=\epsilon}}_{\kk^n_{\mscrS'}} > \multiso{f_1}{f_n}_{\kk^n_{\mscrS'}}$, as required. \qed \subsection{Proof of \cref{non-isolated-cor}}\label{non-isolated-section} \Cref{b-example}\ref{b-example-generic-1} validates the first assertion, and the second assertion follows from \cref{curve-claim}. \qed \section{Proof of the intersection multiplicity formula} \label{multiplicity-section} In this section we prove \cref{multiplicity-thm} using the approach outlined in \cref{stable-remark}. The computation of intersection multiplicity becomes easier if a generic system satisfies a property which is stronger than \eqref{bkk-bar-0}; \cref{pre-multiplicity-section} is devoted to the proof of existence of such systems. The proof of \cref{multiplicity-thm} is then given in \cref{multiplicity-proof-section}. \subsection{Strongly $\mathscr{G}$-non-degenerate systems} \label{pre-multiplicity-section} \begin{defn} \label{strongly-generic-defn} Let $\mathscr{G}:= (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$ and $f_1, \ldots, f_n$ be polynomials in $(x_1, \ldots, x_n)$. We say that $f_i$'s are {\em strongly $\mathscr{G}$-non-degenerate} iff \begin{defnlist} \item \label{admissible} $f_i$'s are $\mathscr{G}$-admissible. \item \label{strongly-non-degenerate} For all non-empty $J \subseteq [n]$, $\{f_j: j \in J\}$ are {\em completely BKK non-degenerate} (\cref{bkk-defn}). \item \label{intersectionally-non-degenerate} For all non-empty $I, J, J' \subseteq [n]$ such that $|J| = n - |I|$, $\kk^I$ is an irreducible component of $V(f_j: j \in J)$, $J' \subseteq [n]\setminus J$ and $|J'| = k-1$, the following holds: \begin{align} \parbox{.57\textwidth}{If $V$ is any irreducible component of $V(f_j: j \in J)$ distinct from $\kk^I$, then $V \cap \kk^I \cap V(f_{j'}: j' \in J')$ is finite.} \label{I-J-J'} \end{align} \item \label{recursively-non-degenerate} For all $I\subsetneqq [n]$ and $J \subseteq [m]$ such that $|I| = |J|$, define \begin{align*} & f_{I,j} := f_j|_{x_{i'_1} = 1, \ldots, x_{i'_{k'}} =1},\ \text{for all $j \in J$,} \end{align*} where $i'_1, \ldots, i'_{k'}$ are elements of $[n]\setminus I$. Then $(f_{I,j}: j \in J)$ are completely BKK non-degenerate. \end{defnlist} \end{defn} We now show the existence of strongly non-degenerate polynomials. Let $\mathscr{G}:= (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$. Fix polytopes $\ensuremath{\mathcal{P}}_j$ such that $\Gamma_j$ is the Newton diagram of a polynomial with Newton polytope $\ensuremath{\mathcal{P}}_j$, $1 \leq j \leq n$. Let $\ensuremath{\mathcal{A}}_j$ be the space of polynomials with support in $\ensuremath{\mathcal{P}}_j$, $1 \leq j \leq n$, and let $\ensuremath{\mathcal{A}} := \prod_{j=1}^n \ensuremath{\mathcal{A}}_j$. \begin{lemma} \label{strong-existence} Assume $\multzero{\Gamma_1}{\Gamma_n} < \infty$. Then the set of strongly $\mathscr{G}$-non-degenerate systems of polynomials contains a non-empty Zariski open subset of $\ensuremath{\mathcal{A}}$. \end{lemma} \begin{proof} \Cref{bkk-existence} implies that the set of systems of polynomials which satisfy properties \ref{admissible}, \ref{strongly-non-degenerate} and \ref{recursively-non-degenerate} contains a non-empty Zariski open subset of $\ensuremath{\mathcal{A}}$. We now show that the same is true for polynomials satisfying property \ref{intersectionally-non-degenerate}. \\ Fix $I, J, J' \subseteq [n]$ as in condition \ref{intersectionally-non-degenerate} of \cref{strongly-generic-defn}. It suffices to show that there is a non-empty Zariski open set $U_{I,J,J'}$ of $\ensuremath{\mathcal{A}}$ such that every $(f_1, \ldots, f_n) \in U_{I,J,J'}$ satisfies \eqref{I-J-J'}. We may assume that $I = \{1, \ldots, k\}$. Let $A_J$ be the coordinate ring of $\ensuremath{\mathcal{A}}_J := \prod_{j \in J} \ensuremath{\mathcal{A}}_j$ and $K_J$ be the field of fractions of $A_J $. Let $R := K_J(x_1, \ldots, x_k)[x_{k+1}, \ldots, x_n]$ and $\ensuremath{\mathfrak{m}}$ be the maximal ideal of $R$ generated by $x_{k+1}, \ldots, x_n$. Since $V(f_j: j \in J)$ is a complete intersection near generic points of $\kk^I$, it follows that the ideal $\ensuremath{\mathfrak{q}}_J$ generated by $f_j$, $j \in J$ has finite codimension in $R_{\ensuremath{\mathfrak{m}}}$. As in the proof of \cref{generic-length}, compute a standard basis of $\ensuremath{\mathfrak{q}}_J$ in $R_{\ensuremath{\mathfrak{m}}}$ and consider the product $h \in A_J[x_1, \ldots, x_k]$ of all the coefficients of all the monomials in $(x_{k+1}, \ldots, x_n)$ that appear in the basis elements. Let $W$ be the zero set of $h$ in $\kk^I \times \ensuremath{\mathcal{A}}_J$. If $(x, (f_j: j \in J)) \in (\kk^I \times \ensuremath{\mathcal{A}}_J) \setminus W$, then $\kk^I$ is the only irreducible component of $V(f_j: j \in J)$ containing $x$. \\ Let $\pi_J: W \to \ensuremath{\mathcal{A}}_J$ be the natural projection. If $\pi_J$ is not dominant, then we are done. So assume there is an irreducible component $W'$ of $W$ which projects dominantly to $\ensuremath{\mathcal{A}}_J$. Let \begin{align*} Y := \{(x, (f_j: j \in J), (f_{j'}: j' \in J')): x \in V(f_{j'}: j' \in J')\} \cap (W' \times \ensuremath{\mathcal{A}}_{J'}) \end{align*} where $\ensuremath{\mathcal{A}}_{J'} := \prod_{j' \in J'} \ensuremath{\mathcal{A}}_{j'}$. Let $\pi: Y \to \ensuremath{\mathcal{A}}_J \times \ensuremath{\mathcal{A}}_{J'}$ be the natural projection. It suffices to show that $\pi$ is generically finite-to-one. Fix a generic $f_J := (f_j :j \in J) \in \ensuremath{\mathcal{A}}_J$ such that $h|_{\kk^I \times \{f_J\}} \not\equiv 0$. Then $Z := \dim(\pi_J^{-1}(f_J)) = k-1$. Since $|J'| = k-1$, it is straightforward to see that $Z \cap V(f_{j'} : j' \in J')$ is finite for generic $(f_{j'}: j' \in J')$, as required. \end{proof} \subsection{Proof of \cref{multiplicity-thm}} \label{multiplicity-proof-section} Due to \cref{generic-thm,strong-existence,bkk-lemma}, it suffices to prove the following: \begin{thm}\label{multiplicity-thm'} Let $\mathscr{G} := (\Gamma_1, \ldots, \Gamma_n)$ be a collection of $n$ diagrams in $\ensuremath{\mathbb{R}}^n$. Let $f_1, \ldots, f_n$ be strongly $\mathscr{G}$-non-degenerate. Then \begin{align} \multzero{f_1}{f_n} &:= \sum_{I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}} \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \times \multstar{\Gamma^{I}_1, \Gamma^{I}_{j'_1}}{\Gamma^{I}_{j'_{k-1}}} \label{mult-formula'} \end{align} where the right hand side of \eqref{mult-formula'} is as in \cref{multiplicity-thm}. \end{thm} \begin{proof} We prove \cref{multiplicity-thm'} by induction on $n$. It suffices to treat the case that $\multzero{\Gamma_1}{\Gamma_n} < \infty$. The theorem is true for $n = 1$ (see \cref{formula-convention}), so assume it is true for all dimensions smaller than $n$. It is straightforward to check that if $0 \in \Gamma_j$ for some $j$, $1 \leq j \leq n$, then both sides of \eqref{mult-formula'} are zero. So assume $0 \not\in \Gamma_j$ for all $j$, $1 \leq j \leq n$.\\ Let $\ensuremath{\mathcal{Z}}_0$ be the union of irreducible components of $V(f_2, \ldots, f_n)$ that contain the origin. Since $0 < \multzero{f_1}{f_n} < \infty$ (combine \cref{generic-thm,bkk-lemma}) it follows that $\dim(\ensuremath{\mathcal{Z}}_0) = 1$. For each $I \subseteq [n]$, let $\scrZ_0^I := \overline{\ensuremath{\mathcal{Z}}_0 \cap (\kk^*)^I}$, and let $\scrB^I_0$ be the (possibly empty) set of branches of $\scrZ_0^I$. For each $\nu \in \scrV^I_0$, let $\scrB^I_{0,\nu}$ be the set of all branches $B \in \scrB^I_0$ such that $\nu^I_B = \nu$. Then \begin{align} \multzero{f_1}{f_n} &= \sum_{I \subseteq [n]} \sum_{\nu \in \scrV^I_0} \sum_{B \in \scrB^I_{0,\nu}} \ord_B(f_2, \ldots, f_n)\nu_B(f_1) \label{all-Bzeroinu-sum} \end{align} where $\ord_B(f_2, \ldots, f_n)$ is the intersection multiplicity of $f_2, \ldots, f_n$ along $B$. We now compute the right hand side of \eqref{all-Bzeroinu-sum}. \begin{lemma} Let $I$ be a non-empty subset of $[n]$ and $\nu \in \scrV^I_0$. \begin{enumerate} \item \label{Bzeroinu-non-empty} If $\scrB^I_{0,\nu} \neq \emptyset$ then $I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}$. \item \label{I_1} Assume $I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}$. Then \begin{align} &\nu_B(f_1) = \nu(f_1|_{\kk^I})\ \text{for each $B \in \scrB^I_{0,\nu}$, and} \label{nu_B-f_1} \\ &\sum_{B \in \scrV^I_0} \ord_B(f_{j'_1}|_{\kk^I}, \ldots, f_{j'_{k-1}}|_{\kk^I}) = \mv(\In_\nu(\Gamma^I_{j'_1}), \ldots, \In_\nu(\Gamma^I_{j'_{k-1}}) ) \label{Bzeroinu-sum} \end{align} where $j'_1, \ldots, j'_{k-1}$ are defined as in \cref{multiplicity-thm}. \end{enumerate} \end{lemma} \begin{proof} At first we prove assertion \ref{Bzeroinu-non-empty}. Let \begin{align*} J &:= \{j: 2 \leq j \leq n,\ f_j|_{\kk^I} \equiv 0\} = \{j: 2 \leq j \leq n,\ \Gamma^I_j = \emptyset \} \\ J' &:= \{2, \ldots, n\} \setminus J \end{align*} Pick $B \in \scrB^I_{0,\nu}$. Since $\In(B) \in (\kk^*)^I \cap V(\In_\nu(f_2|_{\kk^I}), \ldots, \In_\nu(f_n|_{\kk^I}))$ (\cref{branch-lemma}), property \ref{strongly-non-degenerate} of \cref{strongly-generic-defn} implies that \begin{align*} |I| - 1 \geq \dim(\In_\nu(\sum_{j \in J'} \np(f_j|_{\kk^I}))) \geq |J'| = n - 1 - |J|, \end{align*} so that $|J| \geq n-|I|$. On the other hand since $\mathscr{G}$ is $I$-isolated at the origin (\cref{mult-support-solution}), it follows that $|J|\leq n- |I|$. Therefor $|J| = n - |I|$. The $I$-isolation at the origin then also implies that $f_1|_{\kk^I} \not\equiv 0$, so that $I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}$, as required.\\ Now we prove assertion \ref{I_1}. We may assume $J' = \{2, \ldots, k\}$. Let $X_\nu$ be the $\nu$-weighted blow up $\sigma_\nu: X_\nu \to \kk^I$ at the origin and let $U := (\kk^*)^I \cup E^*_\nu \subseteq X_\nu$ (\cref{blow-up-defn}). Then $U$ is open in $X_\nu$. For each $j$, $1 \leq j \leq k$, let $D_j$ be the divisor on $\kk^I$ determined by $f_j$, and $D'_j$ be the strict transform of $D_j$ on $X_\nu$. The complete BKK non-degeneracy of $f_2, \ldots, f_n$ implies that there are only finitely many points in $E_\nu$ common to $D'_2, \ldots, D'_k$, and all these points are in $E^*_\nu$; moreover, these are precisely the points on the strict transforms of the branches in $\scrB^I_{0,\nu}$. Finally, by the BKK non-degeneracy at the origin, none of the points in $E^*_\nu \cap D'_2 \cdots \cap D'_k$ belong to $D'_1$; in particular this implies identity \eqref{nu_B-f_1}. It follows that \begin{align*} &\sum_{B \in \scrV^I_0} \nu_B(f_1)\ord_B(f_{j'_1}|_{\kk^I}, \ldots, f_{j'_{k-1}}|_{\kk^I}) \\ &\qquad\qquad = (\sigma_{\nu}^*(D_1)|_U, D'_2|_U, \ldots, D'_n|_U) = \nu(f_1|_{\kk^I})(E^*_\nu + D'_1|_U, D'_2|_U, \ldots, D'_n|_U) \\ &\qquad\qquad = \nu(f_1|_{\kk^I})(E^*_\nu, D'_2|_U, \ldots, D'_n|_U) = \nu(f_1|_{\kk^I})( D'_2|_{E^*_\nu}, \ldots, D'_n|_{E^*_\nu}) \\ &\qquad\qquad = \nu(f_1|_{\kk^I}) \mult{\In_\nu(f_2|_{\kk^I})}{\In_\nu(f_k|_{\kk^I})}_{(\kk^*)^I} = \nu(f_1|_{\kk^I})\mv(\In_\nu(\Gamma^I_2, \ldots, \In_\nu(\Gamma^I_k) ), \end{align*} where the last equality follows from Bernstein's theorem. This proves identity \eqref{Bzeroinu-sum}. \end{proof} Now let $B \in \scrB^I_{0,\nu}$, where $I \in \ensuremath{\mathcal{I}}_{\mathscr{G},1}$ and $\nu \in \scrV^I_0$. Let $j_1, \ldots, j_{n-k}$ (resp.\ $j'_1, \ldots, j'_{k-1}$) be the elements of $J_I$ (resp.\ $\{2, \ldots, n\}\setminus J_I$). Let $D_j$ be the Cartier divisor on $\ensuremath{\mathbb{K}}^n$ defined by $f_j$, $2 \leq j \leq n$. Since $D_j$'s intersect properly at $B$, $\ord_B(f_2, \ldots, f_n)$ is the coefficient of (the closure of) $B$ in the (proper) intersection product $D_2 \cdot D_3 \cdots D_n$ = $(D_{j_1} \cdots D_{j_{n-k}})(D_{j'_1} \cdots D_{j'_{k-1}})$. Note that \begin{itemize} \item $\kk^I$ is an irreducible component of $D_{j_1} \cdots D_{j_{n-k}}$, and \item property \ref{intersectionally-non-degenerate} of strong non-degeneracy ensures that $B$ is not contained in any other irreducible component of $D_{j_1} \cdots D_{j_{n-k}}$. \end{itemize} It follows that \begin{align} \ord_B(f_2, \ldots, f_n) = \ord_{\kk^I}(f_{j_1}, \ldots, f_{j_{n-k}}) \ord_B(f_{j'_1}|_{\kk^I}, \ldots, f_{j'_{k-1}}|_{\kk^I}) \end{align} Let $i'_1, \ldots, i'_{n-k}$ be the elements of $I' = [n]\setminus I$. As in property \ref{recursively-non-degenerate} of strong non-degeneracy, let $f_{I',j_1}, \ldots, f_{I',j_{n-k}}$ be the polynomials in $(x_{i'}: i' \in I')$ formed by specializing $(x_i: i \in I)$ to $(1, \ldots, 1)$. Then $f_{I',j_1}, \ldots, f_{I',j_{n-k}}$ are completely BKK non-degenerate. \Cref{bkk-deformation} implies that complete BKK non-degeneracy is preserved by specialization of $(x_i: i \in I)$ in the expressions for $f_{j_1}, \ldots, f_{j_{n-k}}$ to generic $\xi \in (\ensuremath{\mathbb{K}}^*)^k$. \Cref{generic-length} and the inductive hypothesis then imply that \begin{align} \ord_{\kk^I}(f_{j_1}, \ldots, f_{j_{n-k}}) = \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \label{ord-Ki} \end{align} \Cref{multiplicity-thm'} follows from combining identities \eqref{all-Bzeroinu-sum}--\eqref{ord-Ki}. \end{proof} \section{Proof of the extended BKK bound} \label{bkk-proof-section} In this section we prove \cref{bkk-thm} following the same approach of the proof of \cref{multiplicity-thm} in \cref{multiplicity-section}. Throughout this section $\mathscr{P}:= (\ensuremath{\mathcal{P}}_1, \ldots, \ensuremath{\mathcal{P}}_n)$ is a collection of $n$ convex integral polytopes in $\ensuremath{\mathbb{R}}^n$. \begin{defn} \label{strongly-bkk-non-degenerate} Given $f_1, \ldots, f_n \in \ensuremath{\mathbb{K}}[x_1, \ldots, x_n]$, we say that they are {\em strongly $\mathscr{P}$-non-degenerate} iff they are $\mathscr{P}$-admissible and satisfy properties \ref{strongly-non-degenerate}, \ref{intersectionally-non-degenerate} and \ref{recursively-non-degenerate} of \cref{strongly-generic-defn}. \end{defn} \Cref{strong-existence} shows that generic systems of polynomials with support $\mathscr{P}$ are strongly $\mathscr{P}$-non-degenerate. We prove \cref{bkk-thm} via the same arguments as in the proof of \cref{multiplicity-thm}. In particular, we show the following: if $f_1, \ldots, f_n$ are strongly $\mathscr{P}$-non-degenerate, then \begin{align} \multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} &= \sum_{I \in \scrI_{\mscrS,\mscrP,1}} \multzero{\pi_{I'}(\Gamma_{j_1})}{\pi_{I'}(\Gamma_{j_{n-k}})} \times \multstarinftySS{\ensuremath{\mathcal{P}}^{I}_1, \ensuremath{\mathcal{P}}^{I}_{j'_1}}{\ensuremath{\mathcal{P}}^{I}_{j'_{k-1}}}{(\overbar \mathscr{S} \cup \scrE_\mscrP)^I} \label{bkk-formula'} \end{align} where the right hand side of \eqref{bkk-formula'} is as in \cref{bkk-thm}. \\ Let $\mathscr{S}' := \overbar \mathscr{S} \cup \scrE_\mscrP$. Define \begin{align*} \scrI_{\mscrS,\mscrP} &:= \{I \subseteq [n]: I \not\in \overbar \mathscr{S} \cup \scrE_\mscrP \cup \{\emptyset\},\ |\scrN^{I}_{\mscrP}| = |I|,\ \text{$\mathscr{P}$ is $(\kk^*)^I$-non-trivial} \} \end{align*} where $\scrN^{I}_{\mscrP}$ is as in \eqref{NiP}, i.e.\ $\scrI_{\mscrS,\mscrP,1}$ is simply the subset of $\scrI_{\mscrS,\mscrP}$ consisting of subsets of $[n]$ containing $1$. \Cref{herrero}, \cref{Pisolated-cor} and assertion \ref{bkk-trivial} of \cref{bkk-lemma} imply that \begin{prooflist} \item \label{isolated-observation-1} all roots of $f_1, \ldots, f_n$ in $X^{\scrI_{\mscrS,\mscrP}} := \bigcup_{I \in \scrI_{\mscrS,\mscrP}}(\kk^*)^I$ are isolated, \item \label{isolated-observation-2} and in turn each isolated root of $f_1, \ldots, f_n$ is contained in $X^{\scrI_{\mscrS,\mscrP}}$. \end{prooflist} Let $\ensuremath{\mathcal{Z}}$ be the union of irreducible components of $V(f_2, \ldots, f_n)$ on $X^{\scrI_{\mscrS,\mscrP}}$. Pick an irreducible component $Z$ of $\ensuremath{\mathcal{Z}}$ and the the smallest $I \in \scrI_{\mscrS,\mscrP}$ such that $(\kk^*)^I$ contains $Z$. \begin{claim}\label{dimension-claim} $\dim(Z) = 1$ and $I \in \scrI_{\mscrS,\mscrP,1} $. \end{claim} \begin{proof} It follows from definition of $\ensuremath{\mathcal{I}}$ that $|\scrN^{I}_{\mscrP}| = |I|$. If $\scrN^{I}_{\mscrP} \subseteq \{2, \ldots, n\}$, then \cref{bkk-lemma} implies that $V(f_2, \ldots, f_n) \cap (\kk^*)^I$ is finite. Since $\dim(Z) \geq 1$, this contradicts the construction of $I$. It follows that $1 \in \scrN^{I}_{\mscrP}$, so that $I \in \scrI_{\mscrS,\mscrP,1} $. This implies that $|\scrN^{I}_{\mscrP}\setminus \{1\}| = |I| - 1$. The same arguments as in the proof of \cref{bkk-dimension} then show that $\dim(V(f_j: 2 \leq j \leq n) \cap (\kk^*)^I) = 1$. \end{proof} For each $I \in \scrI_{\mscrS,\mscrP,1}$, let $\scrZ^I := \overline{\ensuremath{\mathcal{Z}} \cap (\kk^*)^I}$, and let $\scrB^I$ be the set of branches of $\scrZ^I$. For each $\nu \in \scrV^I_{\mscrS'}$ (resp.\ $\omega \in \Omega^{I}_\infty$), let $\scrB^I_{\nu}$ (resp.\ $\scrB^I_{\omega}$) be the set of all branches $B \in \scrB^I$ such that $\nu^I_B = \nu$ (resp.\ $\nu^I_B = -\omega$). Observations \ref{isolated-observation-1}, \ref{isolated-observation-2} and \cref{dimension-claim} imply that \begin{align*} \multiso{f_1}{f_n}_{\kk^n_{\overbar \mscrS}} &= -\sum_{I \in \scrI_{\mscrS,\mscrP,1}} \left( \sum_{\omega \in \Omega^{I}_\infty} \sum_{B \in \scrB^I_{\omega}} \ord_B(f_2, \ldots, f_n) +\sum_{\nu \in \scrV^I_{\mscrS'}} \sum_{B \in \scrB^I_{\nu}} \ord_B(f_2, \ldots, f_n) \right) \nu_B(f_1) \end{align*} Identity \eqref{bkk-formula'} now follows exactly in the same way as the proof of \eqref{mult-formula'} from \eqref{all-Bzeroinu-sum}. \qed
f879e86cb52adf8549de04a2e4701b97210970e1
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Derivation of the effective Hamiltonian}\label{sec:adiabatic} In this section we illustrate the derivation of the OPO Hamiltonian, as given by Eq.~(4) in the main text. We also sketch the extension to the case of an all-optical switch with two input beams. We first describe the adiabatic elimination of the electronically excited states in detail. We assume that the detuning $\Delta=\omega_e-\omega_L$ is large compared with the relevant energy scales for the electronic ground states ($\ket{g^{(i)}}$, $\ket{v^{(i)}}$), and that the interaction terms involving electronic excitations in Eq.~(2) in the main text are perturbative. In the following, we denote these terms as $\hat{V} = \hat{V}_+ + \hat{V}_+^\dagger$, where $\hat{V}_+$ contains all the terms \emph{creating} electronic excitations, The occupation of the electronically excited states $\ket{e^{(i)}}$ with free evolution Hamiltonian $\hat{H}_e=\omega_e \sum_{i=1}^N\hat{\sigma}_{ee}^{(i)}$ is hence vanishingly small. In particular, this assumption enables to avoid the coupling of different electronically excited states, simplifying the following treatment. In the adiabatic elimination procedure \cite{S-Feshbach1958}, the density matrix equations are solved by assuming a slow evolution of the lowest-lying states $\ket{g^{(i)}}$, $\ket{v^{(i)}}$ and the optical modes $\hat{a}_S$, $\hat{a}_L$, determined by $\hat{H}_g = \hat{H}_s + \omega_S \hat{n}_S + \omega_{L} \hat{n}_L$, with $\hat{H}_s$ given in Eq.~(1) in the main text. We here ignore contributions originating from the incoherent dynamics within the electronic excited manifold, which could be introduced by means of effective Lindblad terms \cite{S-Reiter2012}, but are negligible for large detuning. To perform the adiabatic approximation, we \emph{i)} apply the \emph{rotating-frame} transformation $\hat{U}=e^{-i(\omega_{S}\hat{n}_{S}+\omega_{L}\hat{n}_{L})t}$ and \emph{ii)} we work in the eigenbasis of $\hat{H}_s$. The resulting Hamiltonian is \begin{equation} \hat{H}' = \hat{H}_{s} - \frac{1}{2} \hat{V}_{+}^{\prime\dagger}(t) \sum_{f,l}\frac{\omega_e}{\omega_e-\omega_f-\omega_l}\hat{H}_e^{-1} \hat{v}_+^{(l,f)}e^{i\omega_ft}. \label{S-eq:effective} \end{equation} Here, $\hat{V}'_+(t)$ has been expanded in terms of its frequency components $f\in(L,S)$, as well as the system eigenstates $l\in(+,-,\{d\})$ that it couples to, giving \begin{multline} \hat{V}'_{+}(t) = \sum_{f,l}\hat{v}_+^{(f,l)}e^{i\omega_ft} = \sum_{i=1}^{N} \Bigg[ g_{L}\hat{a}_{L}\hat{\sigma}_{Ge}^{\dagger(i)}e^{i\omega_{L}t} \\ + g_{S}\hat{a}_{S} \left( \sum_{\eta=\{\pm\}} \frac{\hat{\sigma}_{\eta e}^{\dagger(i)}}{\sqrt{2N}} + \sum_{d}u_{id}\hat{\sigma}_{de}^{\dagger(i)}\right) e^{i\omega_{S}t}\Bigg]. \label{S-eq:int_eig} \end{multline} Here, $\hat{\sigma}_{le}^{(i)} = \ket{l}\bra{e^{i}}$, and the coefficients appearing between parenthesis follow from the eigenstate expansions $\ket{v^{(i)}}=(2N)^{-\frac{1}{2}}\sum_{\eta=\{\pm\}}\ket{\eta}+\sum_d u_{id}\ket{d}$, where $u_{id}$ is the overlap matrix element between the $i$th vibrational excitation and dark state $d$. These coefficients fulfill $\sum_{i=1}^N u_{id}=0$ and are further constrained by the orthogonality relation $\sum_{i=1}^N u_{di}u_{id'}=\delta_{dd'}$. After going back to the nonrotating frame, the resulting effective interaction reads \begin{equation}\label{S-eq:int_OPO} \hat{H}_{\mathrm{int}}^{\mathrm{eff}} = -\sum_{\eta=\{\pm\}} g_{\mathrm{eff}}^{\eta} (\hat{a}_{L}\hat{a}_{S}^{\dagger}\hat{\sigma}_{G\eta}^{\dagger} + \mathrm{H.c.}), \end{equation} with $g_{\mathrm{eff}}^{\eta} = \frac{g_{S}g_{L}}{2} \sqrt{\frac{N}{2}} \left[(\omega_{e}-(\omega_{\eta}+\omega_{S}))^{-1} + \Delta^{-1}\right]$. Note here that the contribution of the dark states is identically zero in the effective dynamics. This is due to the fact that we assumed perfect overlap between the involved modes, i.e., we took $g$, $g_S$, and $g_L$ to be constant for all involved molecules. Relaxing this condition would give an additional overlap prefactor $g_{\mathrm{eff}}^{\eta}\to \mathcal{S}g_{\mathrm{eff}}^{\eta}$, with $\mathcal{S}\propto \sum_i u_{i\eta} g_{S,i} g_{L,i}^*$, and also give nonzero coupling to the dark states, but would not otherwise change the results presented in the main text. In addition to the effective interaction, we obtain (nonlinear) energy shifts, given by \begin{equation}\label{S-eq:e_shifts} \hat{H}_{\mathrm{shift}}^{\mathrm{eff}} = -\frac{g_L^2N}{\Delta}\hat{n}_L\hat{\sigma}_{GG} -\sum_l \frac{g_{S}^{2} \hat{n}_S \hat{\sigma}_{ll}} {\omega_{e}-(\omega_{l}+\omega_{S})}. \end{equation} Under the assumption that the output modes are not significantly populated ($\hat{\sigma}_{GG}\approx1$, $\hat{n}_S,\hat{\sigma}_{\eta\eta}\ll1$), the first term just gives a constant energy shift (which we assume to be included into $\omega_L$), while the second term can be neglected. This is a good approximation for typical system parameters even when a large number of output photons is generated, due to the relatively short lifetime of the polaritons. Under this approximation, the vibro-polaritons are well-modeled as bosons $[\hat{\sigma}_{G\eta},\hat{\sigma}_{G\eta'}^{\dagger}]\simeq\delta_{\eta\eta'}$, and the trilinear interaction \autoref{S-eq:OPO_eff} corresponds to a nondegenerate OPO under the identifications $\hat{a}_S\rightarrow$ signal, $\hat{\sigma}_{Gp}\rightarrow$ idler. For a laser frequency chosen such that Raman scattering to one of the polaritonic modes is resonant with the cavity mode $S$ in the optical, $\omega_L=\omega_S+\omega_p$, with $p\in\{\pm\}$, there are rapidly oscillating terms in \autoref{S-eq:int_OPO} in the interaction picture with regards to $\hat{H}_g$. Averaging over a time sufficiently big compared to $\tau_{\mathrm{coh}}\sim\Omega_R^{-1}$, these contributions, which correspond to the coupling of the laser field with the detuned polariton, can be neglected under a \emph{second} rotating wave approximation, giving \begin{align} \hat{H}_{\mathrm{eff}}\simeq\omega_L\hat{n}_L+\omega_S\hat{n}_S+\omega_{p}\hat{\sigma}_{pp}-g_{\mathrm{eff}}^{p}\left(\hat{a}_{L}\hat{a}_{S}^{\dagger}\hat{\sigma}_{Gp}^{\dagger}+\mathrm{H.c.}\right).\label{S-eq:OPO_eff} \end{align} We next sketch the derivation of the effective Hamiltonian under pumping of multiple input modes, as given by Eq.~(11) in the main text. The pumping Hamiltonian is then \begin{equation} \hat{H}^{(2)}_{d} = \sum_{\eta=\{\pm\}} \sqrt{\kappa_{L\eta}}\Phi_{\mathrm{in}}^{\eta}(\hat{a}_{L\eta}e^{-i\omega_{L\eta}t}+\hat{a}_{L\eta}^{\dagger}e^{i\omega_{L\eta}t}), \end{equation} while the pump-system interaction is given by \begin{equation} \hat{V}_+ = \sum_{i=1}^{N}\big(g_{S}\hat{a}_S\hat{\sigma}_{v_ie_i}^{\dagger} + \sum_{\eta=\{\pm\}}g_{L\eta}\hat{a}_{L\eta} \hat{\sigma}_{g_ie_i}^{\dagger}\big). \end{equation} The frequencies $\omega_{L\pm}$ are chosen to satisfy the resonance conditions for both polaritons, $\omega_{L\pm}=\omega_{S}+\omega_{\pm}$. The adiabatic elimination of the electronic states proceeds analogously to the single-pump case, giving the effective interaction Hamiltonian \begin{equation}\label{S-eq:int_OPO_2} \hat{H}_{\mathrm{int}}^{\mathrm{eff}(2)} = -\sum_{\eta,\eta'=\{\pm\}}g_{\mathrm{eff}}^{\eta,\eta'} (\hat{a}_{L\eta'}\hat{a}_{S}^{\dagger}\hat{\sigma}_{G\eta}^{\dagger} + \mathrm{H.c.}), \end{equation} where the effective coupling constant of the pump field $\eta$ with the polariton $\eta'$ is (using $\Delta_\pm=\omega_e-\omega_{L\pm}$) \begin{equation} g_{\mathrm{eff}}^{\eta,\eta'} = \frac{g_{S}g_{L\eta}}{2} \sqrt{\frac{N}{2}} \left[\frac{1}{\omega_{e}-(\omega_{\eta'}+\omega_{S})}+\frac{1}{\Delta_\eta}\right]. \end{equation} The off-diagonal terms $\eta\not=\eta'$ can be neglected under the same \emph{second} RWA we invoked for the OPO case. In addition to the effective interaction, we again obtain extra nonlinear terms, given by \begin{multline}\label{S-eq:H2_extra} \hat{H}_{\mathrm{extra}}^{\mathrm{eff}(2)} = -\sum_{\eta,\eta'} \Lambda_{\mathrm{eff}}^{\eta,\eta'} \hat{\sigma}_{GG} (\hat{a}_{L\eta}^{\dagger}\hat{a}_{L\eta'} + \mathrm{H.c.})\\ - \hat{n}_S\sum_{l} \frac{g_{S}^{2}\hat{\sigma}_{ll}}{\omega_{e}-(\omega_{l}+\omega_{S})}, \end{multline} which in addition to the energy shifts already seen in the single-pump OPO case also contains an extra crossed term coupling the two pump fields. The nonlinear terms \autoref{S-eq:H2_extra} can again be neglected under the low-occupation assumption and the second RWA. Finally, we thus obtain \begin{multline} \hat{H}^{(2)}_{\mathrm{eff}}\simeq \omega_S \hat{n}_S + \sum_{\eta=\{\pm\}} \bigg[ \omega_{L\eta}\hat{n}_{L\eta} + \omega_{\eta}\hat{\sigma}_{\eta\eta}\\ - g^{\eta}_{\mathrm{eff}} (\hat{\sigma}_{G\eta}\hat{a}_S\hat{a}_{L\eta}^{\dagger} + \mathrm{H.c.})\bigg], \label{S-eq:H_2} \end{multline} where we used that $g_{\mathrm{eff}}^{\eta,\eta} = g_{\mathrm{eff}}^{\eta}$. \section{Mean-field steady-state solutions} We here discuss the general steady-state solution for the optical-switch setup with multiple pump beams. For clarity, we first recall the well-known Manley-Rowe relations for beam fluxes and powers~\cite{S-Yariv1966} in the single-pump case, which follow straightforwardly from the semiclassical equations Eq.~(7) in the main text. They connect the fluxes of emitted photons $P_i/\omega_i=\gamma_i|\alpha_i|^2$ in the different modes, with the simple relation \begin{align} \frac{P_S}{\omega_S}&=\frac{P_p}{\omega_p}=\frac{P_L^{\mathrm{th}}}{\omega_L}(f-1), & f&=\frac{\Phi_{\mathrm{in}}}{\Phi_{\mathrm{th}}}. \label{S-eq:conservation} \end{align} This explicitly shows that Raman scattering converts each of the incoming pump photons into a Stokes photon/polariton pair. From \autoref{S-eq:conservation} and the resonance condition $\omega_L=\omega_S+\omega_p$, we obtain the power relation $P_{\mathrm{in}}=P_S +P_p+P_L^{\mathrm{th}}$. This expresses the fact that the input power $P_{\mathrm{in}}= \omega_L\Phi_{\mathrm{in}}\Phi_{\mathrm{th}}$ is shared among the three modes, with a maximum clamped power for the $L$ mode at threshold equal to $\Phi_{\mathrm{th}}^2$. From the analogue steady-state relations within the two-pump scenario, a set of generalized Manley-Rowe relations accounting for the exchange of energy between the modes participating in the scattering holds: \begin{align}\label{S-eq:power_rels_2} \frac{P_{S}}{\omega_{S}} &= \sum_{\eta=\{\pm\}} \frac{P_{\eta}}{\omega_{\eta}}, &\frac{P_{\pm}}{\omega_{\pm}} &= \frac{P_{\mathrm{in}}^{\pm}-P_{L\pm}}{\omega_{L\pm}}, \end{align} where the input power in each of the pump modes $L\pm$ is $P_{\mathrm{in}}^{\pm}=\omega_{L\pm}\sqrt{\kappa_{L\pm}}\Phi^{\pm}_{\mathrm{in}}|\alpha_{L\pm}|$. Employing the resonance conditions for the two pumps, we obtain the global power relation $\sum_{\eta=\{\pm\}}P_{\mathrm{in}}^\eta = P_S + \sum_{\eta=\{\pm\}} \left(P_\eta+P_{L\eta}\right)$, a direct generalization of \autoref{S-eq:conservation}. We now proceed to solve the steady-state equations, Eq.~(13) in the main text, under pumping of both modes ($\Phi_\mathrm{in}^\pm>0$). The following relations between the $L\pm$ amplitudes hold, \begin{equation} \frac{\kappa_{L+}|\alpha_{L+}|^{2}}{\Phi_{\mathrm{th}}^{+2}}+\frac{\kappa_{L-}|\alpha_{L-}|^{2}}{\Phi_{\mathrm{th}}^{-2}}=1,\label{S-eq:general_eqs} \end{equation} expressing the fact that the \emph{global} pump amplitude becomes clamped above the threshold due to its connection to a common Stokes mode (with $\Phi_{\mathrm{th}}^{\pm} = \sqrt{\kappa_{L\pm}\kappa_{S}\Gamma_{\pm}}/g^\pm_{\mathrm{eff}}$). As we will see in the following, this is not the case for each of the pumping amplitudes individually. The relation \autoref{S-eq:general_eqs} suggests we can define \begin{align} |\alpha_{L+}|=\frac{\Phi_{\mathrm{th}}^{+}}{\sqrt{\kappa_{L+}}}\sin\Theta,&& |\alpha_{L-}|=\frac{\Phi_{\mathrm{th}}^{-}}{\sqrt{\kappa_{L-}}}\cos\Theta, \end{align} with mixing angle $\Theta$ in the range $\Theta\in(0,\pi/2)$, such that $|\alpha_{L\pm}|>0$. Inserting this into the steady state-equations leads to a quartic equation for $t=\tan(\Theta/2)$, given by \begin{equation}\label{S-eq:quartic} t^{4}+2(\alpha-\beta)t^{3}+2(\alpha+\beta)t-1=0, \end{equation} where $\alpha=\Phi_{\mathrm{in}}^{-}\Phi_{\mathrm{th}}^{-}/(\Phi_{\mathrm{in}}^{+}\Phi_{\mathrm{th}}^{+})$, and $\beta=[(\Phi_{\mathrm{th}}^{+})^{2}-(\Phi_{\mathrm{th}}^{-})^{2}]/(\Phi_{\mathrm{in}}^{+}\Phi_{\mathrm{th}}^{+})$. We have checked that this equation has only one physical solution $0<t<1$ for arbitrary values of $\alpha>0$) and $\beta$. The analytical form of the solution of \autoref{S-eq:quartic} is very lengthy and we thus omit it in the following. However, in the degenerate case with equal thresholds, we get $\beta=0$ and the equation can be factorized as $\left(t^2+1\right) \left(t^2+2 \alpha t-1\right)=0$, leading to the single physical solution analyzed in the main text, $t=\sqrt{1+\alpha ^2}-\alpha$, with $\alpha=f_-/f_+$. \section{Stability of the mean-field solutions} \allowdisplaybreaks We here analyze the stability of the semiclassical steady-state solutions of Eq.~(12) in the main text. Collecting these solutions in the vector $\boldsymbol{v}^{\infty}=(\alpha_S,\psi_+,\alpha_{L+},\psi_-,\alpha_{L-},\mathrm{H.c.})$, inserting the linearized solution $\boldsymbol{v}(t)=\boldsymbol{v}^{\infty}+\delta\boldsymbol{v}(t)$ in Eq.~(12) and keeping terms $O(\delta\boldsymbol{v})$, we obtain the time evolution of the fluctuations, $\partial_{t}\delta\boldsymbol{v}(t) = \boldsymbol{\mathcal{M}}\delta\boldsymbol{v}(t)$. The stability matrix is \begin{equation} \boldsymbol{\mathcal{M}} = \left(\begin{array}{c|cc} -\kappa_{S}\mathbb{1}_{2\times2} & \boldsymbol{v}_{+} & \boldsymbol{v}_{-}\\ \hline \boldsymbol{u}_{+}^T & -\boldsymbol{\mathcal{P}}_{+} & 0\\ \boldsymbol{u}_{-}^T & 0 & -\boldsymbol{\mathcal{P}}_{-} \end{array}\right), \end{equation} where the submatrices are \begin{subequations} \begin{equation} \boldsymbol{v}_\pm = ig_{\mathrm{eff}}^{\pm} \begin{pmatrix} 0 & \alpha_{L\pm} & \psi_\pm^{*} & 0\\ -\alpha_{L\pm}^{*} & 0 & 0 & -\psi_\pm \end{pmatrix}, \end{equation} \begin{align} \boldsymbol{u}_\pm &= ig_{\mathrm{eff}}^{\pm} \begin{pmatrix} \alpha_{L\pm} & 0 & 0 & -\psi_\pm^{*}\\ 0 & -\alpha_{L\pm}^{*} & \psi_\pm & 0 \end{pmatrix},\\ \boldsymbol{\mathcal{P}}_{\pm} &= \begin{pmatrix} \Gamma_{\pm} & 0 & -ig_{\mathrm{eff}}^{\pm}\alpha_{S}^{*} & 0\\ 0 & \Gamma_{\pm} & 0 & ig_{\mathrm{eff}}^{\pm}\alpha_{S}\\ -ig_{\mathrm{eff}}^{\pm}\alpha_{S} & 0 & \kappa_{L\pm} & 0\\ 0 & ig_{\mathrm{eff}}^{\pm}\alpha_{S}^{*} & 0 & \kappa_{L\pm} \end{pmatrix}. \end{align} \end{subequations} The fluctuations $\delta\boldsymbol{v}(t)$ will grow exponentially in time if the real part of any eigenvalue of $\boldsymbol{\mathcal{M}}$ is positive. This can be tested via the Routh-Huwirtz criterion \cite{S-Gradshteyn2014}, which provides necessary and sufficient conditions for the roots of the characteristic polynomial $\det\big({\boldsymbol{\mathcal{M}}-\lambda\mathbb{1}_{10\times10}}\big)$ to have negative real part, without explicit knowledge of their values. Applying this criterion proves that the all-optical switch solutions pictured in the main text are stable. From this, the stability of the OPO solutions under single-mode driving follows automatically. \input{suppl_S.bbl} \end{document}
405f8e8e8f25079c7e869cdb940dc88d0933e1e5
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The noise assessment in physical measurements time series is an important measure of its statistical characteristics and overall quality. Among the most effective approaches to analyzing measurement noise (scatter) is Allan variance (AVAR), which was originally introduced to estimate the frequency standards instability \cite{Allan1966}. Later, AVAR has proved to be a powerful statistical tool for time series analysis, particularly, for the analysis of geodetic and astronomical observations. AVAR has been used for quality assessment and improvement of the celestial reference frame (CRF) \cite{Feissel2000a,Gontier2001,Feissel2003a,Sokolova2007,Malkin2008j,LeBail2010a,LeBail2014a,Malkin2013b,Malkin2015b}, the time series analysis of station position and baseline length \cite{Malkin2001n,Roberts2002,LeBail2006,Feissel2007,LeBail2007,Gorshkov2012b,Malkin2013b,Khelifa2014}, and studies on the Earth rotation and geodynamics \cite{Feissel1980,Gambis2002,Feissel2006a,LeBail2012,Malkin2013b,Bizouard2014a}. AVAR estimates of noise characteristics have important advantages over classical variance estimates such as standard deviation (STD) and weighted root-mean-square (WRMS) residual. The latter cannot distinguish the different significant types of noise, which is important in several astro-geodetic tasks. Another advantage of AVAR is that it is practically independent of the long-term systematic components in the investigated time series. AVAR can also be used to investigate the spectral characteristics of the signal \cite{Allan1981,Allan1987} that is actively used for analysis of astrometric and geodetic data \cite{Feissel2000a,Feissel2003a,Feissel2006a,Feissel2007}. However, the application of original AVAR to the time series analysis of astro-geodetic measurements may not yield satisfactory results. Unlike clock comparison, geodetic and astrometric measurements mostly consist of data points with unequal uncertainties. This requires a proper weighting of the measurements during the data analysis. Moreover, one often deals with multi-dimensional quantities in geodesy and astronomy. For example, the station coordinates $X$, $Y$, and $Z$ form 3D vector of a geocentric station position (although this example is more complicated because the vertical and horizontal station displacements caused by the geophysical reasons may have different statistical characteristics, including AVAR estimates, see \cite{Malkin2001n,Malkin2013c} and references therein). The coordinates of a celestial object, right ascension and declination, also form a 2D position vector. To analyze such data types, AVAR modifications were proposed in \cite{Malkin2008j}, including weighted AVAR (WAVAR), multi-dimensional AVAR (MAVAR), and weighted multi-dimensional AVAR (WMAVAR). These modifications should be distinguished from the classical modified AVAR introduced in \cite{Allan1981}. The rest of the paper is organized as follows. Section~\ref{sect:overview} introduces AVAR and its modification, and gives several practical illustrations of their basic features. In Section~\ref{sect:results}, a brief overview is provided of the works that employ AVAR in geodesy and astrometry, and basic results obtained with the AVAR technique are presented. Additional details and discussion on the use of AVAR in space geodesy and astrometry can be found in \cite{LeBail2004t,Malkin2011c,Malkin2013b}. \section{Overview of AVAR and its modifications} \label{sect:overview} The classical time-domain AVAR applied to the time series $y_i, i=1, \dots, n$ is given by \cite{Allan1966} \begin{equation} AVAR = \frac{1}{2(n-1)}\sum_{i=1}^{n-1}(y_i-y_{i+1})^2\,. \label{eq:AVAR} \end{equation} Allan deviation ADEV = $\sqrt{\rm{AVAR}}$ is used as a noise characteristics in many data analysis applications. Both AVAR and ADEV estimates will be used throughout the paper depending on the context. In metrology, the analyst is normally interested in computing not only the parameter under investigation but also its uncertainty, as a measure of reliability of obtained result. A method of estimating the AVAR confidence interval is proposed in \cite{Howe1981}. The method of estimating the AVAR confidence interval proposed in \cite{Howe1981} was applied in \cite{LeBail2006} for analysis of geodetic time series. The original AVAR definition supposes that all measurements (observations) have uniform precision. In most geodetic and astrometric applications, however, measurements display different level of precision. Therefore, appropriate weighting is necessary during analysis. Given the measurements $y_1,y_2,\dots,y_n$ with the associated uncertainties $s_1,s_2,\dots,s_n$ Then the WAVAR estimate can be defined to treat unequally weighted data \cite{Malkin2008j}: \begin{equation} \begin{array}{rl} & WAVAR = \frac{\displaystyle 1}{\displaystyle 2p} \displaystyle\sum_{i=1}^{n-1}p_i(y_i-y_{i+1})^2 \,, \\[2ex] & p = \displaystyle\sum_{i=1}^{n-1}{p_i}\,, \\[4ex] & p_i = (s_i^2+s_{i+1}^2)^{-1} \,, \end{array} \label{eq:WAVAR} \end{equation} where $p_i$ are the weights. Figure~\ref{fig:WAVAR_example} illustrates the difference between AVAR and WAVAR for three time series with outliers: daily estimates of the station height and two series of celestial pole offsets (CPO). The latter are the differences in $dX$ and $dY$ between the observed $X$ and $Y$ coordinates of the celestial pole from the celestial pole position defined by the International Astronomical Union IAU 2000/2006 precession-nutation model \citep{IERSConv2010} as measured by VLBI (very long baseline interferometry). Here, we consider points with abnormally large deviations from the mean and/or with abnormally large uncertainties to be outliers. Analysts usually aim to identify and reject outliers before conducting statistical analysis; however, generally this is not a simple problem. One can see from Fig.~\ref{fig:WAVAR_example} that WADEV can provide more robust estimate even in the presence of outliers that were not previously detected and eliminated. In other words, WADEV is less dependent on the quality of the outlier detection procedure used prior to data analysis. Notably, WADEV dependence on outliers is low when the outliers exhibit greater uncertainties than the uncertainties of ``good'' measurements. This is mostly the case in practice. Exceptions are possible, as shown in the bottom plot of Fig.~\ref{fig:WAVAR_example} although they are rare. \begin{figure}[ht!] \centering \includegraphics[width=\columnwidth]{wadev_dh.eps} \includegraphics[width=\columnwidth]{wadev_dx.eps} \includegraphics[width=\columnwidth]{wadev_dy.eps} \caption{Three examples of time series of unequally weighted measurements: station height (top) and celestial pole offset (middle and bottom). Two upper cases show series that contain outliers with large measurement errors; for these series the outliers substantially impact ADEV, and WADEV estimate looks more realistic. Third series contains two outliers with uncertainties similar to that of other measurements; in this case WADEV estimate has no advantage over ADEV.} \label{fig:WAVAR_example} \end{figure} It is important to keep in mind that AVAR, as well as any other statistics has limitations depending on data under analysis. Let us consider an example of time series of Z coordinate of the station position presented in Fig.~\ref{fig:hflk}, upper plot. One can see that both ADEV and WADEV give unsatisfactory results. Both estimates are affected by jumps in the time series and are too large as compared with the actual measurement noise. Moreover, WADEV gave even worse result than ADEV. The reason of such a confusing WADEV estimate is the following. As can be clearly seen in Fig.~\ref{fig:hflk}, bottom plot, the measurement errors changed with time. In particular, the smallest errors are observed in the second part of the time series, where the largest jump occurred. As a result, the measurements made around the second, largest jump were entered to the WADEV computation with larger weight. For comparison, if only the part of the time series after the last jump (epoch 2007.5) is used for analysis, ADEV = WADEV = 0.2~cm. So, ADEV, as well as most of other statistics, cannot provide a satisfactory result for non-stationary time series. In particular, if the time series under investigation has heterogeneous uncertainty, the use of dynamic AVAR (DAVAR) \cite{Galleani2009} is clearly advantageous. The DAVAR is defined as AVAR computed over a sliding window moving along the time series, and thus it allows estimating the variation of the time series noise characteristics with time. \begin{figure}[ht!] \centering \includegraphics[width=\columnwidth]{hflk_x.eps} \includegraphics[width=\columnwidth]{hflk_sx.eps} \caption{Time variation of Z coordinate of a GPS station and its uncertainty.} \label{fig:hflk} \end{figure} In astro-geodetic time series analysis, we often address multi-dimensional values, such as geocentric station coordinates $X$, $Y$, and $Z$ (3D vector) and the position of celestial objects (right ascension and declination) in the sky (2D unit vector). Further extension of the WAVAR estimator WMAVAR can facilitate the analysis of such data \cite{Malkin2008j}. Given $k$-dimensional vector of measurements $y_i=(y_i^1,y_i^2,\dots,y_i^k)$ with the $k$-dimensional vector of corresponding uncertainties $s_i=(s_i^1,s_i^2,\dots,s_i^k)$, WMAVAR considers the measurements to be the points in the $k$-dimensional space. Then, the difference between adjacent measurements is taken as the Euclidean length between these values. Thus, the $k$-dimensional WMAVAR estimate can be defined as \begin{equation} \begin{array}{rl} & WMAVAR = \frac{\displaystyle 1}{\displaystyle 2p} \displaystyle\sum_{i=1}^{n-1}p_i \, d_i^2\,, \\ & d_i = |y_i-y_{i+1}|\,, \\[1ex] & p = \displaystyle\sum_{i=1}^{n-1}{p_i}\,, \\[3ex] & p_i = \left( \displaystyle\sum\limits_{j=1}^k \left[(y_i^j-y_{i+1}^j)/d_i\right]^2 \left[(s_i^j)^2+(s_{i+1}^j)^2\right] \right)^{-1} \,, \end{array} \label{eq:WMAVAR} \end{equation} where $|\dots|$ denotes the Euclidean norm, and $p_i$ are the weights. The expression for $p_i$ is derived from classical error propagation low and is therefore theoretically correct. However, this expression cannot be applied in practice because it generates a singular when $d_i=0$, that is, a case in which adjacent measurements are equal. This situation is not extraordinary. To eliminate this problem, a simplified formula was proposed in \cite{Malkin2008j}: \begin{equation} p_i=\left( \displaystyle\sum\limits_{j=1}^k \left[(s_i^j)^2 + (s_{i+1}^j)^2 \right] \right)^{-1} \,. \label{eq:p_alt} \end{equation} Numerous tests with real data revealed no practical difference between the use of Eqs.~(\ref{eq:WMAVAR}) and that of (\ref{eq:p_alt}) in computing for WMAVAR. It is easy to see that WMAVAR (WMADEV) is a universal definition that encompasses all other variants such as AVAR (ADEV), WAVAR (WADEV), and MAVAR (MADEV) as special cases. It can be also noted that for a time series whose $s_i$ values are close to one another WMADEV is approximately $\sqrt{k}$ times as large as the WADEV computed for one (each) vector dimension. For example, given three series of equally precise measurements of station coordinates $X$, $Y$, and $Z$ having about the same WADEV estimate $\sigma_1$, the 3D WMADEV estimate will be equal to roughly $\sqrt{3}\sigma_1$. \begin{figure*}[ht!] \centering \includegraphics[width=\hsize]{cpo_fcn.eps} \caption{The original IVS CPO series (upper row), the same series with the ZM4 CPO model (middle row), and the residuals between CPO and ZM4 model (lower row). Different scatter statistics are presented to the right of each series.} \label{fig:cpo_fcn} \end{figure*} A main advantage of AVAR when used as a noise level estimator is its weak dependence on low-frequency variations. Figure~\ref{fig:cpo_fcn} depicts the results of the CPO determination in the International VLBI Service for Geodesy and astrometry (IVS) \cite{Schuh2012,Boeckmann2010}. One can see that the CPO series include quasi-harmonic signal with period about 430 days corresponding to the free Earth's core nutation (FCN) and low-frequency components caused by the errors in the precession-nutation model. In this study, CPO is modeled by the ZM2 and ZM4 models \cite{Malkin2007i,Malkin2013d} The results presented in Fig.~\ref{fig:cpo_fcn} show that the WRMS estimate depends heavily on the low-frequency components of the analyzed signal. The removal of these components requires highly accurate modeling of the underlying physical processes. By contrast, AVAR is practically unaffected by the presence of trend and low-frequency harmonics. An important AVAR application is in characterizing the spectral behavior of a time series \cite{Allan1981,Allan1987}. The assessment of noise spectral characteristics may be very important in astrometric and geodetic data analyses. For example, a choice of the method for computing the station velocity uncertainty from the station position time series crucially depends on the noise spectral type \cite{Mao1999,Williams2003}. Knowledge of the spectral noise structure facilitates the construction of a proper covariance matrix of a geodetic signal \cite{Zhang1997}. As with clock deviations, the noise processes in geodetic and astronomical time series can generally be described by a power law \begin{equation} S_y(f) = S_0 \, f^\alpha \,, \label{eq:power_law} \end{equation} where $S_0$ is a normalizing constant and $\alpha$ is the spectral index. It was supposed above that $\{y_i\}$ are the original measurements and that AVAR is just used as a scatter level estimate. The original AVAR definition as given in full by \cite{Allan1966} is \begin{equation} AVAR = \sigma^2(\tau) = \frac{1}{2} \langle (\bar y_i - \bar y_{i+1})^2 \rangle \,, \label{eq:original_AVAR} \end{equation} where $\bar y_i$ are averaged values over the sampling interval (number of points) $\tau$. Then, the following relation can be used to identify noise spectral characteristics: \begin{equation} \log(\sigma^2(\tau)) = \mu \, \log(\tau) \,. \label{eq:avar_tau} \end{equation} The $\mu$ value is connected to the spectral index by $\alpha=-(\mu+1)$ for $-3 < \alpha < 1$ ($-2 < \mu < 2$) \cite{Allan1981}. Then, the prevailing noise type in the time series can by classified as white noise ($\mu=-1, \alpha=0$), flicker noise ($\mu=0, \alpha=-1$), random walk ($\mu=1, \alpha=-2$), or another noise type (see Fig.~\ref{fig:noise_type} for illustration). It must be kept in mind, however, that the analogy of the AVAR-derived log-log slope $\mu$ with the spectral index is only justified for stationary stochastic processes. Given real signals, noise often exhibits different spectral characteristics at various frequency bands; thus, a full noise spectrum consists of several components: \begin{equation} S_y = \sum_{j=1}^{N_f}{c_j\,f_j^{\alpha_j} \,.} \label{eq:complex_spectra} \end{equation} \begin{figure}[ht!] \centering \includegraphics[width=0.9\columnwidth]{avar_noise.eps} \caption{Dependence of AVAR on the sampling interval $\tau$ for the three most common typical noise types.} \label{fig:noise_type} \end{figure} It must be noted that this noise classification method must be practiced with care. First, the analyzed time series should be sufficiently long; at least several hundreds of measurements are needed to obtain a reliable result \cite{Williams2003}. Moreover, trend, seasonal and other long-periodic terms (beyond the frequency band of the analyst's interest) from the analyzed signal should be removed prior to spectral analysis; otherwise, the log-log spectrum (\ref{eq:avar_tau}) can be distorted \cite{LeBail2006,LeBail2012,Khelifa2014}. In case of uniform noise, computing and analyzing the classical power spectra may be preferable, such as through Fourier transform. In case of mixed noise, however, AVAR analysis can provide valuable supplemental information for adjusting the analysis procedure. Indeed, the use of the AVAR algorithm in characterizing the spectral behavior of a time series can easily be generalized for weighted and multi-dimensional measurements through the application of WMAVAR and weighted average at each sampling interval $\tau$. It can be mentioned here that the computation of weighted average uncertainty does not generate an unambiguous solution. A possible approach to solving this problem in practical applications is discussed in \cite{Malkin2013e}. Among other applications, AVAR can be used to estimate the Hurst parameter \cite{Bregni2008}. The Hurst parameter (or Hurst exponent) $H$ is related to such time series statistical properties as long-term memory, self-similarity (fractal dimension) and spectral index. Notably, the original AVAR is defined for evenly spaced time series, which generally does not apply to all geodetic and especially astronomical data. The simplest method of solving this problem is to form so-called normal points that average data over equally spaced intervals. Such a method was used in \cite{Feissel2003a} for radio source position time series and in \cite{Feissel2007} for station position time series. Nonetheless, the process of averaging the original series inevitably leads to loss of information regarding high-frequency noise characteristics. However, if AVAR is merely used as a measure of random deviation level, uneven data spacing should not influence analysis results, supposing stationary noise. A special type of unevenly spaced data is the time series that consists of equally spaced measurements with gaps. In \cite{LeBail2006}, a special technique has been developed to fill in data gaps prior to statistical analysis. Additional considerations for the application of AVAR to series with gaps are given in \cite{Sesia2008}. \section{Use of AVAR in geodesy and astrometry} \label{sect:results} This section provides an overview of the applications of AVAR in astrometry, geodesy, and geodynamics. In these fields, AVAR is mainly applied to investigate space geodesy station displacements and the stability of VLBI-derived celestial object positions. Several examples of AVAR applications are discussed in the following subsections to effectively illustrate some features of AVAR estimates. \subsection{Celestial reference frame} The studies described in this subsection are CRF-related investigations, such as the analysis of radio source coordinates catalogs and the improvement of the International Celestial Reference Frame (ICRF) through the detection of radio sources with unstable positional behavior. Accuracy assessment of the CRF catalogs is not a trivial task because only the differences between the radio source positions in various catalogs are known (this situation is very similar to the case of clock comparison). A possible absolute method of catalog comparison is based on the comparison of the noise levels in the CPO series as computed with the compared source catalogs \cite{Malkin2008j}. In this study, the radio source position catalogues obtained in 8 IVS analysis centers were compared by means of two scatter indices. Both indices are based on the analysis of the CPO series computed with different catalogs. The first index is computed from analysis of the residuals of CPO series with respect to the IAU 2000/2006 precession-nutation model. This index occurred to be less sensitive, and, besides, it is not fully independent, since the IAU model was determined form VLBI data analysis using some CRF realizations under study. Conversely, WADEV and WMADEV indices provide an independent estimate of the quality of CRF realizations. The WMADEV occurred to be most sensitive to differences in CRF realizations, and thus can be considered as a preferred scatter index. With the use of this method, the ICRF catalog \cite{Ma1998,Fey2004a}, hereafter referred to as ICRF1, was compared with the catalog RSC(PUL)07C02 derived at the Pulkovo observatory \cite{Sokolova2007}. The latter was computed as combined catalog using several individual catalogs obtained in the IVS analysis centers in 2005. It uses more observations than ICRF1 catalog computed in 1995, and a new method of determination and elimination of the systematic errors in source positions. Two CPO time series for 2002--2006 obtained with two CRF realizations were computed. The results of comparison are presented in Table~\ref{tab:cpo_cat} and show that the noise level of the CPO series computed with the Pulkovo catalog is lower than that of the series computed with ICRF1 This finding suggests that the catalog RSC(PUL)07C02 was more accurate than ICRF1. This conclusion was later confirmed by improvement of the EOP results obtained with the new catalog. The analysis of the astrometric aspects of catalog computation and comparison is beyond the scope of this paper. It is only important here to show that estimation of the noise level in the CPO time series can serve as an external characterization of the accuracy of radio source catalogs (CRF realizations). Subsequently, a similar comparison was made in \cite{Malkin2013b} between two ICRF realizations, ICRF1 \cite{Fey2004a} and ICRF2 \cite{Malkin2015c}. The latter was computed in 2009 and using about 6.5~mln observations collected during 1979--2009 as compared with 1.6~mln observations collected in 1979--1995 used for ICRF1. As a result, several direct catalog comparisons showed much better precision and of ICRF2 with respect to ICRF1 \cite{Malkin2015c}. Two CPO time series for 2002--2010 obtained with two CRF realizations were computed. The result of the CPO noise analysis made in \cite{Malkin2013b} and presented in Table~\ref{tab:icrf2-icrf1} independently confirmed that ICRF2 is more accurate than ICRF1. \begin{table} \centering \caption{Noise level in the CPO time series computed with Pulkovo and ICRF1 catalogs of radio source positions as estimated with weighted AVAR.} \label{tab:cpo_cat} \begin{tabular}{cccc} \hline Catalog & \multicolumn{2}{c}{WADEV} & WMADEV \\ & $dX$ & $dY$ & \\ & $\mu$as & $\mu$as & $\mu$as \\ \hline ICRF1 & 112.8 & 108.7 & 168.1 \\ Pulkovo & 105.0 & 105.8 & 160.9 \\ \hline \end{tabular} \end{table} \begin{table} \centering \caption{Noise level in the CPO time series computed with ICRF1 and ICRF2 catalogs of radio source positions as estimated with weighted AVAR.} \label{tab:icrf2-icrf1} \begin{tabular}{cccc} \hline Catalog & \multicolumn{2}{c}{WADEV} & WMADEV \\ & $dX$ & $dY$ & \\ & $\mu$as & $\mu$as & $\mu$as \\ \hline ICRF1 & 101.6 & 107.3 & 148.4 \\ ICRF2 & ~91.7 & ~89.2 & 128.3 \\ \hline \end{tabular} \end{table} It should be realized that both $dX$ and $dY$ are merely the spherical coordinates of the celestial pole, and their statistical behaviors are closely correlated. Thus, the 2D WMADEV estimate serves as a compact and convenient tool for describing the noise component in CPO (and other similar) series. The noise characteristics of a source position time series can be used for source ranking in terms of temporal and spatial stabilities, as well as for compiling a list of sources that are not stable enough to be included in VLBI global solution and thus requires special handling. The position instabilities of 16 radio sources with long and dense observational history were investigated in detail, including the spectral characteristics of the source position time series through the AVAR analysis \cite{Feissel2000a}. The analysis results showed that the time variabilities of these objects exhibit varied spectral characteristics that may be ascribed to diversified physical processes. Several statistical techniques were applied to source position time series to assess the VLBI-derived CRF in \cite{Gontier2001}. In particular, AVAR analysis was performed to assess the noise spectrum of 60 source position time series. Sampling interval $\tau$ varied from the initial time span of the series to 1/4 $\ldots$ 1/3 of the data span (six months to four years). It was found in \cite{Gontier2001} that roughly 60\% of the series predominantly exhibited white noise and 40\% of the series mainly displayed flicker noise. Approximately 160 sources have sufficiently long and dense observation histories from which the AVAR can be derived for a one-year sampling interval. This information was employed to investigate source position stability further. An extended similar work was conducted to select a set of positionally stable sources through an analysis of source position time series with the use of WRMS and ADEV \cite{Feissel2003a}. This study also included an analysis of the apparent drifts in radio source positions. In this study, 707 sources were analyzed, and 199 were recommended as candidate core sources for the subsequent ICRF realization. In particular, roughly 60\%, 35\%, and 5\% of the series exhibited white noise, flicker noise, and random work noise, respectively. The methods developed in \cite{Feissel2003a} for source stability characterization were also used to analyze recent source position time series \cite{LeBail2010a}. This study investigated the position stability of ICRF2 sources using AVAR and considered the apparent linear source motions. It was found that the selection of stable core sources depends on the configuration of the VLBI solution used to derive the position time series and may differ by up to 20\% of the selected sources. Furthermore, source stability found in \cite{LeBail2010a} improved significantly with respect to that found in \cite{Feissel2003a} because of the advances in observing technology as well as the increase in observing networks, which directly affect the precision and accuracy of VLBI results \cite{Malkin2009g}. Another analysis performed in \cite{LeBail2010a} found that the log-log slope $\mu$ computed according Eq.~(\ref{eq:avar_tau}) may not remain constant with time for long radio source position time series. Subsequently, this work was extended by incorporating the most recent observations in the analysis \cite{LeBail2014a}. AVAR was also applied to the time series analysis of 15 source positions provided by 9 IVS analysis centers in the framework of ICRF2 preparation campaign \cite{Malkin2015b}. In particular, it was found that AVAR generates an inadequate estimate of noise level when jumps are present in a time series, as discussed above. In general cases, a composite index of WRMS and WADEV can be applied to enhance the robustness of the measure of source position stability, as previously discussed in \cite{Feissel2003a}. \subsection{Geodesy} Many studies have confirmed that AVAR is an effective tool for investigating the noise characteristics of station position time series, particularly those obtained with space geodetic methods, such as VLBI, GPS, satellite laser ranging (SLR), and Doppler orbitography and radiopositioning integrated by satellite (DORIS). AVAR was used to estimate the week-to-week repeatability of the station position of the European Permanent GPS Network (EPN) station position in the presence of seasonal station movements \cite{Malkin2001n}. The use of AVAR in this case is reasonable because AVAR does not require the estimation of a systematic trend and seasonal variations in station positions, which are difficult to model. In the process, the WRMS noise estimate that was previously used to compare station position time series is distorted. AVAR was also applied to the time series analysis for baseline length in relation to GPS-based deformation monitoring applications \cite{Roberts2002}, where this approach was used to determine the temporal upper limit accuracy of each monitoring system and to help distinguish a systematic error from a genuine geophysical signal for a new measurement. The noise spectrum in DORIS ground station motion was investigated by applying AVAR to the decomposition of the 3D signal into its principal components in the time domain \cite{LeBail2006}. The noise level was estimated, and white noise was dominant after the linear drift in the series of station coordinates was eliminated. In addition, the AVAR signature was sensitive to periodic terms. To avoid bias in the estimation of noise type, three periodic terms, namely annual, semi-annual, and 117.3-day, were filtered out prior to computation of AVAR. The annual and semiannual seasonal variations are usually observed in the station position series derived from astronomical and space geodesy observations see \cite{Dong2002,Malkin2013c} and literature cited therein. The 117.3-day signal is specific for DORIS observations. It is connected with the 117.3-day period in orbit node movement for the TOPEX/Poseidon satellite extensively used during many years for DORIS. Besides, the method of estimating the AVAR confidence interval proposed in \cite{Howe1981} was applied in \cite{LeBail2006} to the DORIS stations position time series. This became usual practice in future works \cite{LeBail2010a,LeBail2012,LeBail2014a}. The time series of the station coordinates obtained from various space geodetic observations were analyzed in \cite{LeBail2007}. It was found that the station position series obtained from VLBI, SLR, and DORIS contain white noise, whereas the majority of GPS station motions exhibit flicker noise. The atmospheric loading time series (change in the station positions caused by variations of the atmospheric pressure) also contain white noise, as do the series of the transformation parameters in the midst of position time series obtained by different techniques. A similar but more detailed analysis was presented in \cite{Feissel2007}. The noise and low-frequency components of coordinate variations were investigated with the data obtained from three permanent GPS stations located in the territory of Pulkovo Observatory \cite{Gorshkov2012b}. AVAR was used in this work to assess the noise in the station position time series. The results were compared with the spectral index obtained from Hurst exponent analysis. Both methods suggested that the position time series for the three Pulkovo stations are generally dominated by flicker and mixed white-flicker noises. An AVAR analysis of the noise characteristics in the time series for DORIS station position was presented in \cite{Khelifa2014}. Once the trends and seasonal components from the analyzed time series were eliminated, the results of the subsequent AVAR analysis indicated that the three solutions are dominated by white noise in all three components (north, east, and vertical). Upon confirming the dominant noise type in the position time series, an adequate method for further analysis was chosen. \subsection{Earth's rotation and geodynamics} Earth orientation parameters (EOP) define the rotation of the terrestrial reference system ITRS with respect to the celestial reference system ICRS \citep{IERSConv2010}. These parameters include polar motion (movement of the Earth's rotation axis in ITRS), precession-nutation (movement of the Earth's rotation axis in ICRS), and Universal Time, which depends on the rotational speed of the Earth around its axis. EOP are determined through the same space geodetic methods described in the previous subsection. The final EOP series for practical use are derived by combining the results obtained at different EOP analysis centers. The noise characteristics of EOP series should be investigated for different tasks, such as quality assessment of the EOP series and the weighting of these series during combination. The computation method of the combined EOP series at the Bureau International de L'Heure (BIH) involved an AVAR analysis of the input EOP series \cite{Feissel1980}. Specifically, AVAR was used to obtain the frequency dependence of the noise power spectrum in the combined series. In particular, it was found that the yearly bias in a time series with respect to the combined series is dominated by flicker noise, which may indicate accidental local and instrumental changes. However, the individual results obtained at separate stations are not stable enough; thus, corrections for individual stations cannot be reliably applied. AVAR analysis is also a main step in the EOP combination procedure employed by the International Earth Rotation and Reference Systems Service (IERS), which succeeded BIH \cite{Gambis2002}. AVAR is used to characterize the internal random deviations in the EOP time series at various time scales. White noise and flicker noise dominate these series Another application of the AVAR is in comparing EOP series to evaluate their internal precision. These data are used to assign proper weights to EOP series during combination. An analysis of several CPO series computed at different IVS analysis centers was performed in \cite{Malkin2013b}. The results of this study confirmed that the WRMS estimate of the noise in CPO series is heavily dependent on the model of low-frequency variations, whereas WADEV is practically independent of such a model. The use of WADEV estimates also mitigated the influence of outliers significantly. Using WADEV estimates also allowed us to severely mitigate the influence of outliers. Another result of the study is that 2D WMADEV estimates are close to the averaged WADEV estimates computed for $dX$ and $dY$ and multiplied by $\sqrt{2}$, as was discussed in Section~\ref{sect:overview}. Seasonal and non-seasonal components in the geocenter motion signal measured by satellite geodesy techniques were analyzed to determine the connection of these components with geophysical fluids \cite{Feissel2006a}. AVAR was used in this study to compute and compare the spectral characteristics of geodetic and geophysical geocenter time series over a sampling period of one month to one year. In particular, the dependence of the log-log $\tau$-AVAR pattern on the seasonal components in the analyzed signal was determined. Geodetic and modeled excitation functions were compared via AVAR in \cite{Bizouard2014a}. In this study AVAR analysis was applied to Earth's rotation excitation functions, complementing other methods of statistical analysis. Bias and seasonal variations were removed from the signals prior to AVAR computation. Comparison of ADEV obtained for geodetic and geophysical excitations series, and their residuals for different sample interval $\tau$ allowed to draw a conclusion on the geophysical excitation of the Earth's rotation. \section{Conclusions} AVAR is an effective statistical tool for analyzing the time series of observational data in astronomy and geodesy, as well as of all other time series. Important independent characteristics of the noise component in the studied signal can be obtained through this tool. The main application of AVAR to time series analysis is in determining the signal scatter level and spectral analysis with the primary aim to identify the dominant noise type in a time series. This information can be applied to refined data analysis, such as computing realistic uncertainty of station velocities \cite{Mao1999,Williams2003}) and regularization of EOP series \cite{LeBail2012}. AVAR modifications, namely, WAVAR, MAVAR, and WMAVAR, were proposed for processing weighted and/or multi-dimensional data \cite{Malkin2008j}. These modifications serve as effective and convenient tools for data analysis in geodesy and astronomy. WMAVAR is the most general estimator and encompasses AVAR, WAVAR, and MAVAR as special cases. In particular, 2D WMAVAR can be used for complex data processing. Indeed, the AVAR modifications we proposed should not be confused with the ``original'' modified AVAR definition \cite{Allan1981}. An important advantage of AVAR and its more refine versions over WRMS in practical application is its weak sensitivity to low-frequency signal variations. By contrast, WRMS depends heavily on the model used to eliminate the systematic component of the studied signal. Our study showed that WAVAR is more robust to outliers than the classical AVAR is; however, both AVAR and WAVAR may estimate noise level erroneously when jumps occur in a time series. AVAR is also widely used to investigate the spectral characteristics of a time series and is a powerful tool for noise type identification through log-log representation. In particular, AVAR facilitates the effective analysis of signals with different types of noise at various frequency bands. AVAR is supposed to be more computationally effective than classical spectral methods, such as Fourier transform; in our opinion, however, this advantage is no longer significant at present. A detailed comparison of two methods for estimating spectral noise characteristics may be interesting and useful though. It must be noted that the AVAR method needs further investigation and development for some applications. First, many geodetic and astronomical series are unevenly spaced as was discussed in Section~\ref{sect:overview}. Frequent examples include station position time series with gaps, radio source position time series, VLBI-derived session-wise EOP series. Another open issue regarding the application of AVAR to the analysis of geodetic and astronomical time series stems from possible correlations between the measurements that may distort statistical analysis results substantially. Finally, we can conclude that despite its limitations and some unresolved issues, AVAR nevertheless remains one of the most powerful tools for analyzing a wide range of physical measurement time series. \section*{Acknowledgements} This work was partly funded by the Russian Government Program of Competitive Growth of Kazan Federal University. The author is grateful to three anonymous reviewers for careful reading of the manuscript and helpful comments.
f3c41de1f43c2bc817b797a57f33e6bb4d4f67b9
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} There are many ways to define the Fibonacci numbers. An equivalent approach to the standard recurrence relation, where $F_{n+1} = F_n + F_{n-1}$ and $F_1 = 1$ and $F_2=2$, is that they are the unique sequence of integers such that every positive number can be written uniquely as a sum of non-adjacent terms. This expansion is called the Zeckendorf decomposition \cite{Ze}, and much is known about it. In particular, the distribution of the number of summands of $m \in [F_n, F_{n+1})$ converges to a Gaussian as $n\to\infty$, with mean and variance growing linearly with $n$. Similar results hold for a large class of sequences which have a notion of legal decomposition leading to unique decomposition; see \cite{Al, BDEMMTTW, CFHMN1, CFHMN2, Day, DDKMMV, DDKMV, DG, GT, GTNP, Ha, Ho, Ke, LT, Len, Lek, KKMW, MW1, MW2, Ste1, Ste2}. Given a sequence $\{H_n\}$, one can frequently prove that the mean and the variance of the number of summands of $m \in [H_n, H_{n+1})$ grows linearly with $n$. Explicitly, there are constants $a, b, C$ and $d$ such that the mean is $an+b+o(1)$ and the variance is $Cn + d + o(1)$. The difficulty is proving that $a$ and $C$ are positive, which is needed for the proofs of Gaussian behavior. Until recently, the only approaches have been technical and involved generating functions, partial fraction expansions and generalized Binet formulas applied to polynomials associated to the characteristic polynomials of the sequence, which have required a lot of work to show the leading terms are positive for such recurrences. The point of this work is to bypass these arguments through elementary counting. We concentrate on positive linear recurrence sequences (defined below) to highlight the main ideas of the method; with additional work these arguments can be extended to more general sequences (see \cite{CFHMNPX}). In addition to the arguments below, one can also obtain similar results (though not as elementarily) through Markov chains \cite{B-AM} or through an analysis of two dimensional recurrences \cite{LiM}. \begin{definition} \label{defn:goodrecurrencereldef}\label{def:goodrecurrence} A sequence $\{H_n\}_{n=1}^\infty$ of positive integers is a \textbf{Positive Linear Recurrence Sequence (PLRS)} if the following properties hold. \begin{enumerate} \item \emph{Recurrence relation:} There are non-negative integers $L, c_1, \dots, c_L$\label{c_i} such that \begin{equation*} H_{n+1} \ = \ c_1 H_n + \cdots + c_L H_{n+1-L},\end{equation*} with $L, c_1$ and $c_L$ positive. \item \emph{Initial conditions:} $H_1 = 1$, and for $1 \le n < L$ we have \begin{equation*} H_{n+1} \ =\ c_1 H_n + c_2 H_{n-1} + \cdots + c_n H_{1}+1. \end{equation*} \end{enumerate} We define the \textbf{size} of $\{H_n\}$ to be $c_1+\cdots+c_L$ and the \textbf{length} of $\{H_n\}$ to be $L$. \end{definition} \begin{definition} Let $\{H_n\}$ be a PLRS. A decomposition $\sum_{i=1}^{m} {a_i H_{m+1-i}}$\label{a_i} of a positive integer $\omega$ (and the sequence $\{a_i\}_{i=1}^{m}$) is \textbf{legal}\label{legal} if $a_1>0$, the other $a_i \ge 0$, and one of the following two conditions holds. \begin{itemize} \item Condition 1: We have $m<L$ and $a_i=c_i$ for $1\le i\le m$. \item Condition 2: There exists $s\in\{1,\dots, L\}$ such that \begin{equation*} a_1\ = \ c_1,\ a_2\ = \ c_2,\ \dots,\ a_{s-1}\ = \ c_{s-1}\ {\rm{and}}\ a_s<c_s, \end{equation*} and $\{b_i\}_{i=1}^{m-s}$ (with $b_i = a_{s+i}$) is legal. \end{itemize} If $\sum_{i=1}^{m} {a_i H_{m+1-i}}$ is a legal decomposition of $\omega$, we define the \textbf{number of summands}\label{summands} (of this decomposition of $\omega$) to be $a_1 + \cdots + a_m$. Furthermore, we define {two types of blocks}, where a \textbf{block} is a nonempty ordered subset of the coefficients $[a_i, a_{i+1},\ldots,a_{i+j}]$ inclusive: \begin{itemize} \item a \textbf{Type 1 block} corresponds to Condition 1, and has length $m < L$ and size $a_i+ \cdots +a_{i+m-1}$, \item a \textbf{Type 2 block} corresponds to Condition 2, and has length $s\le L$ and size $a_i+ \cdots +a_{i+s-1}$. \end{itemize} \end{definition} \begin{remark} A Type 2 block has three key properties. \begin{itemize} \item A legal decomposition of $\omega$ stays legal if a Type 2 block is inserted (between Type 1 and/or 2 blocks) or removed and indices are shifted appropriately. \item If we know the size of a Type 2 block, the block's content and its length are uniquely determined. So we can define a \textbf{length function} $\ell(t)$ to be the length of a Type 2 block with size $t$. \item A Type 2 block always has nonnegative size and strictly positive length. Specifically, consider a Type 2 block with size 0. Then, in Condition 2, we always have $a_1 = 0 < c_1$, so $s = 1$. Thus a Type 2 block with size 0 has length 1. In other words, $\ell(0) = 1$ holds for all PLRS. \end{itemize} If a legal decomposition contains a Type 1 block, then it must be the last block. Thus any legal decomposition contains at most one Type 1 block. A Type 1 block, according to Condition 1, always has positive size and positive length. \end{remark} The following two examples clarify the above. \begin{example} \textbf{The Fibonacci Sequence} (size 2 and length 2).\\ Type 1 block: [1].\\ Type 2 blocks: [0], [1 0].\\ An example of a legal decomposition: $F_5 + F_3 + F_1$ with block representation: [1 0] [1 0] [1].\\ After removing the second to last block, the new block representation is [1 0] [1].\\ The resulting legal decomposition is $F_3 + F_1$.\\ \end{example} \begin{example}\textbf{PLRS sequence $H_n = 2H_{n-1} + 2H_{n-2} + 0 + 2H_{n-4}$} (size 6 and length 4).\\ Type 1 blocks: [2], [2 2], [2 2 0].\\ Type 2 blocks: [0], [1], [2 0], [2 1], [2 2 0 0], [2 2 0 1].\\ An example of a legal decomposition: $H_7 + 2H_4 + H_1$ with block representation: [1] [0] [0] [2 0] [0] [1].\\ After removing the second to last block, the new block representation is [1] [0] [0] [2 0] [1].\\ The resulting legal decomposition is $H_6 + 2H_3 + H_1$.\\ \end{example} Before we state our main result we first set some notation. \begin{definition} Let $\{H_n\}$ be a Positive Linear Recurrence Sequence. For each $n$, let the discrete outcome space $\Omega_n$ be the set of legal decompositions of integers in $[H_n, H_{n+1})$. By the Generalized Zeckendorf Theorem (see for example \cite{MW2}) every integer has a unique legal decomposition, so $|\Omega_n| = H_{n+1} - H_n$. Define the probability measure on subsets of $\Omega_n$ by \begin{equation*}\mathbb{P}_n(A) \ = \ \sum_{\omega \in A \atop \omega \in \Omega_n} \frac1{H_{n+1}-H_n}, \ \ \ A \subset \Omega_n; \end{equation*} thus each of the $H_{n+1}-H_n$ legal decompositions is weighted equally. We define the random variable $K_n$ by setting $K_n(\omega)$ equal to the number of summands of $\omega \in \Omega_n$. When $n > 2L$ (so there are at least three blocks) we define the random variable $Z_n$ by setting $Z_n(\omega)$ equal to the size of the second to last block of $\omega \in \Omega_n$. Note that the second to last block must be a Type 2 block. Finally, we define the random variable $L_n$ by setting $L_n(\omega)$ equal to the length of the second to last block of $\omega \in \Omega_n$; i.e., $L_n(\omega) = \ell(Z_n(\omega))$. \end{definition} As remarked above, previous work has shown that $\E{K_n} = an + b + f(n)$ where $a > 0$ and $f(n) = o(1)$; this can be proved through very simple counting arguments (see \cite{CFHMNPX}). While it is also known that $\V{K_n} = Cn + d + o(1)$, previous approaches could not easily show $C \neq 0$. We elementarily prove $C > 0$ by giving a positive lower bound $c$ for it. \begin{theorem}\label{thm:main} Let $\{H_n\}$ be a positive linear recurrence sequence with size $S$ and length $L$. Then there is a $c > 0$ such that $\V{K_n} \ge cn$ for all $n > L$. \end{theorem} We sketch the proof. We can remove the second to last block of a legal decomposition to get a shorter legal decomposition, forming relations between longer legal decompositions and shorter legal decompositions. We then use strong induction and conditional probabilities to prove the theorem. \begin{remark} As it is known that $\V{K_n} = Cn + d + o(1)$, to prove that $C > 0$ it would suffice to show $\lim_{n\to\infty} {\rm Var}{[K_n]}$ diverges to infinity. Unfortunately the only elementary proofs we could find of this also establish the correct growth rate; we would be very interested in seeing an approach that yielded (for example) ${\rm Var}[K_n] \gg \log n$ (which would then immediately improve to implying $C>0$). \end{remark} \section{Lemmas derived from Expectation} We first determine a relationship between $K_n$ and $Z_n$. Then, with the help of $\E{K_n} = an + b + f(n)$, we explain how to explicitly determine the positive lower bound $c$. \begin{lemma} Let $n > 2L$. For all $0 \le t < S$, we define $S_t := \{\omega \in \Omega_n|Z_n(\omega) = t\}$, and $h_t(\omega)$ to be the decomposition after removing the second to last block of $\omega$. (When we remove the second to last block with size $t$, we completely remove that block from $\omega$ and shift all the indices to the left of that block by $\ell(t)$.) When we remove the second to last block (a Type 2 block) from $\omega$, then $h_t(\omega)$ is legal and $h_t$ is a bijection between $S_t$ and $\Omega_{n-\ell(t)}$. \end{lemma} \begin{proof} Let $\omega \in \Omega_n$ be arbitrary and consider $h_t(\omega)$. Since the block we remove has size $t$ and thus length $\ell(t)$, $h_t(\omega)$ must be in $\Omega_{n-\ell(t)}$. Next, consider $\omega, \omega' \in S_t$, such that $h_t(\omega) = h_t(\omega')$. As the size determines the composition for Type 2 blocks, we are removing the same block at the same position for $\omega, \omega'$. This implies $\omega = \omega'$. Finally, for any $\omega \in \Omega_{n-\ell(t)}$, if we insert the size $t$ type 2 block before its last block, we get a legal decomposition in $\Omega_n$. Thus $h_t$ is surjective. Therefore, $h_t$ is a bijection between $S_t$ and $\Omega_{n-\ell(t)}$ \end{proof} \begin{corollary} We have \begin{equation*}\begin{split} \PP{Z_n = t} \ = \ \frac{|S_t|}{|\Omega_n|} \ =\ \frac{|\Omega_{n-\ell(t)}|}{|\Omega_n|} \ =\ \frac{H_{n - \ell(t) + 1} - H_{n - \ell(t)}}{H_{n+1} - H_n}. \end{split}\end{equation*} \end{corollary} \begin{remark} As \begin{equation}\PP{Z_n = 0}\ \ge\ \PP{Z_n = 1} \ge \dots \ge \PP{Z_n = S-1}\end{equation} and the sum of these $S$ terms is 1, we have \begin{equation}\label{8}\PP{Z_n = 0}\ \ge\ \frac{1}{S},\end{equation} (which is the consequence we need below). \end{remark} For an arbitrary $\omega \in S_t$, the second to last block has size $Z_n = t$, and the remaining blocks form a legal decomposition in $\Omega_{n-\ell(t)}$ with size $K_{n-\ell(t)}(h_t(\omega))$, so $K_n(\omega) = K_{n-\ell(t)}(h_t(\omega)) + t$. Since $h$ is a bijection, we have the following two equations: \begin{equation}\label{1} \begin{split} \E{K_n|Z_n = t} &\ = \ \E{K_{n-\ell(t)} + t}\\ &\ = \ a(n-\ell(t)) + b + f(n-\ell(t)) + t, \end{split} \end{equation} and \begin{equation}\label{2} \begin{split} \E{K_n^2|Z_n = t} &\ = \ \E{(K_{n-\ell(t)} + t)^2}\\ &\ = \ \E{K_{n-\ell(t)}^2 + 2t{K_{n-\ell(t)}} + t^2}\\ &\ = \ \E{K_{n-\ell(t)}^2} + 2t\E{K_{n-\ell(t)}} + t^2\\ &\ = \ \E{K_{n-\ell(t)}^2} + 2t[a(n-\ell(t)) + b + f(n-\ell(t))] + t^2. \end{split} \end{equation} Furthermore, by $\eqref{1}$ we have \begin{equation}\label{3} \begin{split} \E{K_n} &\ = \ \sum\limits_{t = 0}^{S-1} \PP{Z_n = t}\cdot \E{K_n|Z_n = t}\\ &\ = \ \sum\limits_{t = 0}^{S-1} \PP{Z_n = t}\cdot [a(n-\ell(t)) + b + f(n-\ell(t)) + t]\\ &\ = \ an + b + \sum\limits_{t = 0} ^ {S-1} \PP{Z_n = t}\cdot [t + f(n-\ell(t)) - a\ell(t)]\\ &\ = \ an + b + f(n),\\ \end{split} \end{equation} where the last equality comes from the definition of $f(n)$. If we set $Y_n(\omega) := Z_n(\omega) + f(n-L_n(\omega)) - aL_n(\omega)$, then we have \begin{equation}\label{4} \E{Y_n} \ =\ \sum\limits_{t = 0} ^ {S-1} \PP{Z_n = t}\cdot [t + f(n-\ell(t)) - a\ell(t)] \ = \ f(n). \end{equation} Now that we have $\E{Y_n}$, we use it to estimate $\V{Y_n}$. \begin{lemma}\label{lem:varyn} For $n$ sufficiently large we have \begin{equation}\label{9} \V{Y_n}\ > \ \frac{a^2}{2S}. \end{equation} \end{lemma} \begin{proof} First, for all $n > 2L$ we have \begin{equation*} \begin{split} \V{Y_n} &\ = \ \E{Y_n^2} - \left(\E{Y_n}\right)^2\\ &\ = \ \left(\E{(Z_n - aL_n + f(n - L_n))^2}\right) - \left(f(n)\right)^2 \\ &\ = \ \left(\E{(Z_n - aL_n)^2}+\E{2(Z_n - aL_n)\cdot f(n - L_n)} + \E{f(n - L_n)^2}\right) - \left(f(n)\right)^2. \end{split} \end{equation*} Note that $Z_n - aL_n$ is bounded since $ -aL \le Z_n - aL_n \le S$ for all $n > 2L$. Also we know $f(n) = o(1)$, so $f(n-L_n) = o(1)$ since $L_n \le L$. Hence the following three limits are all zero: \begin{equation} \lim\limits_{n \to \infty}\E{2(Z_n - aL_n)\cdot f(n - L_n)}\ = \ \lim\limits_{n \to \infty}\E{f(n - L_n)^2}\ = \ \lim\limits_{n \to \infty} \left(f(n)\right)^2 \ =\ 0.\end{equation} Further, we know \begin{equation} \V{Y_n} - \E{(Z_n - aL_n)^2}\ =\ \E{2(Z_n - aL_n)\cdot f(n - L_n)} + \E{f(n - L_n)^2} - \left(f(n)\right)^2,\end{equation} so \begin{equation}\label{5} \lim\limits_{n \to \infty}\left(\V{Y_n} - \E{(Z_n - aL_n)^2}\right)\ =\ 0. \end{equation} On the other hand, for all $n > 2L$ we have \begin{equation}\label{6}\begin{split} \E{(Z_n - aL_n)^2} &\ = \ \sum\limits_{t = 0}^{S - 1} \PP{Z_n = t} \cdot \left(t - a\ell(t)\right)^2\\ &\ \ge \ \PP{Z_n = 0} \cdot \left(0 - a\ell(0)\right)^2\\ &\ \ge \ \frac{a^2}{S}, \end{split}\end{equation} where the last inequality follows from $\eqref{8}$. By $\eqref{5}$, we know there must exist $N > 2L$ such that for all $n > N$, $|\V{Y_n} - \E{(Z_n - aL_n)^2}| < \frac{a^2}{2S}$, so $\V{Y_n} - \E{(Z_n - aL_n)^2} > -\frac{a^2}{2S}$. Then, by $\eqref{6}$, we get $\V{Y_n} > \frac{a^2}{2S}$ for all $n > N$. \end{proof} Finally, we choose $c$. Let \begin{equation} c\ =\ \text{min}\left\{\frac{\V{K_{L+1}}}{L+1},\ \frac{\V{K_{L+2}}}{L+2},\ \dots,\ \frac{\V{K_{N}}}{N},\ \frac{a^2}{2SL}\right\},\end{equation} Where $N$ is as determined in Lemma \ref{lem:varyn}. For all $n > L$, $H_{n + 1} - H_n > 1$, so there are at least two integers in $[H_n, H_{n+1})$. Since the legal decomposition of $H_n$ has only one summand while that of $H_n + 1$ has two summands, $\V{K_n}$ is nonzero when $n > L$. Hence, $c > 0$. In the next section we show $\V{K_n} \ge cn$ for all $n > L$. \section{A lower bound for the Variance} We prove Theorem \ref{thm:main} by strong induction. While the algebra is long, the main idea is easily stated: we condition based on how many summands are in the second to last block, which must be a type 2 block, and then use conditional probability arguments (inputting results for the mean and smaller cases) to compute the desired quantities. \begin{proof} The base cases $n = L+1, L+2, \dots, N$ are automatically true by the way we choose $c$. Hence, we only need to consider the cases when $n > N$. In the induction hypothesis, we assume $\V{K_r} \ge cr$ for $L < r < n$. In the inductive step, we prove $\V{K_n} \ge cn$ where $n > N$. For $L < r < n$, we have $\V{K_r} \ge cr$ and $\E{K_r} = ar + b + f(r)$, hence \begin{equation}\label{7} \begin{split} \E{K_r^2} &\ =\ \V{K_r} + \left(\E{K_r}\right)^2\\ &\ \ge\ cr + (ar + b + f(r))^2\\ &\ = \ cr + a^2r^2 + b^2 + (f(r))^2 + 2arb + 2arf(r) + 2bf(r). \end{split} \end{equation} By $\eqref{2}$, we have \begin{equation*}\begin{split} \E{K_n^2} &\ = \ \sum\limits_{t=0}^{S-1}\PP{Z_n = t}\cdot\E{K_n^2|Z_n = t}\\ &\ = \ \sum\limits_{t=0}^{S-1}\PP{Z_n = t} \cdot \left(\E{K_{n-\ell(t)}^2} + 2t[a(n-\ell(t)) + b + f(n-\ell(t))] + t^2\right). \end{split}\end{equation*} Note we only need to consider $n > N > 2L$, so $n > n-\ell(t) \ge n - L > L$ for all $0 \le t \le S-1$. Hence, by $\eqref{7}$, \begin{equation*}\begin{split} \E{K_{n-\ell(t)}^2} &\ \ge\ c(n-\ell(t)) + a^2(n-\ell(t))^2 + b^2 + [f(n-\ell(t))]^2 + 2a(n-\ell(t))b\\ &\ \quad\ + 2a(n-\ell(t))f(n-\ell(t)) + 2bf(n-\ell(t)). \end{split}\end{equation*} After we replace $\E{K_{n-\ell(t)}^2}$ in the conditional expectation $\E{K_n^2|Z_n = t}$ with this lower bound, any term either does not depend on $t$ or can be combined with other terms to form $(t+f(n-\ell(t))-a\ell(t))$. The final equation will then have two parts, one of which does not depend on $t$, while the other can be written in the form of $Z_n + f(n - L_n) - aL_n$, which is exactly $Y_n$. We find \begin{equation*}\begin{split} \E{K_n^2} &\ \ge \ \sum\limits_{t=0}^{S-1}\PP{Z_n = t} \cdot \Bigg[\left(c(n-\ell(t)) + a^2(n-\ell(t))^2 + b^2 + [f(n-\ell(t))]^2 + 2a(n-\ell(t))b\right.\\ &\ \quad\ \left. + 2a(n-\ell(t))f(n-\ell(t)) + 2bf(n-\ell(t))\right)+ 2t[a(n-\ell(t)) + b + f(n-\ell(t))] + t^2\Bigg]\\ &\ = \ \sum\limits_{t=0}^{S-1}\PP{Z_n = t} \cdot \Bigg[\left(c(n-\ell(t)) + a^2(n-\ell(t))^2 + b^2 + [f(n-\ell(t))]^2 + 2a(n-\ell(t))b\right. \\ &\ \quad\ \left. + 2a(n-\ell(t))f(n-\ell(t)) + 2bf(n-l( t))\right)+ \left(2tan - 2ta\ell(t) + 2tb + 2tf(n-\ell(t))\right) + t^2\Bigg]\\ &\ = \ (an+b)^2 + cn + \sum\limits_{t = 0}^{S-1}\PP{Z_n = t}\cdot \Bigg[-c\ell(t) - 2a^2n\ell(t) + a^2(\ell(t))^2 + [f(n-\ell(t))]^2 - 2a\ell(t)b\\ &\ \ \ \ + \ 2anf(n-\ell(t)) - 2a\ell(t)f(n-\ell(t)) + 2bf(n-\ell(t)) + 2tan - 2ta\ell(t) + 2tb + 2tf(n-\ell(t)) + t^2 \Bigg]\\ &\ = \ (an+b)^2 + cn + \sum\limits_{t = 0}^{S-1}\PP{Z_n = t}\cdot \Bigg[\left(a^2(\ell(t))^2 + [f(n-\ell(t))]^2 + t^2 - 2a\ell(t)f(n-\ell(t)) - 2ta\ell(t)\right. \\ & \ \ \ + \ \left. 2tf(n-\ell(t)) \right) + 2an\left(t + f(n-\ell(t)) - a\ell(t)\right) + 2b\left(t+f(n-\ell(t))-a\ell(t)\right) - c\ell(t)\Bigg]\\ &\ = \ (an+b)^2 + cn + \sum\limits_{t = 0}^{S-1}\PP{Z_n = t}\cdot\Bigg[\left(t+f(n-\ell(t))-a\ell(t)\right)^2 \\ &\ \ \ \ + \ 2(an+b)\left(t+f(n-\ell(t))-a\ell(t)\right) - c\ell(t)\Bigg]\\ &\ = \ (an+b)^2 + cn + \sum\limits_{t = 0}^{S-1}\PP{Z_n = t}\cdot\left(t+f(n-\ell(t))-a\ell(t)\right)^2\\ &\ \ \ \ + \ 2(an+b)\sum\limits_{t=0}^{S-1}\PP{Z_n = t}\cdot\left(t+f(n-\ell(t)) -a\ell(t)\right) - c\sum\limits_{t=0}^{S-1}\PP{Z_n = t}\cdot \ell(t)\\ &\ = \ (an+b)^2 + cn + \E{(Z_n+f(n-L_n) -aL_n)^2} + 2(an+b)f(n) - c\E{L_n}, \end{split}\end{equation*} where the last equality comes from $\eqref{4}$. We already know $\left(\E{K_n}\right)^2 = (an+b+f(n))^2 = (an+b)^2 + 2(an+b)f(n) + (f(n))^2$, hence \begin{equation*}\begin{split} \V{K_n} - cn &\ =\ \E{K_n^2} - \left(\E{K_n}\right)^2 - cn\\ &\ \ge\ \E{(Z_n+f(n-L_n) -aL_n)^2} - c\E{L_n}-(f(n))^2\\ &\ = \ \E{Y_n^2} - c\E{L_n} - \left(\E{Y_n}\right)^2\\ &\ = \ \V{Y_n} - c\E{L_n}\\ &\ \ge \ \V{Y_n} - cL\\ &\ \ge \ 0, \end{split}\end{equation*} where the last inequality comes from our definition of $c$ and $\eqref{9}$. Therefore, $\V{K_n} \ge cn$ for all $n > L$. In other words, if $\V{K_n} = Cn + d + o(1)$, then $C \ge c > 0$. \end{proof}
1f650f625c1edf72ae688f5aa5674a3ad56a6edd
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The work presented here, roughly speaking, aims at proving a version of the folklore statement: Gravitational waves carry energy away from an isolated system. More precisely, a system with no radiation emitted to infinity must in fact be stationary. On physical grounds it is expected that truly time-periodic gravitational systems do not exist. This is due to the hyperbolic structure of the Einstein equations; any dynamical solution loses energy through outward radiation. However as remarked by Alexakis-Schlue \cite{AS}, this is a subtle mathematical question. It is true at the linearised level when we consider, for example, the free wave equation $ \square \phi=0$ on a Minkowski background: Formally, an outgoing wave with vanishing radiation field has to be stationary; that is, if the function $\phi$ decays faster than $r^{-1}$ in the null-outgoing direction then it must be time-independent \cite{Friedlander}. This is no longer the case for a suitably perturbed wave operator $L = \square + V$ as remarked by Alexakis, Schlue and Shao in \cite{Alexakis-Shao}. This problem can also be phrased as a question of uniqueness of solutions for differential operators given boundary conditions. H{\"o}rmander, \cite{Hormander}, provided general conditions for the uniqueness property to hold across an hypersurface. In the context of an hyperbolic principal symbol, these conditions reduce to requiring the hypersurface to be pseudo-convex (see section \ref{section:Carleman estimates} for more details). Complementary to this is the work of Alinhac, \cite{Alinhac}, he showed that (generically) if one of H{\"o}rmander's conditions is violated then non-uniqueness of solutions across an hypersurface ensue. \\ In this paper we focus on the consequences of the lack of gravitational waves and radiation. Firstly, we give a brief account of their mathematical meaning. It is easier to start at the linear level: Consider a small perturbation of the Minkowski metric in flat spacetime. The perturbation then is required to satisfy the linearised Einstein equations around the trivial solution. After imposing suitable gauge conditions, it is found that the perturbation obeys a wave equation and therefore it behaves as a wave propagating causally on the background. Moreover, if we impose vanishing boundary conditions at infinity then the leading order term can be interpreted as power radiated to infinity. These perturbations are the so-called \emph{gravitational waves}. The space of solutions, of both linear Einstein and gauge equations, consists of the linear combination of two scalar waves, that is, gravitational waves have 2 degrees of freedom or polarisation modes. \\ Now, a description of gravitational waves in the non-linear case is more delicate. To the author's knowledge there is no way of splitting the gravity field into ``stationary'' and ``dynamic'' parts; this is due to the non-linear nature of Einstein equations. Whence the lack of meaning of the expression: Here is a gravitational wave (to be attributed to the dynamical part) propagating in a background (the stationary part). Nevertheless considerable efforts have been made to understand this statement and it is possible to make sense of it \emph{at infinity} by imposing \textit{asymptotically flat} boundary conditions\footnote{Alternatively, one could impose de Sitter or anti-de Sitter boundary conditions.}. In Section \ref{section:CK analysis} we review briefly the conclusions of the asymptotic analysis carried out by Christodoulou and Klainerman in order to prove the non-linear stability of Minkowski spacetime, \cite{CK}.\\ To the author's knowledge, it was Papapetrou, \cite{Papapetrou62},\cite{Papapetrou65}, who initiated the study of the relation radiation-stationarity for the full non-linear Einstein equations. Although his result is not conclusive, it provided strong evidence for the validity of the theorem presented here. More precisely, Papapetrou showed the incompatibility of the following two conditions: a) A spacetime is stationary below a characteristic hypersurface and non-radiative above it; b) There is a shock wave along that characteristic hypersurface or above it. \\ Later on, Bi\v{c}\'ak, Sholtz and Tod, \cite{BST}, gave a rigorous proof in the class of analytic metrics using ideas from Gibbons and Stewart, \cite{Gibbons-Stewart}. Bi\v{c}\'ak-Scholtz-Tod used the (undesired) hypothesis of analyticity all the way up to infinity to split Einstein equations order by order at infinity and then used the non-radiative condition to conclude time-independence to all orders inductively. Then the desired stationarity conclusion in the interior follows by analytic continuation. \\ The theorem was later improved to cover smooth metrics by Alexakis and Schlue in \cite{AS}. The purpose of this paper is to show that the result holds as well when gravity is coupled to electromagnetism. Alexakis and Schlue approach relies again on first proving stationarity to all orders at infinity as in \cite{BST}. Then, in order to extend this condition to the interior, they use unique continuation from infinity techniques based on Carleman estimates in the spirit of \cite{ASS}. For the first part of the proof, the exact null structure of the equations plays an important role in order to compute the metric, connection coefficients and curvature components to all orders at infinity with just the radiation field. For the second part a wave equation satisfied by the deformation tensor of the Weyl curvature, $\mathcal{L}_T C$, is derived and exploited to conclude stationarity in a neighbourhood of infinity by means of energy estimates. \\ We consider here the extended problem to include a Maxwell field. It is found that the results from \cite{AS} hold as well, that is, gravity and electromagnetism cannot balance each other to produce a periodic solution. The key point is that Maxwell equations also have a special non-linear structure and the coupled Einstein-Maxwell system can be treated in a similar way as before. The two deformation tensors, $\mathcal{L}_T C$ and $\mathcal{L}_T F$, also obey wave equations; however the coupling terms do not decay fast enough for the Alexakis-Schlue argument to work. Hence we are forced to revise and adapt their proof at the level of Carleman estimates to conclude the vanishing of the deformation tensors a little bit into the interior of the spacetime. The main result of this paper is, \begin{thm} \label{thm1} Let $(\mathcal{M}, g, F)$ be an asymptotically flat (in the sense of definition \ref{def:AF}) non-radiating solution of the Einstein-Maxwell equations. Then there exists a time-like vector field $T$ in a neighbourhood of spatial infinity such that \[ \mathcal{L}_T g = 0 = \mathcal{L}_T F . \] \end{thm} We comment briefly about the assumptions. In this paper we will consider the class of asymptotically flat spacetimes admitting coordinates $(t, r, \theta^i)$ which are Minkowski to leading order and the decaying metric coefficients admit an infinite asymptotic expansion in terms of inverse powers of $r$, an area parameter. Moreover, it is required that these expansions are well-behaved with respect to derivatives up to second order (cf. definition \ref{def:AF}). We remark that this condition of asymptotic flatness is morally equivalent to smoothness at null infinity in the conformal picture, the advantage of this formulation lies in its adaptability to more general asymptotic conditions, e.g., polyhomogeneous expansions \cite{Winicour}. Indeed, the methods employed here can be generalised to cope with time-independent logarithmic singularities; and while we take into account the asymptotic behaviour obtained by Klainerman-Christodoulou up to order three, it is a definite hypothesis of Theorem \ref{thm1} that we work with metrics which are smooth at higher orders. For more details about the smoothness assumption we refer the reader to the introduction of Section \ref{chapter:asymptotics}. \\ By non-radiating we mean that the Bondi mass is constant in a neighbourhood of spatial infinity. Recall the Bondi mass loss formula for future null infinity: \[ \partial_u M = - \frac{1}{32\pi} \int_{S^2} (|\Xi|^2 + |\underline{A}(F)|^2) d\mu_{\breve{\gamma}} , \] where $\Xi$ and $\underline{A}(F)$ can be regarded as the gravitational and electromagnetic power radiated to infinity per unit solid angle, respectively. Moreover, they correspond to the leading order components of the Weyl and Faraday tensors, respectively. Hence, the condition of constant Bondi mass is equivalent to requiring the vanishing of the gravitational and electromagnetic radiation fields, i.e., the leading order terms. In fact as we will see in Lemma \ref{lem:decay non-radiating}, the components of the Weyl and Faraday tensors for non-radiating spacetimes are $\mathcal{O}(r^{-3})$ and $\mathcal{O}(r^{-2})$, respectively. We refer the reader to Section \ref{section:CK analysis} for more details. \\ The conclusion of Theorem \ref{thm1} can be seen from two point of views. Firstly, as a rigidity result for metrics with fast decaying curvature. Secondly, it is also a inheritance of symmetry result since the gravity and electromagnetic field both turned out to be time-independent. As we will see shortly, this is not the case for other matter-energy models coupled to gravity. \\ We remark that the above result can be generalised to include a massless Klein-Gordon scalar field coupled with gravity and electromagnetism. A more precise statement accompanied with a sketched proof is presented in the Appendix. We stress that the fall-off conditions required for the whole system have to be assumed at this point due to the lack of a stability result for an Einstein-Klein-Gordon system. In Theorem \ref{prop:Klein-Gordon} we consider the fall-off obtained from the linear analysis performed in Winicour \cite{Winicour} together with a smoothness assumption at higher orders as in definition \ref{def:AF}.\\ Bi{\v{c}}{\'a}k-Scholtz-Tod also tackled this Einstein-Klein-Gordon problem in \cite{BST2} (a continuation of \cite{BST}). They again go around the fall-off condition by requiring analyticity of the fields all the way up to infinity and conclude that time-periodic asymptotically flat Einstein-massless-Klein-Gordon systems are in fact stationary. \\ The results presented in this paper are local around spatial infinity and hence are applicable to any system whose matter source is spatially compact. We mention here what is known about the case when the matter content extends to infinity. The simplest model to consider is that of a massive Klein-Gordon field coupled to Einstein equations. As mentioned above the techniques used in this work carry on to the Einstein-massless-Klein-Gordon system; however as we will see, the massive case is different. \\ In \cite{Herdeiro14}, Herdeiro and Rady constructed time-periodic Einstein-Klein-Gordon solutions. They start by analysing linear perturbations of the Klein-Gordon equation on a fixed Kerr background and take them as starting points to solve the full non-linear equations numerically. The underlying family of spacetimes are asymptotically flat, stationary, axisymmetric and with regular horizons. The scalar field however does not share the symmetries since the ansatz is taken to be $\phi(r,\theta)e^{i(m \varphi - \omega t)}$. Thus, this spacetimes can be regarded as hairy black holes bifurcating off the Kerr solution. Later on, Chodosh and Shlapentokh-Rothman, \cite{Chodosh}, retook the problem and proved, analytically, the existence of such spacetimes in a small neighbourhood of the Kerr family. All this is in contrast with our result where the symmetry is inherited by all the fields. The crucial difference in the assumptions is the presence of a non-zero mass for the scalar field, which acts as a non-decaying potential. \\ Related to the above hairy black holes is the existence of countably many time-periodic, spherically symmetric asymptotically flat \textit{boson stars} given by Bizo\'n and Wasserman in \cite{Bizon}. These are solutions of the Einstein-Klein-Gordon equations where the underlying spacetime is static while the complex scalar field has a positive mass and takes the form of a standing wave $\phi(r)e^{i\omega t}$. Therefore the matter field does not inherit the time-like symmetry. \\ \\ \textit{Outline of the paper.} In \textbf{Section 2} we introduce null coordinates adapted to future null infinity. We do this following the work of Christodoulou-Klainerman \cite{CK}, Christodoulou \cite{Christodoulou91} and Klainerman-Nicol{\`o} \cite{KN}. In addition to providing a background, these coordinates also give us a time-like quasi-symmetry which will play the role of the candidate Killing field. \\ Also, a tetrad adapted to these coordinates is defined and the structure equations are written with respect to this basis. Then the gauge conditions are explicitly spell out and we briefly explain the null hierarchy of the equations. To end this section we summarise the Christodoulou and Klainerman analysis based on the initial value formulation \cite{CK}. In particular their deduced asymptotic expansions and the concepts of mass and radiation are reviewed. \\ In \textbf{Section 3} we use the tetrad formalism to achieve the first step in the proof of the main theorem. That is, it is proven that a non-radiating spacetime is stationary to all orders at infinity. We recursively compute the connection and curvature coefficients to all orders at infinity (here is where the smoothness assumption of asymptotic flatness in definition \ref{def:AF} plays an important role in the form of asymptotic expansions in inverse powers of $r$). The hierarchy found by BMS, also interpreted as signature levels in the language of Christodoulou-Klainerman, helps to understand the limiting structure at future null infinity of the Einstein equations. Then the aforementioned hierarchy helps us identify levels where the equations become linear for the quantities belonging to that level. Moreover, the radiation fields can be regarded as the necessary initial data to run an induction argument and find recurrence relations. In particular, in the absence of radiation fields, all the asymptotic quantities are found to be time-independent. Also, the procedure sheds light on the way the different terms in the $r$-expansion of the metric and connection coefficients couple to each other via the Einstein equations. \\ The above results together with an analytical condition already imply the stationarity of non-radiating electrovacuum spacetimes in a neighbourhood of infinity as in \cite{BST}. However, one of the goals of this paper is to dispense with the analyticity assumption. We are able to remain in the smooth class by using the methods explained in Section 4 and conclude the stationarity of non-radiating spacetimes in the class of smooth metrics. \\ The technique to compute the asymptotic quantities to all orders at infinity is so basic that we feel compelled to give an account here of the main idea with a toy model: An out-going wave on Minkowski spacetime. The reader will find that the procedure is straightforward and the difficulty to apply it to the Einstein equations lies only in the intricacy of the equations themselves. \\ \\ \textit{Toy model.} Consider the Minkowski metric in spherical coordinates: \[ \eta = -dt^2 + dr^2+ r^2 \breve{\gamma} , \] where $\breve{\gamma}$ is the round metric on the sphere $\mathbb{S}^2$. We change to out-going null coordinates which are better suited for the problem at hand. That is, let $u=t-r$, then the metric can be written as \[ \eta = -du^2 - 2dudr + r^2 \breve{\gamma} . \] Note that the level sets of $u$ are null hypersurfaces ruled by $\partial_r$. Now, let $\phi$ be a solution of the free wave equation, $$\square_{\eta} \phi = 0. $$ Assume moreover that $\phi$ admits an expansion of the form\footnote{Such an expansion is morally equivalent to analyticity all the way to infinity; an assumption arguably incompatible with the wave equation. Here we assume it for the sake of simplicity. The role of the wave equation in this argument will be merely computational as opposed to evolutionary. An asymptotic expansion would work as well.} \[ \phi (u,r,\theta^2, \theta^3) = \sum_{n=1}^{\infty} \overset{(n)}{\phi} (u, \theta^2, \theta^3) \frac{1}{r^n} ,\] which is well-behaved with respect to derivatives.\\ We wish to find $\overset{(n)}{\phi}$ in terms of $\overset{(1)}{\phi}$ and $\overset{(n)}{\phi_0} := \lim_{u\rightarrow -\infty} \overset{(n)}{\phi}$. To do so we consider $X = \grad \phi$ and rewrite the wave equation as the 1st order system, \begin{align*} \curl X = 0 , \\ \divergence X = 0 . \end{align*} For easy comparison with the CK notation we define the null components of $X$ as \[ x:= X_r, \quad \underline{x}:=X_u, \quad \cancel{X} := \frac1r \breve{\grad} \, \phi \] For the time being let us denote by $\breve{\grad}$, $\breve{\divergence}$ and $\breve{\Delta}$ the gradient, divergence and Laplace operators on the unit round sphere, respectively. In coordinates, the previous equations read \begin{align*} \partial_r \cancel{X} &= \frac1r \breve{\grad} \, x -\frac1r \cancel{X} , \quad & \partial_u \cancel{X} &= \frac1r \breve{\grad} \, \underline{x} -\frac1r \cancel{X} , \\ \partial_r \underline{x} &= \partial_u x, \quad & \breve{\curl} \, \cancel{X} &= 0 , \\ \partial_u x &= - \partial_r \underline{x} + \frac{1}{r^2}\partial_r(r^2 x) + \frac1r \breve{\divergence} \, X . \end{align*} These equations (the first column) imply the following recurrence relations for the asymptotic quantities, \begin{align*} -(n-1) \overset{(n)}{\cancel{X}} &= \breve{\grad} \overset{(n)}{x}, \\ -n \overset{(n)}{\underline{x}} &= \partial_u \overset{(n+1)}{x}, \\ \partial_u \overset{(n+1)}{x} &= n\overset{(n)}{\underline{x}} -(n-2) \overset{(n)}{x} -\breve{\divergence} \, \overset{(n)}{\cancel{X}}. \end{align*} From these we can deduce an evolution equation for each $\overset{(n+1)}{x}$ in terms of lower order data, \begin{align*} 2\partial_u \overset{(n+1)}{x} = \frac{1}{n-1} \breve{\Delta} \overset{(n)}{x} - (n-2) \overset{(n)}{x} , \end{align*} which can be solved given initial data $\overset{(n+1)}{x_0}$ at $u= -\infty$. Note that the asymptotic expansion for $\phi$ implies $ \overset{(1)}{x}=0$ and $ \overset{(2)}{x}=-\overset{(1)}{\phi}$, in particular, we can compute $\overset{(n+1)}{x}$ inductively for $n>2$ assuming we know the radiation field $\overset{(1)}{\phi}$. \\ Therefore, we can compute $ \overset{(n)}{x}$, $ \overset{(n)}{\underline{x}}$ and $ \overset{(n)}{\cancel{X}}$ for all $n\in \mathbb{N}$ along future null infinity in terms of the radiation field $\overset{(1)}{\phi}$ and the ``pole moments at spatial infinity'' $ \overset{(n)}{x_0}$, $n\in \mathbb{N}$. $_{\blacksquare}$ \\ This toy model reflects perfectly well the asymptotic relations when the equations can be extended smoothly up to infinity. The analysis for Maxwell and Einstein equations is completely analogous albeit more technical. The previous relations will be used in the Appendix when we discuss the extension of the main theorem to include a massless Klein-Gordon scalar field. \\ In \textbf{Section 4} we complete the proof of the main theorem. We explain how the previous asymptotic values at infinity can be translated to a local result in the spacetime. That is, the condition of stationarity to all orders at infinity is extended to a neighbourhood of spatial infinity. \\ The main technical tool is that of Carleman estimates. These are a priori wave estimates for functions with fast decay (faster than any polynomial) towards infinity. They can also be thought as energy estimates adapted to time-like boundaries conditions where now we aim at controlling the bulk term while making the boundary terms vanish. A key ingredient in these inequalities is a 1-parameter weight function which accounts for the exponential rate of decay. With the help of this weight, it is possible to bound the $L^2$-norms of a function $\phi$ and its first derivatives in terms of $\square \phi$, provided that the boundary of the region under consideration satisfies the so-called pseudo-convex condition. \\ At this point we make a parenthesis to look for wave equations satisfied by the components of the deformation tensors $\mathcal{L}_T C$ and $\mathcal{L}_T F$. These are the Ionescu-Klainerman tensorial equations, \cite{IK}. We need to revise them in order to cope with a non-vacuum spacetime. These wave equations, written in suitable coordinates, in conjunction with the Carleman estimates are used to find $L^2$-bounds for the components of the deformation tensors and its first derivatives. These bounds depend on an arbitrarily large parameter which can be taken to infinity to conclude the vanishing of the functions in the interior. \\ Finally, an \textbf{Appendix} is included where we state a variation of Theorem \ref{thm1} to include a Klein-Gordon field, as well as the corresponding recurrence relations. We also include a brief review of the tetrad formalism employed in this paper, as well as the CK notation. Finally a table is included to compare the CK and NP notations. \section{Coordinates and gauge conditions} \label{section:tetrad and coordinates} \setcounter{footnote}{1} Here we construct the necessary out-going null coordinates in the context of asymptotically flat spacetimes. These are given by a foliation by null hypersurfaces complemented with a transversal coordinate given by the candidate Killing field, which is roughly speaking, a time-like symmetry to first order. The existence and construction of such foliation is taken from Christodoulou \cite{Christodoulou91}. \\ Let $(\Sigma_0, h,K)$ be a strongly asymptotically flat initial data set and consider an exhaustion of $\Sigma_0$ by balls $B_d$, where $d$ is an asymptotically flat radial coordinate on $\Sigma_0$. Within the corresponding Cauchy development define $C_d^-$, resp. $C_d^+$, to be the ingoing, resp. outgoing, null hypersurface consisting of geodesic rays emanating from $S_d =\partial B_d$. Fix some $S_0=\partial B_{d_0}$, thanks to Christodoulou-Klainerman \cite{CK}, we can choose it so that the generators of $C_0^{\pm} := C_{d_0}^{\pm}$ have no future end points. Finally let $S_0^* = C_0^+ \cap C_{d^*}^-$ for some $d^*$ large. \\ At $S_0^*$ we can define the future-directed null outgoing and ingoing normal vectors $L$ and $\underline{L}$, respectively, by the conditions: \[ \langle L, \underline{L}\rangle =-2 , \quad \langle L,L \rangle = \langle \underline{L}, \underline{L}\rangle = \langle L, X \rangle = \langle \underline{L}, X \rangle = 0, \quad X\in \mathcal{T}_pS_0^* . \] These conditions fix $L, \underline{L}$ up to rescaling by a function $a$ on $S_0^*$: \[ L \mapsto a L , \qquad \underline{L} \mapsto \frac1a \underline{L} . \] Next, $(L, \underline{L})$ are extended to $C_{d^*}^-$ as follows: Let $\tilde{\underline{L}}$ be the extension of $\underline{L}$ by geodesic flow, that is, \[ \nabla_{\tilde{\underline{L}}}\tilde{\underline{L}} = 0, \quad \textrm{along } C_{d^*}^- \text{, with} \quad \tilde{\underline{L}} = \underline{L} \quad \textrm{on } S_0^* .\] With the help of its affine function $t$ (that is, $\tilde{\underline{L}} \cdot t = 1$ on $C_{d^*}^-$ and $t=0$ on $S_0^*$) we define the \emph{retarded time} function $u$ on $C_{d^*}^-$ to be the area parameter on a level set of $t$, \[ u(t) := 2 (r^*_0 - r^*_t) , \] here $r^*_t$ is the area radius of $S_t^*$, the level set of $t$ on $C_{d^*}^-$. Now, we set $\underline{L}$ to be the velocity field of $u$ on $C_{d^*}^-$, that is, $\underline{L} = \frac{dt}{du}\tilde{\underline{L}}$ along the null geodesics ruling $C_{d^*}^-$. In particular, $\underline{L}(u)=1$.\\ Then, let $L$ be defined on $C_{d^*}^-$ as the null conjugate to $\underline{L}$ with respect to level sets of $u$, $S_u^*$: \[ \langle L, \underline{L}\rangle = -2, \quad \langle L, X \rangle = 0, \quad X\in \mathcal{T}_pS_u^* . \] Finally we extend $u$ by choosing a solution of the eikonal equation, \[ g^{\alpha \beta} \partial_{\alpha} u \partial_{\beta} u = 0 \] with the already prescribed value of $u$ along $C_{d^*}^-$. Let $C_u^+$ be the level sets of $u$, we extend $L$ by geodesics, \[ \nabla_L L = 0 , \quad \textrm{on the relevant open set.}\] Let $s$ be the affine parameter: $L\cdot s = 1$ and $s|_{S_u^*}=r_u^*$. \\ The above procedure also gives coordinates $(s,u, \theta^i)$, by choosing $(\theta^i)$ on $S_0^*$ then $u-$flowing them along $\underline{L}$ on $C_{d^*}^-$ to obtain coordinates $(u, \theta^i)$ and then $s-$flowing the latter along $L$ to cover an open set near infinity. \\ It worth noticing that the above construction is engineered so that $L$, $\underline{L}$ and the candidate Killing field $\partial_u$ are related to leading order in the usual way, that is, \begin{align*} \partial_u = \frac12 \left( L + \underline{L}\right) + \mathcal{O}(r^{-1}). \end{align*} \begin{figure}[h] \begin{center} \begin{tikzpicture}[scale=.8] \draw (-3,0) node {}; \fill (4,0) circle (2pt) node[right] {$i^0$}; \draw [line width = 1pt, dashed] (0,4) -- (4,0) node[pos=.5, above right]{$I^+$}; \draw [line width = 1pt, dashed] (0,-4) -- (4,0) node[pos=.5, below right]{$I^-$}; \draw [line width = 1.5pt, ] (-2,0) -- (4,0) node[pos=0, left]{$\Sigma_0$}; \draw (1,2) circle (3pt) node[left] {$S_0^*$}; \draw [line width = 1pt] (-1,0) -- (1.5,2.5) node[pos=0.3, left ]{$C_0^+$}; \draw [line width = 1.5pt] (0,3) -- (3,0) node[pos=0.8, left]{$C_{d^*}^-$}; \draw[->] [line width = 1pt] (7,1) -- (6,2) node[pos=1, above]{$\underline{L}$}; \draw[->] [line width = 1pt] (7,1) -- (8,2) node[pos=1, above]{$L$}; \draw [line width = 1pt, dashed] (7,-2) -- (6,-1) ; \draw[->] [line width = 1pt] (7,-2) -- (6.7,-.8) node[pos=1, above]{$u$}; \draw[->] [line width = 1pt] (7,-2) -- (8,-1) node[pos=1, above]{$s$}; \end{tikzpicture} \caption{Coordinates in a neighbourhood of future null infinity. The level sets of $u$ are the outgoing null hypersurfaces $C_u^+$ ruled by $L$. The level sets of $s$ are timelike.} \end{center} \end{figure} \subsection{Gauge Conditions} It will be convenient to work with a tetrad \begin{align} \label{tetrad} \begin{aligned} e_0 & := L = \partial_s, \\ e_1 & := \underline{L} = f^0\partial_s + f^1 \partial_u + f^i\partial_{\theta^i}, \\ e_i & := {h_i}^j \partial_{\theta^j} , \end{aligned} \end{align} satisfying \begin{eqnarray} \label{metric-matrix} \eta_{\mu\nu} := \langle e_{\mu}, e_{\nu} \rangle = \left( \begin{array}{ccc} 0 & -2 & 0 \\ -2 & 0 & 0 \\ 0 & 0 & \textrm{Id}_{2}\\ \end{array}\right) . \end{eqnarray} The above change of basis can be written compactly in terms of an \emph{orthonormalisation matrix}, \begin{align*} e_{\mu} = {h_{\mu}}^{a} \partial_{x^{a}} . \end{align*} That is, \begin{eqnarray*} {h_{\mu}}^{a} := \left( \begin{array}{cc} \begin{array}{cc} 1 & f^0 \\ 0 & f^1 \end{array} & \begin{array}{cc} 0 & 0 \\ 0 & 0 \end{array} \\ \begin{array}{cc} 0 & f^2 \\ 0 & f^3 \end{array} & \begin{array}{c} {h_{i}}^k \end{array} \end{array} \right) . \end{eqnarray*} The metric components are related to the orthonormalisation matrix by $g^{ab} = {h_{\mu}}^{a} \eta^{\mu \nu} {h_{\nu}}^{b}$, where $g^{ab}$ is the inverse matrix of $g_{ab} := \langle \partial_{x^{a}}, \partial_{x^{b}} \rangle$. Explicitly, \begin{eqnarray*} g^{ab} = \left( \begin{array}{cc} \begin{array}{cc} 0 & -\frac12 f^1 \\ -\frac12 f^1 & -f^1 f^0 + (f^2)^2 + (f^3)^2 \end{array} & \begin{array}{cc} 0 & 0 \\ f^i {h_i}^2 & f^i {h_i}^3 \end{array} \\ \begin{array}{cc} 0 \qquad \qquad \; & f^i {h_i}^2 \qquad \quad \\ 0 \qquad \qquad \; & f^i {h_i}^3 \qquad \quad \end{array} & \begin{array}{c} {h_{i}}^k \delta^{ij} {h_j}^l \end{array} \end{array} \right) . \end{eqnarray*} \textbf{Remark.} The function $f^1$ is constant and we can choose it so that $f^1=2$. This can be showed by first noting that $\partial_s \langle \partial_u, \partial_s \rangle = \langle \nabla_{\partial_s} \partial_u, \partial_s \rangle + \langle \partial_u, \nabla_{\partial_s} \partial_s \rangle = \langle \nabla_{\partial_u} \partial_s, \partial_s \rangle = \frac12 \partial_u \langle \partial_s, \partial_s \rangle = 0$, hence $\langle \partial_u, \partial_s \rangle$ is constant along $s-$lines. Then the assertion follows from $-2 =\langle L, \underline{L} \rangle = f^1 \langle \partial_u, \partial_s \rangle$. \\ We call all these choices the \emph{null-geodesic gauge conditions}. \\ The connection coefficients are the components of the derivative operator with respect to the tetrad $e_{\mu}$: \begin{eqnarray*} \omega_{\lambda \mu \nu} := \langle \nabla_{e_{\lambda}} e_{\nu}, e_{\mu} \rangle . \end{eqnarray*} They satisfy $\omega_{\lambda \mu \nu} = -\omega_{\lambda \nu \mu}$, this is a consequence of $\eta_{\mu \nu}$ being a constant matrix. Moreover, the connection coefficients and the orthonormalisation matrix are related by the \emph{frame equations} (the torsion-free property of the connection in this language, see Appendix for details). In the null gauge case they read \begin{align} \omega_{001} &= 0, \quad & -2\zeta_i = \omega_{10i} &= \omega_{i01} = -V_i, \\ \omega_{00i} &= 0 , \quad & -\chi_{ij}= \omega_{i0j} &= \omega_{j0i} = -\chi_{ji} , \\ -2\underline{\zeta}_i=\omega_{01i} &= \omega_{i10}=V_i , \quad & - \underline{\chi}_{ij} = \omega_{i1j} &= \omega_{j1i} = - \underline{\chi}_{ji}, \\ e_0(f^0) &= \frac12 \omega_{110} = 2 \underline{\omega}, \quad & e_i(f^0) &= \frac12 \omega_{11i} =\lambda_i, \\ e_0(f^i) &= ( {\omega_0}^j{}_1 - {\omega_1}^j{}_0 ){h_j}^i ,\quad & e_i({h_j}^l) - e_j({h_i}^l) &= ({\omega_i}^k{}_j - {\omega_j}^k{}_i ){h_k}^l, \\ e_0({h_i}^j) &= ( {\omega_0}^l{}_i - {\omega_i}^l{}_0 ){h_l}^j, \quad & e_1({h_i}^j) - e_i(f^j) &= ( {\omega_1}^l{}_i - {\omega_i}^l{}_i ){h_l}^j. \end{align} We refer the reader to the Appendix for a review of the CK notation and a comparison with the NP spin coefficients and the slight variations adopted in this paper. \\ The previous construction of coordinates and tetrad fixes $e_0=L$ and $e_1=\underline{L}$ up to rescaling \[ L \mapsto a L , \qquad \underline{L} \mapsto \frac1a \underline{L} , \] where $a$ is a function of the sphere $S^*_0$. In view of the corresponding scaling of the second fundamental forms, \[ \chi \mapsto a \chi , \qquad \underline{\chi} \mapsto \frac1a \underline{\chi} . \] The function $a$ can be chosen so that \[ \lim_{r \rightarrow \infty} ( \textrm{tr} \chi + \textrm{tr} \underline{\chi} ) = 0 . \] In other words, it is required that the first variation of area of $S_0^*$ along geodesics defined by the direction of $L+\underline{L}$ tends to $0$ at infinity. We define then \[T := \partial_u \] to be the candidate Killing vector field. \\ Finally, we can also arrange for $\omega_{0ij}=0$ as in Lemma \ref{lem:comoving gauge}. This is done by solving an ODE, so we still have the freedom to choose an initial condition for the system. This translates to a freedom to adapt the frame $\{e_i\}$ at infinity. More precisely, we are still free to choose the leading coefficient in the asymptotic expansion of ${h_i}^j$. Since we are approaching the round unit-sphere at infinity, we can set \begin{eqnarray*} \overset{(1)}{{h_i}^j} = \left( \begin{array}{cc} 1 & 0 \\ 0 & \frac{1}{\sin\theta} \\ \end{array}\right) . \end{eqnarray*} We call the choice $\omega_{0ij}=0$ the \emph{co-moving frame gauge}. \\ In the next section we work with the above choice of gauge conditions. While the null-geodesic gauge is essential for the procedure, the co-moving frame gauge is used mostly for convenience. The arguments go through provided that we think of $\omega_{0ij}$ and $ \overset{(1)}{{h_i}^j}$ as given data. \\ Note that the frame equations imply that $\chi_{ij}, \underline{\chi}_{ij}$ are symmetric 2-tensors, $2\zeta_i=V_i=-2\underline{\zeta}_i$, $\xi_i=0$ and $\omega=0$. The evolution of the orthonormalisation matrix is then given by \begin{align} e_0(f^0) &= 2 \underline{\omega}, \quad & e_i(f^0) &= \lambda_i, \\ e_0(f^i) &= -\zeta^k {h_k}^i ,\quad & e_i({h_j}^l) - e_j({h_i}^l) &= ({\omega_i}^k{}_j - {\omega_j}^k{}_i ){h_k}^l, \\ e_0({h_i}^j) &= -{\chi_i}^k{h_k}^j, \quad & e_1({h_i}^j) - e_i(f^j) &= ( {\omega_1}^l{}_i - {\omega_i}^l{}_i ){h_l}^j. \end{align} \subsection{Structure equations} Recall the formula for the Riemann curvature tensor components in terms of the connection coefficients: \[ R_{\rho \sigma \mu \nu} = e_{\rho} (\omega_{\sigma \mu \nu} ) -e_{\sigma} (\omega_{\rho \mu \nu} ) - {\omega_{\rho \mu}}^{\alpha} \omega_{\sigma \nu \alpha} + {\omega_{\sigma \mu}}^{\alpha} \omega_{\rho \nu \alpha} - {\omega_{\sigma \rho}}^{\alpha} \omega_{\alpha \mu \nu} + {\omega_{\rho \sigma}}^{\alpha} \omega_{\alpha \mu \nu} . \] The splitting into trace and traceless part is the usual \[ R_{\rho \sigma \mu \nu} = C_{\rho \sigma \mu \nu} + \frac{1}{2} ( \eta_{\rho \mu} S_{\sigma \nu} - \eta_{\sigma \mu} S_{\rho \nu} + \eta_{\sigma \nu} S_{\rho \mu} - \eta_{\rho \nu} S_{\sigma \mu} ) , \] where $C_{\rho \sigma \mu \nu}$ and $S_{\mu \nu}$ are the components of the Weyl and Schouten tensors, respectively. \\ The \emph{structure equations} are the combination of these two preceding equations, we write them schematically as \begin{align} \label{structure equations} C + \eta \wedge S = e \wedge \omega + \omega \wedge \omega . \end{align} The Einstein equations appear implicitly when we regard the Schouten tensor as the stress-energy tensor of the matter/energy content. \\ In order to understand the structure equations in the null-geodesic gauge we introduce the concept of the signature of a tensor, as discussed in \cite{CK} p 148. We just quote here their clear exposition: \\ \emph{Given any covariant tensor $U$ at a point of spacetime, we define a null component of it to be any tensor tangent to the sphere $S_{u,s}$ at a point, which is derived from $U$ by contractions with either $e_0$ or $e_1$ and projections to $S_{u,s}$. To any such component we assign a signature that it is defined as the difference between the total number of contractions with $e_0$ and the total number of contractions with $e_3$. We are now ready to state the following heuristic principle. \\ \\ \textbf{Principle of Conservation of signature:} Consider an arbitrary covariant tensor $U$ that can be expressed as a multilinear form in an arbitrary number of covariant tensors $U_1 \ldots U_p$, with coefficients depending only on the spacetime metric and its volume form. Then the signature of any null term of U, expressed in terms of the null components of $U_1\ldots U_p$, is equal to the sum of the signatures of each constituent in the decomposition.}\\ In particular the signature of a tensor does not change by lowering/raising indices. We have included in Table 1 of the Appendix the signature of the different null components relevant to us. The above principle will allow us to guess the quadratic terms appearing in the structure equations without performing the calculations. Indeed, we already know that in the null-geodesic gauge the only connection coefficient with signature equal 2 vanishes; this fact is the one responsible for the success of the hierarchy-induction argument explained in Section \ref{section:quantities} to obtain the asymptotic quantities to all orders at infinity.\\ Now we illustrate how to apply this principle to the structure equations. We will denote by $U(l)$ a null component of $U$ with signature $l$. The signature 2 case yields schematically: \begin{align*} C(2) + \eta \wedge S(2) &= e(1) \omega(1) + e(0)\omega(2) + \omega(1) \wedge \omega(1) + \omega(0) \wedge \omega(2), \\ &=e(1) \omega(1) + \omega(1) \wedge \omega(1) . \end{align*} where in the second line we have used the fact that the only connection coefficient with signature 2, $\omega_{0\al0}$, vanishes in the null-geodesic gauge. More generally we have the following: \\ \\ \emph{\textbf{Hierarchy of structure equations:} When working in the null-geodesic gauge there is no connection coefficient of signature $l-2$ on the right-hand side of the structure equation of signature $l$, $l=2,1,0$.}\\ This is precisely the hierarchy found by BMS, \cite{BMS}. The exact form of the structure equations is presented in equations (\ref{hierarchy0})-(\ref{hierarchy4}). \subsection{Christodoulou-Klainerman estimates } \label{section:CK analysis} In their book The Global Nonlinear Stability of the Minkowski Space \cite{CK}, Christodoulou and Klainerman proved that any strongly asymptotically flat initial data set that satisfies a global smallness assumption, leads to a unique globally hyperbolic and geodesically complete solution of the Einstein vacuum equations. Moreover the development is globally asymptotically flat in the sense that the curvature vanishes at infinity in all directions. \\ Here we list some conclusions of their work regarding the asymptotic behaviour of solutions suitably close to Minkowski spacetime. Of particular interest are the decay rates obtained for the connection coefficients and the null components of the Weyl tensor. It is observed that the peeling property may not hold in general. \\ While all the previous was done in the vacuum case, the extension to include an Electromagnetic field was worked out by Zipser in her PhD dissertation \cite{Zipser}. The same conclusion for an Einstein-massless-Klein-Gordon system is expected to hold; however to the author's knowledge such result is still lacking in the literature. \\ Using the null out-going coordinates as constructed in the previous section, Christodoulou-Klainerman showed that, for fixed $u$, $S_{s,u}$ converges to the standard round sphere embedded in Minkowski, that is, let $\cancel{g}$ be the induced metric on $S_{s,u}$, then \begin{align*} \lim_{u; r \rightarrow \infty} r^{-2} \cancel{g} &= \breve{\gamma}, \qquad & \lim_{u; r \rightarrow \infty} K[r^{-2} \cancel{g} ] &= 1, \\ \lim_{u; r \rightarrow \infty} r \tr \chi &= 2, \qquad & \lim_{u; r \rightarrow \infty} r \tr \underline{\chi} &=-2 , \end{align*} here the limit $\lim_{u; r \rightarrow \infty}$ is taken along a level set of $u$ while letting the area function $r$ tend to infinity; $\breve{\gamma}$ is the standard round metric on $\mathbb{S}^2$ and $K$ is the Gauss curvature.\\ Denote by $\hat{\chi}$ the trace-free part of $\chi$, then to next order they obtained \begin{align*} \lim_{u; r \rightarrow \infty} r^{2} \hat{\chi} &= \Sigma, \qquad && \lim_{u; r \rightarrow \infty} r(r\tr\chi - 2) = H, \\ \lim_{u; r \rightarrow \infty} r \underline{\hat{\chi}} &= \Xi, \qquad && \lim_{u; r \rightarrow \infty} r(r\tr\underline{\chi} - 2) = \underline{H} , \end{align*} where $\Sigma$, $\Xi$ are symmetric trace-less $u$-dependent 2-covariant tensors defined on $\mathbb{S}^2$ and $H$, $\underline{H}$ are $u$-independent functions on $\mathbb{S}^2$ of vanishing mean. Moreover, \begin{align*} |\Xi(u, \cdot)|_{\breve{\gamma}} \leq C(1 + |u|)^{-3/2}. \end{align*} Recall the null decomposition the Weyl tensor with respect to a tetrad $\{e_0, e_1, e_i \}$, \begin{align} \label{null Weyl} \alpha_{ij} &= C_{i0j0} , \qquad && \underline{\alpha}_{ij} = C_{i1j1} , \\ 2 \beta_i &= C_{i010} , \qquad && 2 \underline{\beta}_i = C_{i110} , \\ 4 \rho &= C_{1010} , \qquad && \sigma = C_{0123} . \end{align} The curvature null components also decay. Explicitly, there exist a symmetric trace-less 2-tensor $\underline{A}$, a 1-form $\underline{B}$ and functions $P$, $Q$ defined on $\mathbb{S}^2$ and $u$-dependent satisfying \begin{align} \label{null Weyl2} \lim_{u; r \rightarrow \infty} r \underline{\alpha} &= \underline{A}, \qquad & \lim_{u; r \rightarrow \infty} r^2 \underline{\beta} &= \underline{B}, \\ \lim_{u; r \rightarrow \infty} r^{3} \rho &= P , \qquad & \lim_{u; r \rightarrow \infty} r^3 \sigma &= Q . \end{align} Moreover, these limits decay in $u$ as follows, \begin{align} |\underline{A}(u, \cdot)| & \leq C (1+|u|)^{-5/2} , \qquad & |\underline{B} (u, \cdot)| &\leq C(1+|u|)^{-3/2}, \\ |P(u, \cdot) - \overline{P}(u)| & \leq (1+|u|)^{-1/2} , \qquad & |Q(u, \cdot) - \overline{Q}(u)| &\leq (1+|u|)^{-1/2}, \end{align} and $\lim_{u \rightarrow - \infty} \overline{P}(u)=0$, $\lim_{u \rightarrow - \infty} \overline{Q}(u)=0$.\\ This is all consistent with the presence of peeling. In contrast, they could only deduce a weaker fall-off for the remaining curvature components, namely, \begin{align*} |\alpha|_{\cancel{g}} = \mathcal{O} (r^{-7/2}), \qquad |\beta|_{\cancel{g}} = \mathcal{O} (r^{-7/2}), \end{align*} For completeness we mention that the presence of the full peeling property would be achieved by $|\alpha|_{\cancel{g}} = \mathcal{O} (r^{-5})$ and $|\beta|_{\cancel{g}} = \mathcal{O} (r^{-4})$. \\ Finally, the following relations are implied by the structure equations, \begin{align*} \partial_u \Xi &= -\frac12 \underline{A}, \qquad & \breve{ \cancel{\divergence} }\ \Xi &= \underline{B}, \\ \partial_u \Sigma &= -\frac12 \Xi, \qquad & \breve{ \cancel{\divergence} }\ \Sigma &= \frac12 \breve{\cancel{\nabla}} H + Z, \\ \partial_u \underline{H} &= -\frac12|\Xi|^2 , \end{align*} where $\breve{\cancel{\nabla}}$ and $\breve{\cancel{\divergence}}$ stand for the connection and divergence operators on the round sphere, respectively. \subsubsection{Mass} The \emph{Hawking mass} enclosed by a 2-sphere $S_{s,u}$ is defined to be, \begin{align*} m(s,u) = \frac{r(s,u)}{2} \left( 1 + \frac{1}{16\pi} \int_{S_{s,u}} \tr \chi \tr \underline{\chi} \right), \end{align*} Christodoulou and Klainerman also deduced from their analysis the decay rate of the Hawking mass as well as its evolution equation. Explicitly, there exists a function $M(u)$ such that along a level set of $u$ we have \begin{align*} m(s,u) = M(u) + \mathcal{O} (r^{-1}). \end{align*} The function $M(u)$ is called the Bondi mass. Furthermore, it satisfies the Bondi mass formula \begin{align*} \partial_u M = - \frac{1}{32\pi} \int_{\mathbb{S}^2} |\Xi(u, \cdot)|^2 d\mu_{\breve{\gamma}} . \end{align*} \subsubsection{Gravity coupled with electromagnetism} The previous analysis was done for vacuum solutions suitably close to Minkowski spacetime. The generalisation to an Einstein-Maxwell system was carried out by Zipser in her PhD thesis \cite{Zipser}. She found that the null components of the Faraday tensor decay as expected from the linear analysis (cf. \cite{CKlinear}). \\ Recall the null decomposition of the Faraday tensor. \begin{align} \label{null Faraday} \alpha(F)_i &= F_{i0} , \quad& \underline{\alpha}(F)_i &= F_{i1} , \\ \rho (F) &= \frac12F_{10}, \quad& \sigma(F) &= F_{23} . \end{align} As a consequence of Zipser's analysis we have that the following limits exist, \begin{align} \label{null Faraday2} \lim_{u; r \rightarrow \infty} r \underline{\alpha}(F) &= \underline{A}(F), \\ \lim_{u; r \rightarrow \infty} r^{2} \rho(F) &= P(F) , \\ \lim_{u; r \rightarrow \infty} r^2 \sigma (F) &= Q(F) . \end{align} However, full peeling is not available for the remaining component. Instead we have the weaker estimate along a level set of $u$, \begin{align} \label{null Faraday3} \alpha(F) = \mathcal{O}(r^{-\frac52}) . \end{align} The generalisation made by Zipser has as a consequence that the Hawking mass also has a limit, that we may call Bondi mass, and this limit satisfies the following Bondi mass formula: \begin{align} \label{Bondi mass formula} \partial_u M = - \frac{1}{32\pi} \int_{\mathbb{S}^2} |\Xi(u, \cdot)|^2 + |\underline{A}(F)(u, \cdot)|^2 d\mu_{\breve{\gamma}} . \end{align} For this reason we call $\Xi_{ij}$ and $\underline{A}(F)_i$ the \emph{radiation fields}. \begin{definition} \label{non-radiating} \emph{Non-radiating spacetimes.} An asymptotically flat solution to the Einstein-Maxwell equations is called non-radiating (towards future null infinity) if the Bondi mass $M(u)$ is constant as a function of retarded time $u$. \end{definition} This is strongly related to the order of decay of the curvature at infinity. Indeed we have the following lemma which is a consequence of Corollary \ref{cor:non-radiating}. \begin{lem} \label{lem:decay non-radiating} Assume we are working with the tetrad, coordinates and gauge conditions imposed in this section, then an asymptotically flat solution of the Einstein-Maxwell equations is non-radiating towards future infinity if and only if \begin{align*} C_{\rho \sigma \mu \nu} = \mathcal{O}(r^{-3}) \quad \textrm{and} \quad F_{\mu \nu} = \mathcal{O}(r^{-2}) , \end{align*} \end{lem} \section{Asymptotic behaviour of the fields} \label{chapter:asymptotics} \setcounter{footnote}{1} In this section we start by presenting the main technical assumptions required to `push' the Einstein equations to infinity at all orders. Namely we estate the class of functional spaces and asymptotically flat spacetimes that we will consider. Without further comment all the spacetimes in this paper will be assumed asymptotically flat in the sense of definition \ref{def:AF} below. \\ Next, we compute asymptotic quantities at infinity. We follow Bi\v{c}\'ak-Sholtz-Tod \cite{BST} and Alexakis-Schlue \cite{AS} in their asymptotic analysis and extent their results to show that the radiation fields determine the evolution of the metric to all orders at infinity. Indeed, by assuming an asymptotic expansion we can take the limit for each order of the structure equations and find algebraic and $u$-transport relations at infinity at a given order. These relations are well-suited for an induction process provided that we know the radiation fields, $\Xi_{ij}$ and $\underline{A}(F)_i$ along null infinity and initial conditions for $\overset{(2)}{\rho} $ and $\overset{(3)}{\zeta_i} $ at a sphere. The main results of this section state that these data determines the solution to all orders at infinity, see Proposition \ref{prop:recurrence relations} and its ensuing corollaries. \\ Before giving our definition of asymptotic flatness, we digress briefly about the different coordinate systems that we will be using. These are time-space, null out-going and double-null coordinates. \begin{itemize} \item \emph{Time-space coordinates.} These are important as they split naturally the initial data and the evolution equations. It is a definite hypothesis of this work that the asymptotic properties at spatial infinity of the Cauchy data are preserved under the evolution; this is the content of part $i)$ in definition \ref{def:AF}, below. It would be desirable to deduced this fact\footnote{That is, that smoothness at spatial infinity is preserved for finite but arbitrarily large time. Then, the desired expansion would hold for, e.g., time-periodic spacetimes.} from the Einstein equations; this would require a `preservation of regularity estimate for all orders at spatial infinity' which appears to be missing in the literature. \\ We stress here that the CK analysis is not suited for this problem as their gauge conditions are naturally adapted to null inifnity. On the other hand, Friedrich's analysis of his conformal equations, \cite{Friedrich02}, provides strong evidence about the preservation of regularity at spatial infinity, since the equations can be extended regularly there.\\ \item \emph{Null out-going coordinates.} Their construction is one of the crucial steps in the CK analysis. These coordinates are used to define the 15-parameter family of quasi-symmetries (corresponding to translations, Lorentz rotations, scaling and inverted translations in Minkowski spacetime) which serve to compare all geometric quantities to that of flat spacetime. Of particular importance to us is the construction of the time-like quasi-translations, which will play the role of candidate Killing field.\\ In \cite{CK} they define $u$ by setting initial conditions on a ``last Cauchy hypersurface'' and then extending by geodesics. Here we have followed Christodoulou \cite{Christodoulou91} and Klainerman-Nicol\`o \cite{KN} in that initial conditions for $u$ are set on a null hypersurface by means of the area radius parameter and then are extended by geodesics as before. With respect to this coordinate system the candidate Killing field is given by $T=\partial_u$. \item \emph{Double-null coordinates.} In order to capture the decaying properties of a function towards both past and future null infinities one is forced to use double-null coordinates. Indeed, Klainerman-Nicol\`o revised the methods of \cite{CK} using a double-null foliation which is better suited for the analysis of the structure equations. \\ The relevance of these coordinates in this paper relies on the fact that the Carleman estimates, Theorem \ref{thm:Carleman}, are stated with respect to double-null coordinates. It is important to stress, however, that the type of metrics allowed by definition \ref{def:AF} do not satisfy a priori the decay conditions, (\ref{metric AS}), required for the Carleman estimates to hold. However, in the context of non-radiating spacetimes, it can be proved that the Cauchy development of initial data close to Kerr-Newman which happens to be non-radiating do admit double-null coordinates of the desired form (\ref{metric AS}). We will prove this in Lemma \ref{lem:doublenull}. \end{itemize} \subsection{Asymptotic conditions} Now we present the functional spaces relevant for our class of asymptotically flat spacetimes. \\ We say that a function $f \colon \mathbb{R} \times [R, \infty) \times \mathbb{S}^2 \longrightarrow \mathbb{R}$ is in $\mathcal{O}^{\infty}_{k} (r^{-q})$, $q\geq 1$, if it admits an infinite asymptotic expansion in $r^{-n}$, $n=q,q+1,\ldots$, and this expansion is well-behaved with respect to derivatives up to order $k$. Formally we require that there exist functions\footnote{By the symbol $\overset{(n)}{f}$ we mean the function that achieves the desired approximation to $f$ at a given order, as opposed to the $n$th derivative of $f$} $\overset{(n)}{f}(t,\vartheta^i)\in C^k(\mathbb{R} \times \mathbb{S}^2)$ such that for all $t\in \mathbb{R}$, $r\geq R$, $\vartheta^i \in \mathbb{S}^2$ the following holds \[ f(t,r,\vartheta^i) \sim \sum_{n=0}^{\infty} \overset{(n-q)}{f} (t,\vartheta^i) r^{q-n} ,\] \[ \partial_t^l \partial^{\alpha}_{\vartheta} f(t,r,\vartheta^i) \sim \sum_{n=0}^{\infty} \partial_t^l \partial^{\alpha}_{\vartheta} \overset{(n-q)}{f} (t,\vartheta^i) r^{q-n}, \] \textrm{ where } $\partial^{\alpha}_{\vartheta} = \dfrac{\partial^{\alpha_1}}{\partial_{\vartheta^{1}}} \dfrac{\partial^{\alpha_2}}{\partial_{\vartheta^{2}}}, \, l + \alpha_1 + \alpha_2 \leq k$. \[ \partial^l_{r} f(t,r,\vartheta^i) \sim \sum_{n=0}^{\infty} \prod_{i=0}^{l-1} (q-n-i) \overset{(n-q+l)}{f} (t,\vartheta^i) r^{q-n-l}, \, \quad 1\leq l \leq k.\] Here $f \sim \sum_{n=0}^{\infty} \overset{(n-q)}{f} r^{q-n}$ means that for all $N\in \mathbb{N}$ there is $C_N>0$ so that \[ \big| f - \sum_{n=0}^N \overset{(n-q)}{f} r^{q-n} \big| \leq C_N r^{q-N-1} . \] \begin{definition} \label{def:AF} Throughout this paper we will call a Lorentzian manifold $(\mathcal{M},g)$ \emph{asymptotically flat} if the following conditions are satisfied: \begin{enumerate} [i)] \item \emph{Regularity at spatial infinity.} We say that $(\mathcal{M}, g)$ is regular at spatial infinity if there are coordinates $(t,r , \vartheta^i)$ such that for large $r$, the metric admits an expansion of the form \begin{eqnarray*} g = -dt^2 + dr^2 + r^2 \breve{\gamma}_{ij} d\vartheta^i d\vartheta^j + g^{\infty} \end{eqnarray*} where $\breve{\gamma}$ is the round metric on $\mathbb{S}^2$ and the components $g_{\alpha\beta}^{\infty}$ of $g^{\infty}$ belong to the class $\mathcal{O}^{\infty}_{2} (r^{-1})$. \item \emph{Smoothness at null infinity.} $(\mathcal{M},g)$ is called \emph{smooth at future null infinity} if, with respect to the null out-going coordinates $(u,s, \theta^2, \theta^3)$ previously defined, the vanishing of a component of a geometric quantity, $w = w (s, u, \theta^1, \theta^2)$, in the sense that \begin{equation} \label{vanish} \lim_{u; r \rightarrow \infty} r^q w = 0, \end{equation} implies $r^q \omega = \mathcal{O}(r^{-1})$ and the existence of the limit of $w$ to order $q+1$, in the sense that \[ \partial_s (r^{q+1} w) = \mathcal{O}(r^{-2}) . \] Moreover, we require a mild angular regularity condition, namely, if $w$ vanishes in the sense of (\ref{vanish}) then so do its angular derivatives: \[ \lim_{u; r \rightarrow \infty} r^q \partial_{\theta^i} w = 0 . \] Analogously, we require these conditions to hold at past null infinity. \item \emph{Weak peeling.} We also require that the Weyl tensor decays accordingly to \begin{align*} |\underline{\alpha} |_{\cancel{g}} &= \mathcal{O}(r^{-1}) \\ |\underline{\beta} |_{\cancel{g}} &= \mathcal{O}(r^{-2}) \\ |\rho|, |\sigma| &= \mathcal{O}(r^{-3}) \\ |\partial_u \beta |_{\cancel{g}} &= \mathcal{O}(r^{-4}) \\ |\partial_u \alpha|_{\cancel{g}} &= \mathcal{O}(r^{-5}) \end{align*} \item \emph{Compatibility of $I^+$ and $I^-$ at spatial infinity.} We say that the \emph{poles are finite towards spatial infinity} if all the components of the curvature $\kappa = \alpha, \beta, \rho, \sigma, \underline{\beta}, \underline{\alpha}$ admit an expansion for large $r$ (near future and past null infinities $I^+$, $I^-$), \[ \kappa \sim \sum_{n=0}^{\infty} \overset{(n+1)}{\kappa}_{\pm} (u_{\pm}, \xi) r^{-1-n}, \qquad I^{\pm} \simeq \{ (u_{\pm}, \xi): u_{\pm}\in \mathbb{R}, \xi \in \mathbb{S}^2 \} , \] with the property that \[ \lim_{u_{\pm} \rightarrow \mp \infty} | \overset{(n)}{\kappa_{\pm}} (u_{\pm} ,\xi)| < \infty , \qquad \forall n\in \mathbb{N}. \] \end{enumerate} \end{definition} These are all technical conditions which are sufficient to allow us to deal with the Einstein equations one order at a time at infinity. Some remarks are in order. \\ \\ \textbf{Remarks.} \begin{itemize} \item The first condition states that the initial regularity assumptions at spatial infinity are propagated through out the evolution. This property may not hold in general since logarithmic divergences are expected to appear when $t\rightarrow \infty$. However, it should be possible to prove its validity for arbitrarily large $t$ by using Friedrich's conformal equations, \cite{Friedrich02}. Then condition $i)$ would hold for time-periodic spacetimes. \item The purpose of condition $ii)$ is to rule out \emph{logarithmic singularities}. For example, the function $\frac{\log r}{r}$ decays at infinity but it is not $\mathcal{O}(r^{-1})$. It is known that these kind of singularities are present in Cauchy developments of asymptotically flat initial data \cite{Winicour} and actually are expected to be ubiquitous in dynamical spacetimes. However, as remarked by Christodoulou \cite{Christodoulou02} the coefficients accompanying the logarithmic singularities are time-independent for relevant systems, e.g., an $N$-body system with no incoming radiation. Therefore we can get rid off the logarithmic singularity by taking an extra time-derivative. In the context of non-radiating spacetimes taking an extra time-derivative does not affect the general argument thanks to the \emph{finiteness of poles condition}, condition $iv)$ which prohibits polynomial growth. \\ Summarising, at first glance it would appear that smoothness at null infinity is too strong an assumption; however, thanks to the CK analysis, we can circumvent the problem by noting that an extra time-derivative makes the geometric quantities smooth at null infinity. The same issue arises when discussing the Peeling property. It does not hold in general for all the components of the Weyl tensor, but when the failure is understood as being time-independent, an extra time-derivative allows to state the weak peeling property as in $iii)$ of the above definition.\\ Moreover, in the context of time-periodic spacetimes, it is easy to prove that the affine parameter $s$ is comparable to $r$ by appealing to a fundamental domain argument and condition $i)$. That is, in a time-periodic spacetime the regularity assumption at null infinity follows from the regularity condition at spatial infinity.\\ At this point it is also worth stressing that the $r$ function appearing in time-space coordinates is also compatible with the $r$ function appearing in double-null coordinates. This is a consequence of Lemma \ref{lem:doublenull}. Therefore, when discussing asymptotic properties, we can freely use any of these $r$ or $s$ as parameters along the outgoing null geodesics. \item Finally, there are dynamical spacetimes satisfying the above regularity assumptions. Indeed, Cutler and Wald, \cite{Cutler-Wald}, and Chru\'sciel and Delay, \cite{Chrusciel-Delay} provided examples of electrovacuum and vacuum spacetimes, respectively, having a smooth null infinity. \end{itemize} \subsection{Quantities to all orders at infinity} \label{section:quantities} Following the the work of Bi\v{c}\'ak-Sholtz-Tod \cite{BST} and Alexakis-Schlue \cite{AS} we now proceed to show that the radiation fields $\Xi_{ij}:=\overset{(1)}{\hat{\chi}_{ij}}$ and $\underline{A}(F)_i :=\overset{(1)}{\underline{\alpha}(F)_i} $ characterise the metric at infinity to all orders (up to stationary data to be described below). To do so we rely on the hierarchy discovered by BMS which is also tied to the splitting of the Weyl tensor into its null components and can be interpreted as signature levels. \\ Formally, in Proposition \ref{prop:recurrence relations} we state recurrence relations satisfied by the different orders of the physical quantities. This is done by translating the structure equations to infinity, hence obtaining a non-linear algebraic system of equations for the asymptotic coefficients. Then the aforementioned hierarchy helps us to identify levels where the equations become linear for the quantities belonging to that level. Moreover, the structure of that hierarchy leads to identifying the radiation fields as the necessary initial data to run an induction argument. \\ Then we further specialise to the case when $\Xi_{ij}=0$ and $A(F)_i=0$, that is, when no radiation is escaping to infinity (Definition \ref{non-radiating}). Then by running the hierarchy-induction procedure we prove in Corollary \ref{cor:non-radiating} that $\partial_u$ is a Killing symmetry to all orders at infinity. \\ We remark here that we will be using generically the word \emph{quantities} to refer to the components of either the orthonormalisation matrix, the connection coefficients or the curvature components. \\ Before stating the main results we set some notation. Recall that $\overset{(n)}{f}$ denotes the best $r^{-n}$-approximation of $f$ (see discussion preceding definition \ref{def:AF}). The symbol $\{\phi_1, \ldots, \phi_n\}$ will denote any expression involving the functions $\phi_1, \ldots, \phi_n$. The symbol $\lfloor n \rfloor $ denotes an expression involving the connection coefficients up to order $n$ and the orthonormalisation matrix up to order $n-1$. That is, $$ \lfloor n \rfloor = \{ \overset{(0)}{\omega_{\mu \nu \lambda}}, \ldots , \overset{(n)}{\omega_{\mu \nu \lambda}}, \overset{(0)}{{h_\mu}^{\nu}}, \ldots \overset{(n-1)}{{h_\mu}^{\nu}} \} , $$ with the convention that $\lfloor 0 \rfloor =0$. The symbol $Q(\phi,\ldots; \varphi, \ldots)$ stands for a quadratic expression containing terms of the form $\phi \varphi$. We remind the reader that asymptotic flatness is to be understood as stated in definition \ref{def:AF}. \\ \begin{prop} \label{prop:recurrence relations} Let $(M, g, F)$ be an asymptotically flat solution of the Einstein-Maxwell equations. Then the asymptotic quantities satisfy the following recurrence relations along the out-going null hypersurfaces $C_u^+$ for any $n\in \mathbb{N}$, \begin{subequations} \label{recursion0} \begin{align} \overset{(n+1)}{\alpha_{ij}} &= (n-1) \overset{(n)}{\chi_{ij}} - 2 \overset{(n)}{\alpha(F)_k} \overset{(1)}{\alpha(F)^k} \eta_{ij} + \lfloor n-1 \rfloor , \\ \overset{(n)}{{h_i}^j} &= \overset{(n)}{{\chi_i}^k} \overset{(1)}{{h_k}^j} + \lfloor n-1 \rfloor , \end{align} \end{subequations} \begin{subequations} \label{recursion1} \begin{align} \overset{(n+1)}{\beta_i}, \overset{(n)}{\rho(F)}, \overset{(n)}{\sigma(F)}, \overset{(n)}{\omega_{jji}}, \overset{(n)}{\zeta_i}, \overset{(n)}{f^i} &= \{ \overset{(n)}{\chi_{ij}}, \overset{(n)}{\alpha(F)_i}, \lfloor n-1 \rfloor \} , \quad n> 3 \end{align} \end{subequations} \begin{subequations} \label{recursion2} \begin{align} \overset{(n+1)}{\rho}, \overset{(n+1)}{\sigma}, \overset{(n)}{\underline{\alpha}(F)}, \overset{(n)}{\underline{\omega}}, \overset{(n)}{\omega_{123}}, \overset{(n)}{\underline{\chi}_{ij}}, \overset{(n)}{\lambda_i}, \overset{(n-1)}{f^0} &= \{ \overset{(n)}{\chi_{ij}}, \overset{(n)}{\alpha(F)_i}, \lfloor n-1 \rfloor \} , \quad n> 2 . \end{align} \end{subequations} Moreover, \begin{subequations} \label{recursion3} \begin{align} 2\partial_u \overset{(n+1)}{\chi_{ij}} &= -n \overset{(n)}{\underline{\chi}_{ij}} + \{ \overset{(n)}{\chi_{ij}}, \overset{(n)}{\alpha(F)_i}, \lfloor n-1 \rfloor \} , \\ 2\partial_u \overset{(n+1)}{\alpha(F)_i} &= \{ \overset{(n)}{\alpha(F)_i} , \overset{(n)}{\rho(F)}, \overset{(n)}{\sigma(F)}, \lfloor n \rfloor \}, \end{align} \end{subequations} \begin{subequations} \label{recursion4} \begin{align} \overset{(n+1)}{\underline{\beta}_i} &= \partial_u \overset{(n+1)}{\zeta_i} + \lfloor n \rfloor , \\ \overset{(n+1)}{\underline{\alpha}_{ij}} &= \partial_u \overset{(n+1)}{\underline{\chi}_{ij}} + \lfloor n \rfloor . \end{align} \end{subequations} \end{prop} This proposition can be interpreted as saying that the asymptotic quantities can be computed recursively starting from the radiation fields. The cases $n=1,2,3$ are special. Roughly speaking $n=1$ corresponds to the choice of gauges at infinity. On the other hand, for $n=2$ we have to specify a mass aspect function as data at the $(\rho, \sigma)$-level; also an EM-charge aspect function at the $\beta_i$-level. Finally, for $n=3$ at the $\beta_i$-level an angular momentum aspect vector is required. It is worth remarking that these functions are not freely specifiable but they obey constraint/evolution equations given by the Einstein and Maxwell equations of signature $-1$ and $-2$. We deal with these cases during the proof of Proposition \ref{prop:recurrence relations}. Here we state the main relations found there in: \begin{lem} \label{lem:evolution equations} Under the same hypothesis of Proposition \ref{prop:recurrence relations} the following relations hold, \begin{subequations} \label{mass and angular momentum} \begin{align} \partial_u \tr \overset{(2)}{\underline{\chi}} &= Q(\overset{(1)}{\underline{\chi}}; \overset{(1)}{\underline{\chi}}) + Q(\overset{(1)}{\underline{\alpha}(F)}; \overset{(1)}{\underline{\alpha}(F)}), \\ \overset{(3)}{\rho}, \overset{(3)}{\sigma}, \overset{(2)}{\underline{\alpha}(F)}, \overset{(2)}{\underline{\omega}}, \overset{(2)}{\omega_{123}}, \overset{(2)}{\hat{\underline{\chi}}_{ij}}, \overset{(2)}{\lambda_i} , \overset{(1)}{f^0} &= \{ \overset{(2)}{\chi_{ij}}, \overset{(2)}{\alpha(F)_i}, \tr \overset{(2)}{\underline{\chi}}, \lfloor 1 \rfloor \} , \\ \overset{(3)}{\omega}_{jji} &= \overset{(1)}{h^{-1}} \cdot ( \breve{e}_k \overset{(3)}{{h_i}^j} + \lfloor 2 \rfloor ), \\ \partial_u\overset{(3)}{\zeta_i} &= - \partial_u\overset{(3)}{\omega_{jji}} + Q(\overset{(1)}{{ \underline{\alpha}(F)}}; \overset{(2)}{\rho(F)}, \overset{(2)}{\sigma(F)}) + \lfloor 2 \rfloor , \\ 2 \partial_u\overset{(2)}{\rho(F)} &= \breve{\cancel{\divergence}} \, \underline{A}(F) , \\ 2 \partial_u\overset{(2)}{\sigma(F)} &= \breve{\cancel{\curl}} \, \underline{A}(F) . \end{align} \end{subequations} \end{lem} Here $\breve{e}_k := \overset{(1)}{{h_k}^j} \partial_j$ is the standard orthonormal basis on the round sphere and $\breve{\cancel{\divergence}}$ and $\breve{\cancel{\curl}}$ are the corresponding operators. We remark that equations (\ref{mass and angular momentum}a), (\ref{mass and angular momentum}d), (\ref{mass and angular momentum}e) and (\ref{mass and angular momentum}f) can be interpreted as evolution formulas for the mass, angular momentum and EM-charges. From the point of view of a characteristic initial value formulation, these are constraint equations at null infinity. \\ In particular, combining these relations with the non-radiating condition we get stationarity to all orders at infinity: \\ \begin{cor} \label{cor:non-radiating} Let $(M, g, F)$ be an asymptotically flat electrovacuum spacetime. Assume it is non-radiating, that is, $\Xi_{ij}=0=\underline{A}(F)_{i}$. Then all the asymptotic quantities are $u$-independent. \\ \end{cor} Along the same lines we can prove a related result, namely, that all the asymptotic quantities can be found recursively in terms of the radiation fields and the ``pole moments'' of the initial Cauchy data. In other words, the radiation fields determine completely the dynamics of the spacetime at infinity. \\ \begin{cor} Let $(M, g, F)$ be an asymptotically flat electrovacuum spacetime. Then all the asymptotic quantities depend solely on the radiation fields $\Xi_{ij}$, $\underline{A}(F)_{i}$, the initial mass and angular momentum and the pole moments of $\chi_{ij}$ and $\alpha(F)_i$ (The latter are the limits as $u\rightarrow -\infty$ of the asymptotic components $\overset{(n)}{\chi_{ij}}$ and $\overset{(n)}{\alpha(F)_i}$, $n \in\mathbb{N}$). \end{cor} Before proving these results we present the structure equations below. We have grouped them in levels representing the hierarchy found first by BMS, \cite{BMS}, \cite{Sachs}. Note that (\ref{hierarchy0}b), (\ref{hierarchy1}e), (\ref{hierarchy1}f), (\ref{hierarchy2}i) and (\ref{hierarchy2}j) are frame equations while (\ref{hierarchy1}g), (\ref{hierarchy2}k) are 1st Bianchi identities. All of these are satisfied on any Lorentzian manifold. In contrast, the Einstein equations appear implicitly when we regard $S_{\mu\nu}$ as a quadratic expression in $F_{\mu \nu}$. Finally, (\ref{hierarchy1}c), (\ref{hierarchy1}d), (\ref{hierarchy2}g), (\ref{hierarchy2}h), (\ref{hierarchy3}c), (\ref{hierarchy3}d) are the Maxwell equations. \begin{enumerate} [a)] \item The $\alpha_{ij}$ or $\Psi_0$ or signature $2$ level. This includes the 2nd fundamental form $\chi_{ij}$ and ${h_i}^j$: \begin{subequations} \label{hierarchy0} \begin{align} \alpha_{ij} + \frac12 \eta_{ij} S_{00} &= -e_0(\chi_{ij}) - \chi_{ik} {\chi^k}_j , \\ e_0({h_i}^j)& = -{\chi_i}^k {h_k}^j. \end{align} \end{subequations} \item The $\beta_i$ or $\Psi_1$ or signature $1$ level. This includes the torsion $\zeta_i$ and the coefficients of the induced connection on $S_{s,u}$, that is, $\omega_{iij}$, and $f^i$: \begin{subequations} \label{hierarchy1} \begin{align} 2\beta_i + S_{0i} &= -2 e_0(\zeta_i) - 4 \zeta^k \chi_{ki} , \\ \beta_i - \frac12 S_{0i} &= e_0(\omega_{jji}) - \zeta_j\chi_{ii} + \chi_{jj}\zeta_i + {\chi_j}^k\omega_{kji} , \quad i\neq j , \\ e_0(\rho(F)) &= - \cancel{\divergence} \, \alpha(F) - \tr \chi \rho(F) - \zeta_i\alpha(F)^i , \\ e_0(\sigma(F)) &= - \cancel{\curl} \, \alpha(F) - \tr \chi \sigma(F) + \varepsilon^{ik}\zeta_i\alpha(F)_k , \\ e_i({h_j}^k) - e_j({h_i}^k) &= ({\omega_i}^n{}_j - {\omega_j}^n{}_i) {h_{n}}^k, \\ e_0(f^i) &= -\zeta^k {h_k}^i , \\ e_1(\omega_{jji}) - e_j(\omega_{1ji}) &= -e_j(\underline{\chi}_{ji}) + e_i(\underline{\chi}_{jj}) \nonumber \\ & \qquad - (\omega\wedge\omega)_{ji1j} + (\omega \wedge \omega)_{1jji} , \quad i\neq j . \end{align} \end{subequations} with no summation on repeated $j$'s. \item The $(\rho, \sigma)$ or $\Psi_2$ or signature $0$ level. This includes $\underline{\omega}$, $\omega_{123}$, $\underline{\chi}_{ij}$, $\lambda_i$ and $f^0$: \begin{subequations} \label{hierarchy2} \begin{align} -\rho + \frac12(S_{22} + S_{33}) &= e_2(\omega_{323}) - e_3(\omega_{223}) + \frac12 \chi_{22} \underline{\chi}_{33} + \frac12 \chi_{33} \underline{\chi}_{22} \nonumber \\ & \qquad - \chi_{23} \underline{\chi}_{23} - (\omega_{232})^2 - (\omega_{323})^2 , \\ - \sigma &= -e_2(\zeta_3) + e_3(\zeta_2) + \omega_{323}\zeta_3 - \omega_{232}\zeta_2 \nonumber \\ & \qquad - {\chi_2}^k \underline{\chi}_{3k} + {\chi_3}^k \underline{\chi}_{2k} , \\ \rho +\frac12 S_{01} &= e_0(\underline{\omega}) - 3\zeta^k\zeta_k, \\ - 2 \sigma &= e_0(\omega_{123}) + 4 \zeta^k \omega_{k23} , \\ -4 \rho - S_{jj} + \frac12 S_{01} &= -e_0(\underline{\chi}_{jj}) - 2 e_j(\zeta_j) + 2 \zeta^k \omega_{jkj} - {\chi_j}^k\underline{\chi}_{kj} - 2\zeta_j \zeta_j, \\ - \sigma - S_{23} &= -e_0(\underline{\chi}_{23}) - 2 e_2(\zeta_3) + 2 \zeta^k\omega_{2k3} - 2 \zeta_2\zeta_3 - {\chi_2}^k\underline{\chi}_{k3} , \\ e_0(\underline{\alpha}(F)_i) &= -e_i(\rho(F)) + {\varepsilon_i}^je_j(\sigma(F)) - (F\wedge \omega)_{i} , \\ e_1(\alpha(F)_i) &= e_i(\rho(F)) + {\varepsilon_i}^k e_k(\sigma(F)) - ({}^* F\wedge \omega)_{i} , \\ e_0(f^0) &= \underline{\omega}, \\ e_i(f^0) &= \lambda_i , \\ e_1(\chi_{ij}) -e_i(\zeta_j) &= e_0(\underline{\chi}_{ij}) - e_j(\zeta_i) - (\omega \wedge \omega)_{1i0j} + (\omega \wedge \omega)_{0j1i} . \end{align} \end{subequations} With no summation on repeated $j$'s. \item The $\underline{\beta}_i$ or $\Psi_3$ or signature $-1$ level: \begin{subequations} \label{hierarchy3} \begin{align} 2 \underline{\beta}_i - S_{1i} &= e_0(\lambda_i) +2 e_1(\zeta_i) - (\omega\wedge \omega)_{01i1} \\ - \underline{\beta}_i - \frac12 S_{1i} &= e_1(\omega_{jji}) - e_j(\omega_{1ji}) - (\omega \wedge \omega)_{1jji} . \\ e_1(\rho(F)) & = \cancel{\divergence} \, \underline{\alpha}(F) + \tr \underline{\chi} \rho(F) + \zeta_i\alpha(F)^i , \\ e_1(\sigma(F)) & = -\cancel{\curl} \, \underline{\alpha}(F) - \tr \underline{\chi} \sigma(F) + \varepsilon^{ik}\zeta_i\alpha(F)_k , \end{align} \end{subequations} With no summation on repeated $j$'s. \item Finally, the $\underline{\alpha}_{ij}$ or $\Psi_4$ or signature $-2$ level: \begin{align} \label{hierarchy4} \underline{\alpha}_{ij} + \frac12 \eta_{ij}S_{11} = -e_i(\lambda_i) - e_1(\underline{\chi}_{ij}) - (\omega\wedge \omega)_{i1j1}. \end{align} \end{enumerate} Now we proceed to prove the results stated above. It is convenient to think of the structure-Einstein-Maxwell equations as a 1st order system schematically of the form: \begin{align*} \nabla h + \omega \wedge h &= 0 , \\ \nabla \omega + \omega \wedge \omega &= C + \eta \wedge S , \\ S &= F^2, \\ \nabla F + \omega \wedge F &= 0 . \end{align*} The procedure will be analogous to the one described in the Introduction for a wave equation. \\ \textbf{Proof of Proposition \ref{prop:recurrence relations}.} Firstly, we state the 1st order values of the relevant coefficients that we will be using to obtain the recurrence relations and also in the induction argument. These are a consequence of the CK analysis and the gauge conditions. \begin{align*} \overset{(0)}{f^0} &= 1 , \qquad f^1 \equiv 2 , \qquad \overset{(1)}{{h_i}^j} = \mathrm{diag}(1 , \frac{1}{\sin \theta}), \\ \overset{(1)}{\underline{\chi}_{ij}} & = \Xi_{ij} + \eta_{ij}, \qquad \overset{(1)}{\chi_{ij}} = \eta_{ij}. \end{align*} Recall the Einstein equations: $$ S_{\mu \nu } = 2 F_{\mu \sigma} {F^{\sigma}}_{\nu} - \frac12 \eta_{\mu \nu} F_{\alpha \beta} F^{\alpha \beta} . $$ They are quadratic in $F_{\mu \nu}$, in particular then $\overset{(n+1)}{S_{\mu \nu}} = \{ \overset{(1)}{F_{\mu \nu}}, \ldots, \overset{(n)}{F_{\mu \nu}} \}$. A more detailed null decomposition is given by \begin{align*} S_{00} &= 2 \alpha(F)_i \alpha(F)^i = \mathcal{O} (r^{-5}), \\ S_{01} &= \rho(F)^2 + \sigma(F)^2 = \mathcal{O} (r^{-4}), \\ S_{ij} &= Q (\rho(F), \sigma(F) ; \rho(F), \sigma(F)) + Q(\alpha(F); \underline{\alpha}(F)) = \mathcal{O}(r^{-\frac72}) , \\ S_{0i} &= Q (\alpha(F)_i ; \rho(F), \sigma(F)) = \mathcal{O}(r^{-\frac92}), \\ S_{1i} &= Q (\underline{\alpha}(F)_i ; \rho(F), \sigma(F)) = \mathcal{O}(r^{-3}) \\ S_{11} &= 2 \underline{\alpha}(F)_i \underline{\alpha}(F)^i = \mathcal{O}(r^{-2}) , \end{align*} where we have included the peeling behaviour of the Faraday tensor, (\ref{null Faraday2})-(\ref{null Faraday3}). Now we proceed to derive the recurrence relations. \begin{enumerate} [a)] \item It can be seen that the $(n+1)$-order of the $\alpha$-level equations together with $$ \overset{(n+1)}{S_{00}} = \overset{(n)}{\alpha(F)_i} \overset{(1)}{\alpha(F)^i} + \{ \overset{(1)} \alpha(F)_i, \ldots, \overset{(n-1)} \alpha(F)_i \} $$ imply the recurrence relations (\ref{recursion0}) for $\overset{(n+1)}{\alpha_{ij}}$ and $\overset{(n)}{{h_i}^j}$. \item Now we look at the $(n+1)$-order of the $\beta$-level equations. From this level onwards the equations become more intricate. The important thing to remember is that we only need to keep track of the coefficients accompanying the variables that we want to find at a given order and level. \\ We start with the $(n+1)$-order of the frame equation: $$e_i({h_j}^k) - e_j({h_i}^k) = ({\omega_i}^n{}_j - {\omega_j}^n{}_i) {h_{n}}^k . $$ Note that the left-hand-side does not contain terms with $\overset{(n+1)}{{h_j}^k}$ since $e_i = {h_i}^j\partial_j$ and ${h_i}^j$ decays. Writing down the equations explicitly for $k=2,3$ we get the system \begin{eqnarray*} \left( \begin{array}{cc} \overset{(1)}{{h_2}^2} & - \overset{(1)}{{h_2}^3}\\ -\overset{(1)}{{h_2}^2} & \overset{(1)}{{h_2}^3} \end{array}\right) \left( \begin{array}{c} \overset{(n)}{\omega_{223}} \\ \overset{(n)}{\omega_{332}} \end{array}\right) = \breve{e}_k \overset{(n)}{{h_i}^j} + \lfloor n-1 \rfloor . \end{eqnarray*} Here we recall that $\breve{e}_k := \overset{(1)}{{h_k}^j} \partial_j$ is the standard orthonormal basis on the unit round sphere. This system tells us that $\overset{(n)}{\omega_{jji}} = \{ \breve{e}_k \overset{(n)}{{h_i}^j}, \lfloor n-1 \rfloor \} = \{ \breve{e}_k \overset{(n)}{\chi_{ij}}, \lfloor n-1 \rfloor \}$, where we have used the recurrence relation for $\overset{(n)}{{h_i}^j}$. In particular for $n=3$ we get equation (\ref{mass and angular momentum}c).\\ We are left with the variables $\beta_i$, $\rho(F)$, $\sigma(F)$ and $\zeta_i$. The $(n+1)$-order of equations (\ref{hierarchy1}a)-(\ref{hierarchy1}d) give the following linear system valid for $n\geq 3$, \begin{eqnarray*} \left( \begin{array}{cccc} 2 & 0 & 0 & -(2n-4) \\ 1 & 0 & 0 & -1 \\ 0 & -(n-2) & 0 & 0 \\ 0 & 0 & -(n-2) & 0 \end{array}\right) \left( \begin{array}{c} \overset{(n+1)}{\beta_i} \\ \overset{(n)}{\rho(F)} \\ \overset{(n)}{\sigma(F)} \\ \overset{(n)}{\zeta_i} \end{array}\right) = \{ \overset{(n)}{\chi_{ij}},\overset{(n)}{\alpha(F)_i}, \lfloor n-1 \rfloor \} , \end{eqnarray*} where the zeros on the first two rows correspond to the fact that $\overset{(1)}{\alpha(F)} = 0 = \overset{(2)}{\alpha(F)} $ and the quadratic character of $S_{0i}$ stated before. Similarly the zeros on the last column come also from $\overset{(1)}{\alpha(F)} = 0 = \overset{(2)}{\alpha(F)}$. The above system can be solved (note that the degenerate cases correspond to $n =2,3$) to obtain the desired recurrence relations. \\ It is worth noticing that $\overset{(1)}{\zeta_i} = 0$, while $\overset{(1)}{\omega_{jji}}$ corresponds to the connection coefficients of the standard round sphere.\\ \\ \textbf{Remark.} The case $n=2,3$ are special. For $n=2$ the Maxwell equations degenerate and we have to specify $\overset{(2)}{\rho(F)}$ and $\overset{(2)}{\sigma(F)}$ as initial data, these can be regarded as electromagnetic charges. For $n=3$, the degeneracy is telling us that we have to prescribe $\overset{(3)}{\zeta_i}$ or $\overset{(4)}{\beta_i}$ as initial data. However, these data are not freely specifiable, in the sense that they satisfy evolutions equations (\ref{mass and angular momentum}). We will find the same situation on the $(\rho,\sigma)$-level where the corresponding initial data can be regarded as a mass aspect function. \item This is the 0-signature case; a signature count gives us that $\lambda_i$ does not appear on the $(\omega \wedge \omega)$-term. Thus, the $(n+1)$-order of equations (\ref{hierarchy2}a)-(\ref{hierarchy2}g) give us 9 equations for 9 variables, namely, $\overset{(n+1)}{\rho}$, $\overset{(n+1)}{\sigma}$, $\overset{(n)}{\alpha(F)_i}$, $\overset{(n)}{\underline{\omega}}$, $\overset{(n)}{\omega_{123}}$ and $\overset{(n)}{\underline{\chi}_{ij}}$. The linear system can be solved provided it is non-degenerate, which can be checked by direct computation for $n>2$. This gives the desired recurrence relation. Here we state the linear system obtained by considering the $(n+1)$-order of equations (\ref{hierarchy2}a)-(\ref{hierarchy2}g): \begin{eqnarray*} \left( \begin{array}{cccccccc} 1 & 0 & \frac12 & \frac12 & 0 & 0 & 0 & * \\ 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \\ 1 & 0 & 0 & 0 & 0 & n & 0 & * \\ 0 & 1 & 0 & 0 & 0 & 0 & -n & 0\\ 1 & 0 & n-1 & 0 & 0 & 0 & 0 & * \\ 1 & 0 & 0 & n-1 & 0 & 0 & 0 & *\\ 0 & 1 & 0 & 0 & 0 & 0 & n-1 & *\\ n-1 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \end{array}\right) \left( \begin{array}{c} \overset{(n+1)}{\rho} \\ \overset{(n+1)}{\sigma} \\ \overset{(n)}{\underline{\chi}_{22}} \\ \overset{(n)}{\underline{\chi}_{33}} \\ \overset{(n)}{\underline{\chi}_{23}} \\ \overset{(n)}{\underline{\omega}} \\ \overset{(n)}{\underline{\omega}_{123}} \\ \overset{(n)}{\underline{\alpha}(F)_i} \\ \end{array}\right) = \textrm{Known data} , \end{eqnarray*} where `Known data' can be described more precisely as a term of the form $$ \{ \overset{(n)}{\chi_{ij}}, \overset{(n)}{\zeta_i},\overset{(n)}{\omega_{jji}}, \overset{(n)}{\alpha(F)_i}, \overset{(n)}{\rho(F)},\overset{(n)}{\sigma(F)}, \lfloor n-1 \rfloor \} .$$ Once again the case $n=2$ is special. The 1st, 5th and 6th rows become linearly dependent. Due to this degeneracy we have to specify $\tr \overset{(2)}{\underline{\chi}}$ along null infinity, this corresponds to the Bondi mass (up to gauge terms). Finally we remark that the Einstein equation at the $\underline{\alpha}$-level provides us with an evolution equation, (\ref{mass and angular momentum}a), for $\tr \overset{(2)}{\chi}$, which corresponds to the Bondi mass formula (\ref{Bondi mass formula}). \item Finally, equations (\ref{recursion3}) are obtained by taking the $(n+1)$-order of equations (\ref{hierarchy2}k) and (\ref{hierarchy2}h), respectively. \item Now we focus on Lemma \ref{lem:evolution equations}. The evolution equations for the EM-charges, (\ref{mass and angular momentum}e) and (\ref{mass and angular momentum}f), follow from Maxwell's equations (\ref{hierarchy3}c) and (\ref{hierarchy3}d), respectively. \item As explained before we can obtain an evolution equation for $\overset{(3)}{\zeta_i}$ by considering the system $(\ref{hierarchy3}a)$ and $(\ref{hierarchy3}b)$. The equation obtained after getting rid of the Weyl terms is of the form: $$ \partial_u\overset{(3)}{\zeta}_i = - \partial_u \overset{(3)}{\omega_{jji}} + Q(\overset{(1)}{\underline{\chi}}; \overset{(2)}{\zeta}) + Q(\overset{(1)}{\underline{\alpha}(F)};\overset{(2)}{\rho(F)}, \overset{(2)}{\sigma(F)}) + \lfloor 2 \rfloor . $$ This proves the identity (\ref{mass and angular momentum}d). \item Again a signature analysis as above helps us to find the structure of the $\underline{\alpha}$-equation. We have that to 2nd order the trace of (\ref{hierarchy4}) reads $$ \partial_u \tr \overset{(2)}{\underline{\chi}} = Q(\overset{(1)}{\underline{\chi}}; \overset{(1)}{\underline{\chi}}) + Q(\overset{(1)}{\underline{\alpha}(F)}; \overset{(1)}{\underline{\alpha}(F)}) , $$ where we have used $\overset{(1)}{\zeta_i}=0$. The remaining system for $\overset{(3)}{\rho}$, $\overset{(3)}{\sigma}$, etc, is non-degenerate and can be solved. \\ This finishes the proof of Proposition \ref{prop:recurrence relations} and Lemma \ref{lem:evolution equations}. $_{\blacksquare}$\\ \end{enumerate} \textbf{Proof of Corollary \ref{cor:non-radiating}.} All the hard work was done in the proof of Proposition \ref{prop:recurrence relations}. Now we have to check the lower order cases. \\ Firstly, we show that the 1st order asymptotic quantities are all $u-$independent. Recall that $\overset{(1)}{{h_i}^j}=\mathrm{diag(1, 1/\sin\theta)}$ and $\overset{(1)}{\chi_{ij}}=\eta_{ij}$. Moreover, $\overset{(1)}{\underline{\alpha}(F)_i}=0$ and $\overset{(1)}{\underline{\chi}_{ij}}=\eta_{ij}$ by the non-radiating condition. We also know that $\overset{(n)}{\alpha_{ij}}= \overset{(n)}{\beta_i}=0$ for $n=0,1,2,3$; $\overset{(n)}{\alpha(F)_i} = \overset{(n)}{\rho}= \overset{(n)}{\sigma}=0$ for $n=0,1,2$; and $\overset{(n)}{\rho(F)}= \overset{(n)}{\sigma(F)}=0$ for $n=0,1$.\\ Now we focus on $\overset{(1)}{\zeta_i}$ and look at equations (\ref{hierarchy3}), we can get rid of the Weyl term by subtracting them. To 1st order we find that $$ \partial_u \overset{(1)}{\zeta_i} = \{ \overset{(1)}{S_{1i}} \} , $$ which vanishes since $S_{1i} = \mathcal{O}(r^{-3})$ due to Einstein-Maxwell equations. Hence $\overset{(1)}{\zeta_i}$ is also $u$-independent. The frame equation $e_0(f^i) = -\zeta^k {h_k}^i$ to 2nd order gives us that $\overset{(1)}{f^i} = \overset{(1)}{\zeta^k} \overset{(1)}{{h_i}^i}$ and thus it is also $u$-independent. This finishes the $\beta$-level.\\ Moving on to the $(\rho,\sigma)$-level: Recall that $\overset{(1)}{\underline{\chi}_{ij}} = \Xi_{ij} + \eta_{ij}$. By assumption $\Xi_{ij} =0$, so $\overset{(1)}{\underline{\chi}_{ij}}$ is $u$-independent. Moreover, $\overset{(2)}{\rho} =\overset{(2)}{\sigma} = \overset{(2)}{S_{01}} =0$, then the structure equations (\ref{hierarchy2}c) and (\ref{hierarchy2}d) to 2nd order give us that $\overset{(1)}{\underline{\omega}}$ and $\overset{(1)}{\omega_{123}}$ depends on previous data, so they must be $u$-independent. Finally, to 1st order, the frame equations $ e_i(f^0) = \lambda_i$ tell us that $\overset{(1)}{\lambda_i}=0$, since $\overset{(0)}{f^0}=1$. Therefore all the 1st order quantities at this level are $u$-independent.\\ We will see that we can repeat the argument now for asymptotic quantities of order 2. Firstly, the $u$-derivative of the recurrence relation (\ref{recursion3}) for $n=1$ tell us that \begin{align*} \partial_u^2\overset{(2)}{\chi_{ij}} &= \partial_u^2\overset{(2)}{\alpha(F)_i}, \\ \partial_u^2\overset{(2)}{\alpha(F)_i} &= 0 , \end{align*} where we have used that the 1st order quantities are $u$-independent. Now we make use of the hypothesis that $\overset{(3)}{\alpha_{ij}}$ and $\overset{(2)}{\alpha(F)_i}$ must remain finite\footnote{At this point is not really necessary to use this hypothesis since we already know that $\overset{(2)}{\alpha(F)_i}=0$. We have phrased it this way for the sake of generality in the induction argument.} as $u \rightarrow -\infty$ to rule out linear growth of $\overset{(2)}{\chi_{ij}}$ and $\overset{(2)}{\alpha(F)_i}$. Hence, $\overset{(2)}{\chi_{ij}}$ and $ \overset{(2)}{\alpha(F)_i}$ are $u$-independent. \\ Now, at the $\beta$-level we find that the equation (\ref{hierarchy1}a) degenerates at this order; the $\overset{(2)}{\zeta_i}$-terms cancel. Also $\overset{(3)}{S_{0i}}=0$, this together with equation (\ref{hierarchy1}b) gives a relation of the form: $$\overset{(2)}{\zeta_i} = \overset{(2)}{\omega_{jji}}. $$ On the other hand we can use the frame equation (\ref{hierarchy1}e) as before to find $\overset{(2)}{\omega_{jji}}$ in terms of $\overset{(2)}{{h_i}^j}$ and 1st order terms. This proves that $\overset{(2)}{\zeta_i}$ and $\overset{(2)}{\omega_{jji}} $ are $u$-independent. \\ As forecast, the Maxwell equations (\ref{hierarchy1}c) and (\ref{hierarchy1}d) also degenerate. This suggests that we need to provide $\overset{(2)}{\rho(F)}$ and $\overset{(2)}{\sigma(F)}$ as initial data. However, equations (\ref{mass and angular momentum}e) and (\ref{mass and angular momentum}f) impose evolution equations for these. Therefore they are $u$-independent in the non-radiating case. This finishes the $\beta$-level at this order.\\ Next, as seen in part c) of the proof of Proposition \ref{prop:recurrence relations}, the equations at the $(\rho, \sigma)$-level also degenerate. Quick inspection and the fact that $\overset{(3)}{S_{01}}=\overset{(3)}{S_{ij}}=0$ lead to identifying $\overset{(2)}{\tr \underline{\chi}}$ as initial data. Then all the remaining quantities can be computed in terms of this. \\ Moreover, the evolution equation for $\overset{(2)}{\tr \underline{\chi}}$ implies that it is $u$-independent in the non-radiating case. Therefore all the quantities at this level are also $u$-independent. \\ Thus we have established $u$-independence of all the quantities to 1st and 2nd order. In general, we can prove that the 1st and 2nd order quantities depend solely on the radiation fields and three aspect functions: mass and EM-charges. \\ It is clear now how to repeat the procedure inductively for $n>2$: Assume that $\overset{(n)}{\chi_{ij}}$, $\overset{(n)}{\alpha(F)_i}$ and $\lfloor n-1 \rfloor$ are $u$-independent then run the hierarchy-recursion argument to conclude that $\lfloor n \rfloor$ is $u$-independent. \\ At this point we encounter one difficulty when analysing the $\beta$-level to order 3. It degenerates, but we already know that we have to prescribe an angular momentum aspect vector $\overset{(3)}{\zeta_i}$ which comes with an evolution equation (\ref{mass and angular momentum}d). We now that we can compute $\overset{(3)}{\omega_{jji}}$ from previous data, equation (\ref{mass and angular momentum}c), hence it is $u$-independent in the non-radiating case. Therefore the evolution equation for $\overset{(3)}{\zeta_i}$ implies that it is also $u$-independent and the induction works through. \\ Finally, the evolution equations for $\overset{(n+1)}{\chi_{ij}}$ and $\overset{(n+1)}{\alpha(F)}$, (\ref{recursion3}a) and (\ref{recursion3}b), ensure that they will also be $u$-independent (provided they have finite limits as $u \rightarrow -\infty$) and the inductive step is proved. $_{\blacksquare}$ \section{Unique continuation from infinity } \setcounter{footnote}{1} In the previous section we showed that all the asymptotic quantities depend only on $\chi_{ij}$ along the out-going null hypersurface $C_0^+$ and the radiating fields $\Xi_{ij}$ and $\underline{A}(F)_i$. In particular, if $\Xi_{ij}$ and $\underline{A}(F)_i$ vanish then all the quantities are $u$-independent; that is, $T=\partial_u$ is a Killing field to all orders at infinity. The goal now is to extend this symmetry into the spacetime. We do this in Proposition \ref{prop:unique continuation2}, which in conjuction with Corollary \ref{cor:non-radiating}, allows us to prove our main result, Theorem \ref{thm1}.\\ We start here by presenting the necessary techniques to extend the time-like symmetry into a neighbourhood of infinity. The motivation comes from successful applications of the so-called Carleman estimates to prove uniqueness of solutions in the context of hyperbolic equations with pseudo-convex boundary conditions. For example, Ionescu-Klainerman used this approach to prove local unique extension of Killing vector fields across a pseudo-convex hypersurface in \cite{IK}. Also, in \cite{AIK}, Alexakis-Ionescu-Klainerman showed uniqueness of smooth stationary black holes for small perturbations of Kerr by proving the unique continuation property for the Simon-Mars tensor. Here we present the main results when the boundary condition imposed is that of asymptotic flatness. Alexakis-Schlue-Shao showed in \cite{ASS} that linear waves satisfy the unique continuation from infinity property provided they decay faster than any polynomial. The main technical tool is the new Carleman estimates they derived in the context of asymptotically flat spacetimes, we include them here as Theorem \ref{thm:Carleman}.\\ In the second part of this section we revise, and extend to the non-vacuum case, the tensorial equations satisfied by the deformation tensors $\mathcal{L}_Tg$, $\mathcal{L}_TC$ and $\mathcal{L}_T F$ obtained by \cite{IK}. These take the form of transport equations for $\mathcal{L}_T g$ and its first derivatives along the outgoing null direction and wave equations for $\mathcal{L}_TC$ and $\mathcal{L}_T F$. The latter are a consequence of the usual wave equations satisfied by $C$ and $F$, which in turn are implied by the Bianchi and Maxwell equations.\\ Then, in the third section we use Cartesian coordinates to cast the equations in a suitable form (with ``fast decaying coefficients'') such that the Carleman estimates from Theorem \ref{thm:Carleman} can be applied. Finally, a standard argument is used to bound the weighted $L^2$-norms of the deformation tensors and hence conclude its vanishing in a neighbourhood of infinity. Special care has to be taken regarding the wave equation for $\mathcal{L}_T F$ whose coefficients do not decay fast enough. To deal with this problem, we ``borrow'' some decay from the coupling coefficient appearing in the wave equation for $\mathcal{L}_T C$. \subsection{Carleman estimates} \label{section:Carleman estimates} In \cite{AS}, Alexakis and Schlue proved Theorem \ref{thm1} for a vacuum spacetime. In order to generalise the proof to include a Maxwell field we need to adapt their argument at the level of Carleman estimates. These are inequalities that can be thought of as a priori estimates for functions decaying faster than any polynomial at infinity. In conjunction with a wave equation satisfied by the function, this method can be used to prove the vanishing of the function in a neighbourhood of infinity. \\ We present here the Carleman estimates obtained by Alexakis, Schlue and Shao in \cite{ASS} for spacetimes of positive mass. To do so we introduce some notation first. \\ The class of spacetimes considered in \cite{ASS} are incomplete 4-manifolds of the form $(-\infty,0) \times (0, \infty) \times \mathbb{S}^2$ with coordinates $(u,v, y^i)$ equipped with a metric given by: \begin{align} \label{metric AS} g & = g_{uu}du^2 - 4 K du dv + g_{vv} dv^2 + \sum_{i.j=2}^3 \gamma_{ij} dy^i dy^j \\ \nonumber & \qquad \sum_{i=2}^3 (g_{iu} dy^i du + g_{iv} dy^i dv) , \end{align} where $m$ and $r$ are smooth positive functions. In addition Alexakis-Schlue-Shao require the following bounds to hold: \begin{itemize} \item The metric components satisfy \begin{align*} K & = 1- \frac{2m}{r} ,\qquad & g_{uu}, g_{vv} &= \mathcal{O}_1(r^{-3}), \\ g_{iu}, g_{iv} &= \mathcal{O}_1(r^{-1}) , \qquad & \gamma_{ij} &= \breve{\gamma}_{ij} + \mathcal{O}_1(r^{-1}). \end{align*} \item The function $m$ is uniformly bounded away from $0$, $m\geq m_{min} >0$, and for some $\eta>0$, the differential of $m$ satisfies \begin{align*} |\partial_i m| = \mathcal{O}(r(-uv)^{-\eta}), \qquad |\partial_um|, |\partial_v m| = \mathcal{O}(r^{-2}). \end{align*} \item The function $r$ is bounded on a level set of $v-u$, that is, there exist constants $C_1$, $C_2$ such that \begin{align*} v(p)-u(p) = C_1 \quad \textrm{implies} \quad 1 \ll |r(p)| < C_2 . \end{align*} The differential of $r$ satisfies the following estimate \begin{align*} \left( 1+ \frac{2m}{r}\right) dr = (1+\mathcal{O}(r^{-2})) dv - (1+\mathcal{O}(r^{-2}))du + \sum_{i=2}^3 \mathcal{O}(r^{-1})dy^i. \end{align*} Note that these two conditions imply that $r$ and $v-u$ are comparable, that is, there exists a constant $C$ such that $r \leq C |v-u|$. \item Finally they also require \begin{align*} | \square_g \left( \frac{m}{r}\right)| = \mathcal{O}((-uv)^{-1-\eta}). \end{align*} \end{itemize} \textbf{Remark.} It is not clear that an asymptotically flat spacetime in the sense of definition \ref{def:AF} satisfies the above condition when cast in double-null coordinates. While this may not be true in general, it holds for non-radiating spacetimes. We prove this in Lemma \ref{lem:doublenull}. \\ Now, an important ingredient in the deduction of the Carleman estimates is a \emph{pseudo-convex} function. Geometrically, in the Lorentzian context, these are functions whose level sets are convex with respect to null geodesics; that is, any null geodesic tangent to a level set locally remains on one side of that level set. This is equivalent to the following quantitative condition, \cite{Hormander}: A function $f$ on $(\mathcal{M},g)$ is pseudo-convex if there exists a function $h$ such that \[ \pi := h g - \nabla^2 f \] is positive definite when restricted to the tangent space of the level sets of $f$. Alexakis-Schlue-Shao proved that, with respect to the above coordinates, the function $$ f:= \frac{1}{(-u)v}, $$ is pseudo-convex. Therefore we expect the unique continuation property to hold across its level sets. One technicality arises in that this pseudo-convexity degenerates towards infinity. To cope with this degeneracy, which takes the form of vanishing/blowing up weights towards infinity, Alexakis-Schlue-Shao rely on a reparametrisation of $f$, $$ F(f) := \log f - f^{2\delta} ,$$ for some $\delta>0$ to be chosen later. Then they are able to conclude the unique continuation from infinity property in a neighbourhood of the form \[ \mathcal{D}_{\omega} := \{ (u,v,y^i) : 0 < f(u,v) < \omega \} . \] In order to understand the geometry of pseudo-convex timelike hypersurfaces we start by analysing them in Minkoswki spacetime. Consider double null coordinates $(u,v, \theta^2, \theta^3)$. The pseudo-convex timelike hypersurfaces considered by Alexakis-Schlue-Shao \cite{ASS} are given by the positive level sets of the function $f_{\epsilon} = \frac{1}{(-u + \epsilon)(v + \epsilon)}$, $\epsilon>0$. The $\epsilon$-perturbation is necessary to accomplish the pseudo-convexity condition in the absence of a mass. See Figure \ref{fig:pseudo-convex}. \\ \begin{figure} \label{fig:pseudo-convex} \begin{minipage}{8cm} \begin{center} \begin{tikzpicture}[scale=1] \draw (-3,0) node {}; \fill (0,4) circle (2pt) node[above] {$i^+$}; \fill (0,-4) circle (2pt) node[below] {$i^-$}; \fill (4,0) circle (2pt) node[right] {$i^0$}; \draw [line width = 1.5pt, red](1.7,2.3) ..controls (2.7,0).. (1.7,-2.3); \draw [red] (1.7,2.3) (0,0.6); \draw [red] (1.7,-2.3) (0,-0.6); \draw [line width = 1.5pt, dashed] (0,-4) -- (0,4) node[pos=.5,sloped,above] {$r=0$}; \draw [line width = 1.5pt] (0,4) -- (4,0) node[pos=.7, above right]{$I^+$}; \draw [line width = 1.5pt] (0,-4) -- (4,0) node[pos=.7, below right]{$I^-$}; \draw [<->] (1.85, 2.45) -- (2.15, 2.15) node[above]{$\epsilon$}; \draw [<->] (1.85, -2.45) -- (2.15, -2.15) node[below]{$\epsilon$}; \draw[dotted] (0,0) -- (2,2); \draw[dotted] (0,0) -- (2,-2); \draw [dotted](0,3.4) ..controls (1,2.8).. (2,2); \draw [dotted](0,2.8) ..controls (.9,2.6).. (2,2); \draw [dotted](0,2) -- (2,2); \draw [dotted](0,1.3) ..controls (.9,1.4).. (2,2); \draw [dotted](0,.6) ..controls (1,1.2).. (2,2); \draw [dotted](0,-3.4) ..controls (1,-2.8).. (2,-2); \draw [dotted](0,-2.8) ..controls (.9,-2.6).. (2,-2); \draw [dotted](0,-2) -- (2,-2); \draw [dotted](0,-1.3) ..controls (.9,-1.4).. (2,-2); \draw [dotted](0,-.6) ..controls (1,-1.2).. (2,-2); \draw [dotted](2,2) -- (2,-2); \draw [dotted](2,2) ..controls (0,0).. (2,-2); \draw [dotted](2,2) ..controls (1,0).. (2,-2); \draw [dotted](2,2) ..controls (3,0).. (2,-2); \draw [dotted](2,2) ..controls (4,0).. (2,-2); \end{tikzpicture} \end{center} \end{minipage} \begin{minipage}{8cm} \begin{center} \begin{tikzpicture}[scale=1] \shade [shading=axis,left color=white, right color=gray] (4,0)--(2.5,-1.5) .. controls (3,0).. (2.5,1.5); \fill (0,4) circle (2pt) node[above] {$i^+$}; \fill (0,-4) circle (2pt) node[below] {$i^-$}; \fill (4,0) circle (2pt) node[right] {$i^0$}; \draw [line width = 1.5pt, dashed] (0,4) -- (-1.5,2.5); \draw [line width = 1.5pt, dashed] (0,-4) -- (-1.2,-2.5); \draw [line width = 1.5pt, red] (2.5,-1.5) .. controls (3,0).. (2.5,1.5); \draw [line width = 1.5pt] (0,4) -- (4,0) node[pos=.5, above right]{$I^+$}; \draw [line width = 1.5pt] (0,-4) -- (4,0) node[pos=.5, below right]{$I^-$}; \draw [line width = 1.5pt, dashed] (0,4) -- (-1.5,4); \draw [line width = 1.5pt, dashed] (0,-4) -- (-1.5,-4); \draw (1, 0) -- (2.5,1.5); \draw (1, 0) -- (2.5,-1.5); \draw [dashed] (0,-4) .. controls (1.3,0).. (0,4) node[pos=.5, above, sloped]{$r=r_0$}; \draw (3.4,0) node {$\mathcal{D}_{\omega}$}; \end{tikzpicture} \end{center} \end{minipage} \caption{\emph{Left.} Dotted lines represent the level sets of $\frac{1}{(-u)v}$ in Minkowski spacetime. One level set of $f_{\epsilon} = \frac{1}{(-u + \epsilon)(v + \epsilon)}$ is also shown (red line). In order to ensure the unique continuation property for a wave in Minkowski spacetime one has to prescribe initial data (to all orders) on more than half of null infinity. \emph{Right.} One level set of $f= \frac{1}{(-u)v}$ is shown in Schwarzschild spacetime as well as the corresponding neighbourhood of infinity $\mathcal{D}_{\omega}$. The pseudo-convex function depends now on a parameter $r_0>2M$; different choices of $r_0$ give rise to `parallel' foliations. This behaviour is responsible for the localised result around spatial infinity for positive-mass spacetimes: Data require for unique continuation from infinity can be provided on small portions of null infinity.} \end{figure} On positive-mass spacetimes the situation is qualitatively different. Consider for example Schwarzschild spacetime in double null coordinates, $(u,v,\theta^2, \theta^3)$, recall that these are defined by $$ u = \frac{t-r_* }{2} , \qquad v = \frac{t+r_* }{2} , $$ where $$ r_*(r) = \int_{r_0}^r\left( 1- \frac{2M}{s}\right)^{-1}ds, \qquad r_0>2M . $$ Due to the presence of a mass now the function $f = \frac{1}{(-u )v}$ is pseudo-convex. \\ We are now in position to state the main technical tool of this section. For convenience we introduce the weight function $\mathcal{W}$ and associated weighted norms. For any $\lambda>0$ and domain $\mathcal{D}=\mathcal{D}_{\omega}$, $\omega>0$ $$ \mathcal{W} = \textrm{e}^{-\lambda F} f^{\frac12}, \quad \| \cdot \|_{\mathcal{W}} = \| \mathcal{W} \cdot \|_2, \quad \| \phi \|^2_2 = \int_{\mathcal{D}} |\phi|^2 d\mu_g . $$ \begin{thm} \label{thm:Carleman} \emph{(Carleman estimate near infinity for linear waves, \cite{AS}) } Let $(\mathcal{M},g)$ be an asymptotically flat spacetime with positive mass $m\geq m_{min}>0$ in the sense of (\ref{metric AS}) and $\mathcal{D}_{\omega}$ a neighbourhood of infinity for some $\omega>0$. Let $\delta>0$ and let $\phi$ be a smooth function on $\mathcal{D}_{\omega}$ that vanishes to all orders at infinity in the sense that for each $N\in \mathbb{N}$ there exist an exhaustion\footnote{A nested family of subsets, with piece-wise smooth time-like boundaries, whose union is all of $\mathcal{D}_{\omega}$.} $(\mathcal{U}_k)$ of $\mathcal{D}_{\omega}$ such that \begin{align} \label{vanishing condition} \lim_{k\rightarrow \infty} \int_{\partial \mathcal{U}_k} r^N (\phi^2+|\partial \phi|^2) = 0 . \end{align} Then, for $\omega>0$ sufficiently small and $\lambda>0$ sufficiently large, \begin{align} \label{Carleman} \lambda^3 \| f^\delta \phi \|_{\mathcal{W}} + \lambda \| f^{-\frac12} \Psi^{\frac12} \nabla \phi \|_{\mathcal{W}} \lesssim \| f^{-1} \square \phi \|_{\mathcal{W}} , \end{align} where $\Psi$ is defined by \[ \Psi := \frac{ m_{min} \log r}{r} . \] \end{thm} \textbf{Sketch of proof.} Consider $\varphi = \textrm{e}^{-\lambda F(f)} \phi$. The idea is to obtain an energy estimate for $\varphi$, but here we wish for the bulk terms of the integral to be positive and for the boundary terms to vanish. That is, consider the modified energy current \[ J_{\beta}^{w}[\varphi] = Q_{\alpha \beta}[\varphi] \nabla^{\alpha} f + \left( \frac12 \partial_{\beta} w + \frac12 \lambda^2 (\nabla^{\alpha}f)(\nabla_{\alpha}f)(F')^2 \partial_{\beta} f \right) \varphi^2 - \frac12 w\partial_{\beta}(\varphi^2) , \] where $Q_{\alpha \beta}[\varphi]$ is the standard energy-momentum tensor for $\square \varphi=0$. \\ The function $w$ is to be chosen appropriately so that $\divergence J_{\beta}^{w}$ produces the tensor $\pi$, hence capturing the pseudo-convexity of $f$. Specifically, the choice $$ w = h - \frac12 \square f - \frac12, $$ produces positive bulk terms that are quadratic in $\partial_X \varphi$ for $X$ tangent to the level sets of $f$; in order to obtain positivity in the normal direction one relies on the choice of reparametrisation $F(f)$. \\ The above procedure ultimately results in an inequality of the form \begin{align*} \int_{\mathcal{D}_{\omega}} \mathcal{W}_L |\mathcal{P} \varphi|^2 \geq & \, C \lambda \int_{\mathcal{D}_{\omega}} \left( \mathcal{W}_N |\nabla_N \varphi|^2 +\mathcal{W}_T |\nabla_T \varphi|^2 +\mathcal{W}_T \sum |\nabla_{e_i} \varphi|^2 \right) \\ & C \lambda^3 \int_{\mathcal{D}_{\omega}} \mathcal{W}_0 |\varphi|^2 + \int_{\mathcal{D}_{\omega}} \mathcal{E}, \end{align*} where $\mathcal{P} (\varphi) = \textrm{e}^{-\lambda F(f)}\square(\textrm{e}^{\lambda F(f)} \varphi)$ and $\mathcal{W}_L$, $\mathcal{W}_N$, $\mathcal{W}_T$, $\mathcal{W}_0$ are positive weights. The only non-positive term is the error $\mathcal{E}$ which must be absorbed into the remaining positive terms. $_{\blacksquare}$\\ Alexakis-Schlue-Shao used the above Carleman estimate to extend the vanishing of a function $\phi$ into a neighbourhood of infinity, \cite{ASS}. The proof is based on standard arguments using cut-off functions to cope with the ``inner decay'' necessary for Theorem \ref{thm:Carleman} to hold. The wave equation is used to substitute $\square \phi$ by lower order derivatives which can be absorbed on the left-hand side of (\ref{Carleman}) due to the decaying conditions. Finally the result follows by taking $\lambda$ to infinity. We omit a detailed proof since we will follow the same program in the proof of Proposition \ref{prop:unique continuation2}. \\ We will also need Carleman estimates for transport equations involving $\mathcal{L}_T g$ and its first derivatives. These are covered by the following lemma proved in \cite{AS}. \begin{lem} \label{lem:CarlemanODE} Let $(\mathcal{M}, g)$ be an asymptotically flat spacetime and $L=\partial_v$ be the outgoing null vector field in coordinates (\ref{metric AS}). Let $\phi$ be a smooth function on $\mathcal{D}_{\omega}$ that vanishes to all orders at infinity, in the sense that for any $N\in \mathbb{N}$ there is an exhaustion $(\mathcal{U}_k)$ of $\mathcal{D}_{\omega}$ such that $$\lim_{k\rightarrow \infty} \int_{\partial \mathcal{U}_k} r^N\phi^2 =0 . $$ Then for any $q\geq 1$ and $\lambda>0$ sufficiently large, $$ \lambda \|\frac1r f^{-1}r^{-q} \phi \|_{\mathcal{W}} \lesssim \|f^{-1} r^{-q} \nabla_L \phi \|_{\mathcal{W}} . $$ \end{lem} To conclude this section we present a technical result necessary to ensure that we can change from time-space coordinates in the asymptotically flat sense of definition \ref{def:AF} to the double-null coordinates required for the Carleman estimates. \begin{lem}\label{lem:doublenull} Let $(\mathcal{M},g, F)$ be the Cauchy development of initial data suitably close to Kerr-Newman. Specifically, let $g|_{\Sigma}$ and $\partial_t g|_{\Sigma}$ be induced on $\Sigma= \{ t=0 \}$ by \begin{align} \label{Kerr-Newman} g &= g_{KN} + g^{\infty} \nonumber \\ & = - \left( 1- \frac{2M}{r} + \frac{e^2}{r^2} + \mathcal{O}^{\infty}_2(r^{-3}) \right) dt^2 + \mathcal{O}^{\infty}_2(r^{-4}) dt dr \nonumber \\ & + \left( 1 + \frac{2M}{r} - \frac{a^2 + e^2}{r^2}\sin^2\vartheta^2 + \frac{2Ma^2}{r^3}\cos^2\vartheta^2 + \mathcal{O}^{\infty}_2(r^{-4}) \right) dr^2 \nonumber \\ & + \left( \breve{\gamma}_{ij} + \mathcal{O}^{\infty}_2(1) \right) d\vartheta^i d\vartheta^j + \sum_{i=2}^3 \left( \mathcal{O}^{\infty}_2(r^{-1})dt d\vartheta^i + \mathcal{O}^{\infty}_2(r^{-3})dr d\vartheta^i \right) \end{align} Suppose $(\mathcal{M},g)$ is asymptotically flat, in the sense of Definition \ref{def:AF}. Moreover, assume that the vector field $T$, as constructed in Section \ref{section:tetrad and coordinates}, is Killing to order 4, that is, \[ \lim_{r \rightarrow \infty} r^N \mathcal{L}_T g = 0 \qquad N=0,\ldots,4. \] Then we can change to the double-null coordinates required for the Carleman estimates of Theorem \ref{thm:Carleman}, i.e., there exists a domain $\mathcal{D}=(-\infty, 0)\times (0, \infty) \times \mathbb{S}^2 \subset \mathcal{M}$ with coordinates $(u,v, y^i)$ such that the metric components verify condition (\ref{metric AS}) and the corresponding bounds. \end{lem} \textbf{Proof.} We will prove that the asymptotic expansion (\ref{Kerr-Newman}) holds through out the development when written in $T$-adapted coordinates. Then a transformation to `comoving coordinates', as in \cite{ASS}, shows that $(\mathcal{D}, g)$ is of the desired form. \\ First we need to find estimates for $T$ at $\Sigma$ with respect to the given coordinate system. Let us write, \[ T = \tau \partial_t + \rho \partial_r + \theta^i \partial_{\vartheta^i} . \] Recall that $T = \partial_u = \frac{L + \underline{L}}{2} + \sum_{\mu=0}^3 \mathcal{O}(r^{-1})e_\mu $. \\ We have abused notation by using the same symbol, $r$, for the area function of Section \ref{section:tetrad and coordinates} and the coordinate function in (\ref{Kerr-Newman}). However, due to the asymptotic expansion (\ref{Kerr-Newman}) assumed on $\Sigma$ we have that they agree up to $\mathcal{O}(1)$, hence the symbol $\mathcal{O}(r^{-k})$ is unambiguous. Moreover, the cone $C^-_{d^*}$ was chosen so that intersects $\Sigma$ on $r=d^*$, this guarantee that $L$ and $\underline{L}$ agree to leading order with the in-going and out-going null directions defined by the level sets of $r$. Therefore, given the expansion (\ref{Kerr-Newman}), $T$ must agree with $\partial_t$ to leading order, that is, $\tau \rightarrow 1$ and $\rho, \theta^i \rightarrow 0$ as $r \rightarrow \infty$ along $\Sigma$.\\ Now we use the hypothesis that $T$ is Killing to order 4 to find more refined estimates for $\tau, \rho, \theta^i$ at $\Sigma$. Indeed, we have, \begin{align*} (\mathcal{L}_T g)_{\mu \nu} & = (\partial_{\mu} \tau) g_{t \nu} + (\partial_{\nu} \tau) g_{\mu t} + \tau(\partial_t g)_{\mu \nu} \\ & \quad + (\partial_{\mu} \rho) g_{r \nu} + (\partial_{\nu} \rho) g_{\mu r} + \rho(\partial_r g)_{\mu \nu} \\ & \quad + (\partial_{\mu} \theta^i) g_{\vartheta^i \nu} + (\partial_{\nu} \theta^i) g_{\mu \vartheta^i} + \theta^i(\partial_{\vartheta^i} g)_{\mu \nu} . \end{align*} Also, (\ref{Kerr-Newman}) gives control on $\partial g$ along $\Sigma$: \begin{align*} (\partial_t g)_{rr} &= \mathcal{O}^{\infty}_1(r^{-4}), &\quad (\partial_t g)_{rt} &= \mathcal{O}^{\infty}_1(r^{-4}), &\quad (\partial_t g)_{r \vartheta^i} &= \mathcal{O}^{\infty}_1(r^{-3}), \\ (\partial_r g)_{rr} &= -\frac{2M}{r^2} + \mathcal{O}^{\infty}_1(r^{-4}), &\quad (\partial_r g)_{rt} &= \mathcal{O}^{\infty}_1(r^{-5}), &\quad (\partial_r g)_{r \vartheta^i} &= \mathcal{O}^{\infty}_1(r^{-4}), \\ (\partial_{\vartheta^i} g)_{rr} &= \mathcal{O}^{\infty}_1(r^{-2}), &\quad (\partial_{\vartheta^i} g)_{rt} &= \mathcal{O}^{\infty}_1(r^{-4}), &\quad (\partial_{\vartheta^i} g)_{r \vartheta^j} &= \mathcal{O}^{\infty}_1(r^{-3}) . \end{align*} Therefore, to leading order, the $(r,{\vartheta^i})$, $(r,r)$ and $(r,t)$ components of $\mathcal{L}_T g$ imply the following differential estimates: \begin{align*} \partial_r \theta^i + \frac{1}{r^2}(\partial_{\vartheta^i} \rho) + \frac{1}{r^2}(\partial_r \tau)\mathcal{O}(r^{-1}) &= \mathcal{O}(r^{-4}) , \\ 2(\partial_r \rho)\left( 1 +\frac{2M}{r} \right) - \left( \frac{2M}{r^2} + \mathcal{O}(r^{-2}) \right) \rho + \theta^i \mathcal{O}(r^{-2}) & = \mathcal{O}(r^{-4}) , \\ -2(\partial_r \tau)\left( 1-\frac{2M}{r} \right) + (\partial_r \theta)\mathcal{O}(r^{-1}) + \theta^1\mathcal{O}(r^{-3}) & = \mathcal{O}(r^{-4}) . \end{align*} Now, the a priori estimates, $\tau \rightarrow 1$ and $\rho, \theta^i \rightarrow 0$, together with the above bounds show that $\partial_r\theta^i, \partial_{r} \rho, \partial_r \tau = \mathcal{O}(r^{-2})$. Thus, \begin{align} \tau-1, \rho, \theta^i = \mathcal{O} (r^{-3}). \end{align} Finally, we define $T$-adapted coordinates by setting $\bar{t}=0$ on $\Sigma$ and $T(\bar{t})=1$, while coordinates $(r, y^2, y^3)$ are defined to be constant along $T$-lines with $(r, y^2, y^3)= (r, \vartheta^2, \vartheta^3)$ on $\Sigma$. The goal of the previous analysis was to achieve the correct asymptotic behaviour of the metric in this new coordinates, indeed, it is easy to check that \begin{align*} g = & - \left( 1- \frac{2M}{r} + \frac{e^2}{r^2} + \mathcal{O}^{\infty}_2(r^{-3}) \right) d\bar{t}^2 + \mathcal{O}^{\infty}_2(r^{-4}) d\bar{t} dr \nonumber \\ & + \left( 1 + \frac{2M}{r} - \frac{a^2 + e^2}{r^2}\sin^2y^2 + \frac{2Ma^2}{r^3}\cos^2y^2 + \mathcal{O}^{\infty}_2(r^{-4}) \right) dr^2 \nonumber \\ & + \left( \breve{\gamma}_{ij} + \mathcal{O}^{\infty}_2(1) \right) dy^i dy^j + \sum_{i=2}^3 \left( \mathcal{O}^{\infty}_2(r^{-1})d\bar{t} dy^i + \mathcal{O}^{\infty}_2(r^{-3})dr dy^i \right) \end{align*} In these coordinates, the metric components are easily controlled by the initial data and $\mathcal{L}_T g$, indeed, \[ \left| g_{\mu \nu} (\bar{t}, r, y^2, y^3) - g_{\mu \nu} (0,r, y^2, y^3) \right| \leq \int_0^{\bar{t}} | (\mathcal{L}_T g)_{\mu \nu} (t,r,y^2, y^3)| dt . \] Therefore on a domain of the form, $$ \mathcal{D} = \{ (\bar{t}, r, y^2, y^3) : r\geq R, |\bar{t}| \leq r + 2M\log r - (R+2M\log R) \} , $$ we have that $|g_{\mu \nu} (\bar{t}) - g_{\mu \nu}(0)| = \mathcal{O}(r^{-3})$. \\ Then a transformation to comoving coordinates as in \cite{ASS} achieves the desired double-null form (\ref{metric AS}). $_{\blacksquare}$ \subsection{Proof of the main Theorem} Now we proceed to the last part of the proof of Theorem \ref{thm1}. Namely, we show that the vector field $T=\partial_u$ is indeed a symmetry of the gravitational and electromagnetic fields provided that it is a symmetry to all orders at infinity. More precisely we have the following: \begin{prop} \label{prop:unique continuation2} Let $(M,g,F)$ be an asymptotically flat electrovacuum spacetime with rapidly decaying curvature in the sense that uniformly with respect to an orthonormal frame \begin{align} \label{decay} \begin{aligned} |C_{\alpha \beta \gamma \delta}| & = \mathcal{O}(r^{-3}), \\ |\nabla_\epsilon C_{\alpha \beta \gamma \delta}| &=\mathcal{O}(r^{-4}), \end{aligned} && \begin{aligned} |F_{\alpha \beta}| &=\mathcal{O}(r^{-2}), \\ |\nabla_{\epsilon} F_{\alpha \beta}| &=\mathcal{O}(r^{-3}). \end{aligned} \end{align} Let $(u,s, \theta^2,\theta^3)$ be coordinates on $D_{\omega}$, $\omega>0$, constructed as in Section \ref{section:tetrad and coordinates} with $L:=\partial_s$ and $T:=\partial_u$ satisfying \[ \nabla_L L=0 , \qquad [L,T]=0 . \] Assume that $T$ is a symmetry to all orders at infinity, i.e., for all $N\in \mathbb{N}$ there is an exhaustion $(\mathcal{U}_k)$ of $D_{\omega}$ such that \begin{align} \label{killing to all orders} \lim_{k\rightarrow \infty} \int_{\partial \mathcal{U}_k} r^N \mathcal{L}_T g = 0, \qquad \lim_{k\rightarrow \infty} \int_{\partial \mathcal{U}_k} r^N \mathcal{L}_T C = 0, \qquad \lim_{k\rightarrow \infty} \int_{\partial \mathcal{U}_k} r^N \mathcal{L}_T F = 0. \end{align} Then $T$ is in fact locally a genuine symmetry for $(\mathcal{M},g,F)$, namely \[ \mathcal{L}_Tg \equiv 0, \quad \mathcal{L}_T C \equiv 0, \quad \mathcal{L}_T F \equiv 0, \quad \textrm{ on } \mathcal{D}_{\omega'}\] for some $0<\omega'<\omega$. \end{prop} Note that conditions (\ref{decay}) corresponds precisely to the non-radiating hypothesis of Theorem \ref{thm1}; while conditions (\ref{killing to all orders}) are those deduced from Corollary \ref{cor:non-radiating}. \\ The strategy to prove Proposition \ref{prop:unique continuation2} is analogous to the one used by Ionescu-Klainerman in \cite{IK} and Alexakis-Schlue in \cite{AS}. We start by deducing tensorial wave equations for $\mathcal{L}_T C$ and $\mathcal{L}_T F$. When written in Cartesian coordinates those equations fit into the framework provided by the Carleman estimates of Theorem \ref{thm:Carleman}. Once we get those differential estimates a standard argument to deduce unique continuation follows through.\\ We emphasise here that the full system of equations, (\ref{coupled system}) below, does not fit into the Alexakis-Schlue argument since one of the coupling terms does not decay fast enough. We deal with this problem by borrowing decay from the well-behaved coefficients and by using different Carleman weights for each of $\mathcal{L}_T C$ and $\mathcal{L}_T F$. See equations (\ref{CarlemanW3}), (\ref{CarlemanE}) and ensuing discussion. \subsection{Ionescu-Klainerman tensorial equations} In this section we recall the equations satisfied by the deformation tensors deduced by Ionescu and Klainerman in \cite{IK}. We remind the reader that their work assumes a vacuum spacetime, that is, the Riemann and Weyl tensors are equal, $R_{abcd}=C_{abcd}$. This is in contrast with our current approach where the Riemann tensor is coupled to the Faraday tensor via the Einstein equations. Hence we need to be careful now to distinguish between $R$ and $C$. \\ \\ \textit{Notation.} Through out this section we will denote schematically by \[ U \cdot V \] any linear combination of the product of two tensors $U$ and $V$, and contractions thereof. For example the relation \[ R_{abcd} = C_{abcd} + \frac{1}{n-1} \left( g_{ac} S_{bd} - g_{dc} S_{ba} + g_{db} S_{ca} - g_{ab} S_{cd} \right) . \] will be abbreviated to \[ R = C+ g \cdot S . \\ \] We proceed now to compute the Ionescu-Klainerman tensorial wave equations for $\mathcal{L}_T C$ and $\mathcal{L}_T F$. The idea is to use the wave equations satisfied by $C$ and $F$ and then commute the covariant and Lie derivatives with the help of Lemma \ref{lem:commutator}. The relevant equations are naturally coupled to $\pi_{ab}:=\mathcal{L}_T g_{ab}$ and $\nabla_a \pi_{bc}$, hence the necessity of finding (transport) equations for them. In order to get a closed system of equations we need to perform some algebraic tricks (related to the symmetries of the Weyl tensor) and work instead with auxiliary variables, $B_{ab}$ and $P_{abc}$; see Proposition \ref{prop:tensorial equations}. This section is entirely based on \cite{IK}.\\ We begin by noticing that $\mathcal{L}_T C$ is not trace-less. Indeed, $$ g^{ac} \mathcal{L}_T C_{abcd} = (-\mathcal{L}_T g^{ac})C_{abcd} = \pi^{ac}C_{abcd} . $$ To remedy this we define the modified Lie derivative \begin{align} \hat{\mathcal{L}}_T C := \mathcal{L}_T C - B\odot C . \end{align} where $B$ is a 2-covariant tensor and $(B\odot C)_{abcd} := {B_a}^eC_{ebcd} + {B_b}^eC_{aecd}+ {B_c}^eC_{abed}+ {B_d}^eC_{abce}$. If we take $$B = \frac12( \pi + \omega) ,$$ with $\omega$ any antisymmetric 2-form, a simple calculation leads to \begin{lem} The modified Lie derivative of the Weyl tensor, $W:= \hat{\mathcal{L}}_T C$, is a Weyl field, i.e., \begin{enumerate}[i)] \item $W_{abcd} =-W_{bacd} = -W_{abdc} = W_{cdab}$. \item $W_{abcd} + W_{acdb} + W_{abdc} = 0 $. \item $g^{ac}W_{abcd} = 0$. \end{enumerate} \end{lem} Next, we need to compute the commutator of the Lie and covariant derivatives. This is given by the following: \begin{lem} \label{lem:commutator} For an arbitrary $k$-covariant tensor $V$ and vector field $T$ we have, \begin{align} \nabla_b \mathcal{L}_T V_{a_1...a_k} - \mathcal{L}_T \nabla_b V_{a_1...a_k} &= \sum_{j=1}^k \Pi_{a_jb\rho} {V_{a_1...}}^{\rho}{}_{...a_k} , \\ \Pi_{abc} & := \frac12 \left( \nabla_a \pi_{bc} + \nabla_b \pi_{ca} - \nabla_c \pi_{ab} \right) . \end{align} Schematically we write, $[\nabla_b, \mathcal{L}_T] V = \Pi_b \odot V$. \end{lem} \textbf{Proof.} We compute \begin{align*} \mathcal{L}_T V_{a_1 \ldots a_n} &= T^c \nabla_c V_{a_1 \ldots a_n} + \sum_i (\nabla_{a_i}T^c)V_{a_1 \ldots c \ldots a_n} , \\ \nabla_b \mathcal{L}_T V_{a_1 \ldots a_n} &= (\nabla_b T^c ) \nabla_c V_{a_1 \ldots a_n} + T^c\nabla_b \nabla_c V_{a_1 \ldots a_n} \\ & \qquad + \sum_i (\nabla_b \nabla_{a_i}T^c)V_{a_1 \ldots c \ldots a_n} + (\nabla_{a_i}T^c) \nabla_b V_{a_1 \ldots c \ldots a_n}, \\ \mathcal{L}_T \nabla_b V_{a_1 \ldots a_n} &= T^c \nabla_c \nabla_b V_{a_1 \ldots a_n} + (\nabla_b T^c) \nabla_c V_{a_1 \ldots a_n} + \sum_i (\nabla_{a_i}T^c) \nabla_b V_{a_1 \ldots c \ldots a_n} . \end{align*} Then \begin{align*} [\mathcal{L}_T, \nabla_b] V_{a_1 \ldots a_n} = \sum_i \left ( T^c {R_{cba_i}}^f V_{a_1 \ldots f \ldots a_n} + (\nabla_b \nabla_{a_i} T^c) V_{a_1 \ldots c \ldots a_n} \right). \end{align*} And the results follows from the identity \begin{align*} \nabla_b\nabla_a T_c = R_{cabd}T^d + \Pi_{abc}. \end{align*} To prove this equality we just evaluate, commute derivatives and make use of the 1st Bianchi identity: \begin{align*} R_{cabd}T^d + \Pi_{abc} &= R_{cabd}T^d + \frac12 \left( \nabla_a (\nabla_b T_c + \nabla_cT_b) + \nabla_b(\nabla_aT_c + \nabla_cT_a) - \nabla_c(\nabla_aT_b + \nabla_bT_a)\right), \\ &= R_{cabd}T^d + \frac12 \left( R_{acbd}T^d + R_{bcad}T^d + 2 \nabla_a\nabla_bT_c - R_{bacd}T^d \right), \\ &= R_{cabd}T^d + \frac12 \left( 2R_{acbd}T^d + 2 \nabla_a\nabla_bT_c \right), \\ &= \nabla_a\nabla_bT_c . \qquad {}_{\blacksquare} \end{align*} Now we present the variables and equations which play the crucial role in the unique continuation analysis. The variables are minor modifications of $\pi$, $\nabla \pi$ and $\mathcal{L}_T C$ which give us a closed system of equations. \begin{prop} \label{prop:tensorial equations} Let $\pi_{ab}:= \mathcal{L}_Tg_{ab}$ and $\omega_{ab}$ be a 2-form solution of the transport equation $$ \nabla_L \omega_{ab} = \pi_{ac} \nabla_b L^c - \pi_{bc} \nabla_a L^c .$$ Define the tensors $B$, $P$ and $W$ as follows, \begin{align} B_{ab} &:= \frac12 \left( \pi_{ab} + \omega_{ab} \right), \\ P_{abc} &:= \frac12 \left( \nabla_a \pi_{bc} - \nabla_b \pi_{ac} -\nabla_c \omega_{ab} \right) = \Pi_{acb} - \nabla_c B_{ab}, \\ W_{abcd} &:= \mathcal{L}_T C_{abcd} - (B\odot C)_{abcd} , \end{align} Then the following equations hold \begin{align} \nabla_L B_{ab} &= L^c P_{cba} - B_{cb}\nabla_aL^c, \\ \nabla_L P_{abc} & = L^d (\mathcal{L}_T R_{abcd} - {B_a}^e R_{ebcd}- {B_b}^e R_{aecd} + P_{abd} \nabla_cL^d, \\ \square W & = \mathcal{L}_T (\square C) + \nabla P \cdot C + B\cdot \nabla C + \nabla B \cdot \nabla C + B \cdot \square C. \end{align} \end{prop} During the proof we will make use of the following identities: \begin{lem} Assume $[T,L]=0$, $L^c\nabla_c L_a = 0$ and $\pi$ vanishes to first order at infinity. Then $$L^b\pi_{ab}=0 , \quad P_{abc}L^c = 0, \quad L^b \omega_{ab}=0 .$$ \end{lem} \textbf{Proof of Lemma.} We start by showing $L^aL^b\pi_{ab}=0$. Indeed, \begin{align*} L^aL^b\pi_{ab} &= L^aL^b (\nabla_a T_b + \nabla_b T_a ) = L^b T^a\nabla_a L_b + L^a T^b\nabla_b L_a \\ &= T^a \nabla_a (L_b L^b) = 0. \end{align*} Now, by commuting derivatives we can find a transport equation for $L^b\pi_{ab}$. \begin{align*} L^c\nabla_c (L^b\pi_{ab} ) &= L^c \nabla_c (L^b (\nabla_a T_b + \nabla_b T_a ) ,\\ &= L^c L^b ( \nabla_a \nabla_c T_b + {R_{cab}}^e T_e) + L^c \nabla_c (T^b \nabla_b L_a) , \\ &= \nabla_a(L^cL^b) \nabla_c T_b + {R_{cab}}^e L^c L^b T_e) + {R_{cba}}^e L^cT^bL_e + T^c \nabla_c (L^b \nabla_b L_a),\\ &= (\nabla_aL^c) (L^b \pi_{cb}) , \end{align*} which is a homogeneous equation for $L^b \pi_{ab}$. In our context, the choice of $T$ implies that the deformation tensor $\pi$ vanishes to first order at infinity. In particular $L^b\pi_{ab}$ also vanishes to first order at infinity; this together with the above transport equation implies that $L^b\pi_{ab}\equiv 0$ as desired.\\ Next, a straightforward computation yields, \begin{align*} 2 P_{abc}L^c &= L^c(\nabla_a \pi_{bc} - \nabla_b \pi_{ac} -\nabla_c \omega_{ab}) , \\ &= -\pi_{bc}\nabla_aL^c + \pi_{ca}\nabla_bL^c - \pi_{ac} \nabla_b L^c + \pi_{bc} \nabla_aL^c = 0, \end{align*} where we have used $L^c\pi_{ac}=0$ and the definition of $\omega_{ab}$.\\ Finally, for the last equality we deduce a transport equation for $L^b\omega_{ab}$ using its definition, \begin{align*} L^c \nabla_c(L^b\omega_{ab}) &= L^b L^c \nabla_c \omega_{ab} , \\ &= L^b (\pi_{ac} \nabla_bL^c - \pi_{bc} \nabla_aL^c) = 0 , \end{align*} Hence $L^b\omega_{ab}$ is constant and we can choose initial conditions for $\omega_{ab}$ such that $L^b\omega_{ab}$ vanishes. $_{\blacksquare}$ \\ \textbf{Proof of Proposition \ref{prop:tensorial equations}.} For the transport equation for $B_{ab}$ we compute: \begin{align*} 2(L^c P_{cba} - B_{cb}\nabla_aL^c) &= L^c(\nabla_c\pi_{ba} - \nabla_b\pi_{ac} - \nabla_a\omega_{cb}) - (\pi_{cb}+\omega_{cb})\nabla_aL^c \\ &= \nabla_L \pi_{ab} + \pi_{ac}\nabla_b L^c + \omega_{cb}\nabla_aL^c - (\pi_{cb}+\omega_{cb})\nabla_aL^c \\ &= 2 \nabla_L B_{ab}, \end{align*} where we have used $L^c\pi_{cb}=L^c\omega_{cb}=0$ and the transport equation defining $\omega_{ab}$. \\ Next we deduce the transport equation for $P_{abc}$. Recall the following identity\footnote{Which is basically an antisymmetrised identity for $\nabla_d \nabla_a (\nabla_{(b} T_{c)})$ necessary to cope with the symmetries of $R_{abcd}$. The idea to prove it is to commute derivatives.} proved in \cite{IK} for $\tilde{P}_{abc} := \frac12 (\nabla_a\pi_{bc} - \nabla_b\pi_{ac})$, $$ \nabla_d\tilde{P}_{abc} - \nabla_c\tilde{P}_{abd} = \mathcal{L}_T R_{abcd} - \frac12 {\pi_a}^f R_{fbcd} - \frac12 {\pi_b}^f R_{afcd}. $$ Note that $P_{abc} = \tilde{P}_{abc} - \frac12 \nabla_c \omega_{ab}$, hence \begin{align*} L^d(\nabla_d P_{abc} - \nabla_c P_{abd}) &= L^d \left(\mathcal{L}_T R_{abcd} - \frac12 {\pi_a}^f R_{fbcd} - \frac12 {\pi_b}^f R_{afcd} + \frac12 (\nabla_c\nabla_d \omega_{ab} - \nabla_d\nabla_c \omega_{ab}) \right), \\ &= L^d \left(\mathcal{L}_T R_{abcd} - \frac12 {\pi_a}^f R_{fbcd} - \frac12 {\pi_b}^f R_{afcd} - \frac12 {\omega_a}^f R_{fbcd} - \frac12 {\omega_b}^f R_{afcd} \right), \\ &= L^d \left(\mathcal{L}_T R_{abcd} - {B_a}^f R_{fbcd} - {B_b}^f R_{afcd} \right), \end{align*} the final result follows by noticing that $L^d\nabla_c P_{abd} = - P_{abd}\nabla_cL^d$ since $L^dP_{abd}=0$.\\ Finally, we deduce the wave equation for $W$ by commuting Lie and covariant derivatives. We are interested only in the general structure of the equations, in particular, in the coefficients accompanying $W$ and $\nabla W$. Hence we do not any longer keep track of the different contractions but just on the bilinear structure of products and the different terms involving our variables $P$, $B$ and $W$. \\ \\ \textit{Notation.} During the following computations we will substitute freely $B$ instead of $\pi$ since $\pi_{ab} = B_{ab} + B_{ba}$. \\ We start now with a divergence equation for $W$. \begin{lem} The following holds: \begin{align*} \nabla^a W_{abcd} = B^{ae}\nabla_e C_{abcd} + g^{af}({P_f}^e{}_a C_{ebcd} + {P_b}^e{}_f C_{aecd} + {P_c}^e{}_f C_{abed} + {P_d}^e{}_f C_{abce} ). \end{align*} \end{lem} \textbf{Proof.} We will prove the schematic version: $$ \divergence W = B \cdot \nabla C + P \cdot C . $$ We have that \begin{align*} \nabla_e W = \mathcal{L}_T (\nabla_e C) + \Pi_e \odot C - \nabla_e (B \odot C), \end{align*} hence, \begin{align*} \divergence W &= (\mathcal{L}_T g) \cdot \nabla C + \mathcal{L}_T \divergence C + (\Pi_e - \nabla_e B) \cdot C - B\cdot \nabla C, \\ &= B \cdot \nabla C + P \cdot C , \end{align*} where we used $\pi_{ab} = B_{ab} + B_{ba}$, $\Pi_{aeb} - \nabla_e B_{ab} = P_{abe}$ and $\divergence C = 0$. $_\blacksquare$ \\ We proceed similarly to obtain the wave equation for $W$, \begin{align*} \square W &= \nabla ^e( \mathcal{L}_T \nabla_e C + \Pi_e \odot C - \nabla_e(B \odot C) ) , \\ &= \mathcal{L}_T \square C + \Pi^e \cdot \nabla_e C + \nabla ^e(\Pi_e \odot C - \nabla_e(B \odot C)) , \\ &= \mathcal{L}_T \square C + \nabla B \cdot \nabla C + \nabla^e( \Pi_e - \nabla_e B) \odot C + B\cdot \square C , \\ &= \mathcal{L}_T \square C + \nabla B \cdot \nabla C + \nabla P \cdot C + B\cdot \square C, \end{align*} where we have used once more $\Pi_{aeb} - \nabla_e B_{ab} = P_{abe}$ and $\Pi=\nabla \pi = \nabla B$ (schematically). This finishes the proof of Proposition \ref{prop:tensorial equations}. $_{\blacksquare}$ \\ Using the same argument as above we can prove the following general statement: \begin{lem} Let $F$ be a $k$-covariant tensor, then $E:= \mathcal{L}_T F- B\odot F $ obeys \begin{align} \square E = \mathcal{L}_T \square F + \nabla B \cdot \nabla F + \nabla P \cdot F + B\cdot \square F. \end{align} \end{lem} This last lemma can be applied to the Faraday tensor. To conclude this section we state the full system of equations relevant for the Carleman estimates. We start by recalling the wave equations satisfied by the Weyl and Faraday tensors, as well as the Einstein equations in schematic form: \begin{align} \square C &= R \cdot C = (C + g\cdot S) \cdot C, \\ \square F &= R \cdot F = (C + g\cdot S) \cdot F, \\ S &= F \cdot F \end{align} Hence \begin{align*} \mathcal{L}_T\square C & = (\mathcal{L}_T C + \pi \cdot F^2 + \mathcal{L}_T F \cdot F)\cdot C + R \cdot \mathcal{L}_T C , \\ & = (W + B\cdot C + B \cdot F^2 + E \cdot F + B \cdot F^2)\cdot C + R \cdot W + R\cdot B \cdot W. \end{align*} Similarly, for the modified Lie derivative of the Faraday tensor we obtain $$\mathcal{L}_T\square E = (W + B\cdot C + B \cdot F^2 + E \cdot F + B \cdot F^2)\cdot F + R \cdot E + R\cdot B \cdot F . $$ Therefore, we have proved: \begin{lem} The deformation tensors $W = \mathcal{L}_T C - B \odot C $ and $E = \mathcal{L}_T F - B\odot F$ satisfy the following wave equations \begin{align} \square W &= (R+C) \cdot W + (C^2 + F^2\cdot C + R\cdot C)\cdot B \nonumber \\ & \quad + (F\cdot C) \cdot E + \nabla C \cdot \nabla B + C\cdot \nabla P, \\ \square E &= F \cdot W + (F^3+ F\cdot C + R\cdot F)\cdot B \nonumber \\ & \quad + (F^2 + R)\cdot E + \nabla F\cdot \nabla B + F \cdot \nabla P . \end{align} \end{lem} \subsection{Estimates for the deformation tensors} Now we use the Carleman estimates of Theorem \ref{thm:Carleman} applied to the wave equations for $W$ and $E$. We start by writing them in Cartesian coordinates. This type of coordinates are chosen due to its uniform decay in all directions. We also have to check that the fields $W$ and $E$ satisfy the vanishing condition (\ref{vanishing condition}). This latter condition is fulfilled on the ``exterior'' part thanks to the hypothesis of Proposition \ref{prop:unique continuation2}, that is, the decaying properties corresponds to the fields vanishing to all orders at infinity. However in order to cope with the ``interior'' decay also included in the vanishing condition (\ref{vanishing condition}) a cut-off function needs to be introduced. Finally, by absorbing the lower order terms into the terms already present in the Carleman estimates we can conclude the vanishing of $W$, $E$, $B$ and $P$ in a neighbourhood of spatial infinity. \\ Firstly, we pass to Cartesian coordinates $(x^0, x^1, x^2, x^3)$ such that the metric takes the form $$ g = -(dx^0)^2 + (dx^1)^2 + (dx^2)^2 + (dx^3)^2 + \sum_{\mu, \nu=0}^3 \mathcal{O}_2(r^{-1}) dx^{\mu}dx^{\nu} , $$ and $\partial_{x^0}|_{x^1, x^i}$ coincides with $\partial_u|_{s,\theta^i}$ as $r\rightarrow \infty$, where $$r = \sqrt{(x^1)^2 + (x^2)^2 + (x^3)^2}. $$ Then, in these coordinates the Christoffel symbols decay suitably fast. That is, $$ \Gamma_{\mu \nu}^{\alpha} = \mathcal{O}_1(r^{-2}) \quad \textrm{and} \quad \partial_{\beta} \Gamma_{\mu \nu}^{\alpha} = \mathcal{O}(r^{-3}) .$$ In what follows, for brevity, we will denote by $(V) $ the components of a tensor $V$ with respect to these Cartesian coordinates. In particular, we have schematically \begin{align*} (\nabla V) &= \partial (V) + \Gamma \cdot (V) , \\ (\square V) &= \square (V) + \Gamma \cdot \partial (V) + \partial \Gamma \cdot (V). \end{align*} Hence, the wave equations for the components of $W$ and $E$ are \begin{align*} \square (W) & = [(R) + \partial \Gamma]\cdot (W) + \Gamma \cdot \partial (W) + [(F)\cdot (C)] \cdot (E) + [(R)\cdot (C) + \partial (C) \cdot \Gamma]\cdot (B) \\ & \quad + \partial (C) \cdot \partial (B) + (C) \cdot \partial (P) + (C)\cdot \Gamma \cdot (P) , \\ \square (E) & = [(F) + \partial \Gamma] \cdot (W) + \Gamma \cdot \partial (E) + [(F)^2 + \partial\Gamma] \cdot (E) + [(R)\cdot (F) + \partial (F) \cdot \Gamma] \cdot (B) \\ & \quad + \partial (F) \cdot \partial (B) + (F) \cdot \partial (P) + (F) \cdot \Gamma \cdot (P) , \end{align*} where we have just kept the leading order terms multiplying $W$, $E$, etc. In view of the asymptotic behaviour assumed for the Weyl and Faraday tensors we have the following estimates \begin{align} \label{coupled system} \begin{aligned} \square (W) & = \mathcal{O}(r^{-3}) (W) + \mathcal{O}(r^{-2})\partial (W) + \mathcal{O}(r^{-5}) (E) + \mathcal{O}(r^{-6}) (B) \\ & \qquad + \mathcal{O}(r^{-4}) \partial (B) + \mathcal{O}(r^{-5}) (P) + \mathcal{O}(r^{-3}) \partial (P) , \\ \square (E) & = \mathcal{O}(r^{-2}) (W) + \mathcal{O}(r^{-3}) (E) + \mathcal{O}(r^{-2})\partial (E) + \mathcal{O}(r^{-5}) (B) \\ & \qquad + \mathcal{O}(r^{-3}) \partial (B) + \mathcal{O}(r^{-4}) (P) + \mathcal{O}(r^{-2}) \partial (P) . \end{aligned} \end{align} While these are morally the reason for the unique continuation we still need to compensate for the fact that the coefficient accompanying $(W)$ in the second equation does not decay fast enough\footnote{We need a power strictly greater than 2 in order to run the Alexakis-Schlue argument, see proof below where the corresponding term is controlled by choosing different $\lambda$-weights for each Carleman inequality.}. However the coupling term in the first equation, $\mathcal{O}(r^{-5}) (E)$, allows us to borrow some decay by modifying the Carleman weight.\\ Before applying the Carleman estimates of Theorem \ref{thm:Carleman} we have to guarantee that all the quantities vanish to all orders in the sense of (\ref{vanishing condition}), $$ \lim_{k\rightarrow \infty } \int_{\partial \mathcal{U}_k} r^N(\phi^2 + |\partial\phi|^2) = 0. $$ We deal with the ``interior'' and ``exterior'' parts of the the boundary of $\mathcal{U}_k$ differently. \\ The ``exterior'' boundary approaches infinity as $r\rightarrow \infty$ and so it captures the idea of $\phi$ vanishing to all orders at infinity. By assumption the quantities $\pi =\mathcal{L}_T g$, $\mathcal{L}_T C$ and $\mathcal{L}_T F$ vanish to all orders at infinity. Now we check the modified versions. Firstly, $(\omega)$ satisfies the transport equation $$ \nabla_L (\omega) = \mathcal{O}(r^{-1})(\pi). $$ with $(\omega)=0$ at infinity by construction. Thus, $(\omega)$ vanishes to all orders at infinity. It follows that \begin{align*} (B) &= (\pi) + (\omega), \\ (P) &= \partial (\pi + \omega) + \Gamma \cdot (\pi + \omega), \\ (W) &= (\mathcal{L}_T C) + (R) \odot (B), \\ (E) &= (\mathcal{L}_T F) + (F) \odot (B), \end{align*} also vanish to all orders at infinity.\\ To deal with the ``interior'' part of $\partial \mathcal{U}_k$ a cut-off function is used. This technique is standard for unique continuation problems. \\ Let $\chi$ be a cut-off function whose level sets coincide with those of $F\circ f$, $$ \chi =1 \, \textrm{on } \mathcal{D}_{\omega_0}, \qquad \chi=0 \, \textrm{on } \mathcal{M}\setminus \mathcal{D}_{\omega_1}, \qquad \omega_0<\omega_1<\omega . $$ Then the functions $\chi \cdot (W)$, $\chi \cdot (E)$, etc., satisfy the vanishing condition (\ref{vanishing condition}). The price to pay is that we have introduced extra terms in the wave equations, however these are easy to treat since they are supported only in the cut-off region. Indeed, \begin{align*} \square (\chi\cdot (W)) &= (\square \chi) \cdot (W) + (\partial \chi) \cdot \partial (W) + \chi \cdot \square (W) , \\ &= (\square \chi) \cdot (W) + (\partial \chi) \cdot \partial (W) + (\partial \chi) \cdot \{(W), \partial(W), \ldots \} \nonumber \\ & \quad+ \{ \chi \cdot (W), \partial (\chi \cdot (W)), \ldots \} , \\ &= \nabla \chi \mathbf{M} + \{ \chi \cdot (W), \partial (\chi \cdot (W)), \ldots \} . \end{align*} Hereafter we will use the symbol $\nabla \chi \mathbf{M} $ to denote multiples of $(W)$, $(E)$, etc., which are only supported in the cut-off region $\mathcal{D}_{\omega_1} \setminus \mathcal{D}_{\omega_0}$. Recall also that we have used the notation $\{ (W), (E), \ldots \}$ to denote a function involving $(W)$, $(E)$, etc. Hence, after applying the Carleman estimates we can focus only on the terms supported on $\mathcal{D}_{\omega_0}$.\\ \\ \textbf{Weighted Carleman estimates.}\\ \\ We are now ready to apply Theorem \ref{thm:Carleman} to the functions $\chi \cdot (W)$, $\chi \cdot (E)$, etc. To keep the notation simple and readable we will omit in the next argument the cut-off function and the parenthesis denoting Cartesian components. We follow the standard procedure to bound the $L^2$-norms of $W$, $B$, $P$ and its first derivatives. \\ The Carleman estimate for $W$ combined with its wave equation reads \begin{align} \label{CarlemanW} \lambda^3 \| f^{\delta} W \|_{\mathcal{W}} + \lambda \| f^{-\frac12} \Psi^{\frac12} \partial W \|_{\mathcal{W}} & \lesssim \| f^{-1} \square W \|_{\mathcal{W}} , \nonumber \\ & \lesssim \| f^{-1} r^{-3} W \|_{\mathcal{W}} + \| f^{-1} r^{-2} \partial W \|_{\mathcal{W}} \nonumber \\ & \quad + \| f^{-1} r^{-6} B \|_{\mathcal{W}} + \| f^{-1} r^{-4} \partial B \|_{\mathcal{W}} \nonumber \\ & \quad + \| f^{-1} r^{-5} P \|_{\mathcal{W}} + \| f^{-1} r^{-3} \partial P \|_{\mathcal{W}} \nonumber \\ & \quad + \| f^{-1} r^{-5} E \|_{\mathcal{W}} + \| \nabla\chi \textbf{M} \|_{\mathcal{W}} . \end{align} The following estimates will be used throughout; they follow from the definition of $f$ and $\Psi$: \begin{align} \label{estimates f} f \gtrsim \frac{1}{r^2}, \qquad f\Psi \gtrsim \frac{1}{r^3}. \end{align} They imply \begin{align} f^{-\frac12} \Psi^{\frac12} & =f^{-1}(f\Psi)^{\frac12} \gtrsim f^{-1}r^{-\frac32}, \\ \label{estimates f2} f^{\delta} &\gtrsim \frac{1}{r^{2\delta}} > \frac{1}{r}, \qquad \textrm{for } 0<2\delta < 1 . \end{align} These last inequalities tell us that the first two terms on the right hand side of (\ref{CarlemanW}) can be absorbed into the corresponding terms on the left hand side, since $f^{-1}r^{-3} \lesssim f^{\delta}$ and $f^{-1}r^{-2} \lesssim f^{-\frac12} \Psi^{\frac12}$. This will be the main trick during the proof. \\ Next, we aim at controlling the $B$ terms. The Carleman estimate from Lemma (\ref{lem:CarlemanODE}) together with the transport equation for $B$ read \begin{align*} \lambda \|\frac1r f^{-1} r^{-4} B \|_{\mathcal{W}} & \lesssim \| f^{-1} r^{-4} \nabla_L B \|_{\mathcal{W}} , \nonumber \\ & \lesssim \| f^{-1} r^{-4} P \|_{\mathcal{W}} + \| f^{-1} r^{-5} B \|_{\mathcal{W}} + \| \nabla\chi \textbf{M} \|_{\mathcal{W}} . \end{align*} We add this inequality to (\ref{CarlemanW}) and observe that the terms $\| f^{-1} r^{-5} B \|_{\mathcal{W}}$ and $\| f^{-1} r^{-6} B \|_{\mathcal{W}}$ can be absorbed into the left hand side. We have thus obtained, \begin{align} \label{CarlemanW2} \lambda^3 \| f^{\delta} W \|_{\mathcal{W}} + \lambda \| f^{-\frac12} \Psi^{\frac12} \partial W \|_{\mathcal{W}} + \lambda\| f^{-1} r^{-5} B \|_{\mathcal{W}} & \lesssim \| f^{-1} r^{-5} E \|_{\mathcal{W}} + \| f^{-1} r^{-4} \partial B \|_{\mathcal{W}} \nonumber \\ & \quad + \| f^{-1} r^{-4} P \|_{\mathcal{W}} + \| f^{-1} r^{-3} \partial P \|_{\mathcal{W}} \nonumber \\ & \quad + \| \nabla\chi \textbf{M} \|_{\mathcal{W}} . \end{align} Now we proceed similarly to bound $P$, $\partial B$ and $\partial P$. We have \begin{align*} \lambda \|\frac1r & f^{-1} r^{-3} P \|_{\mathcal{W}} \lesssim \| f^{-1} r^{-3} \nabla_L P \|_{\mathcal{W}} , \nonumber \\ & \lesssim \| f^{-1} r^{-3} W \|_{\mathcal{W}} + \| f^{-1} r^{-6} B \|_{\mathcal{W}} + \| f^{-1} r^{-5} E \|_{\mathcal{W}} + \| f^{-1} r^{-4} P \|_{\mathcal{W}} + \|\nabla\chi \textbf{M} \|_{\mathcal{W}} , \\ \lambda \|\frac1r & f^{-1} r^{-3} \partial B \|_{\mathcal{W}} \lesssim \| f^{-1} r^{-3} \nabla_L \partial B \|_{\mathcal{W}} , \nonumber \\ & \lesssim \| f^{-1} r^{-5} B \|_{\mathcal{W}} + \| f^{-1} r^{-4} \partial B \|_{\mathcal{W}} + \| f^{-1} r^{-4} P \|_{\mathcal{W}} + \| f^{-1} r^{-3} \partial P \|_{\mathcal{W}} + \|\nabla\chi \textbf{M} \|_{\mathcal{W}} , \\ \lambda \|\frac1r &f^{-1} r^{-2} \partial P \|_{\mathcal{W}} \lesssim \| f^{-1} r^{-2} \nabla_L \partial P \|_{\mathcal{W}} , \nonumber \\ & \lesssim \| f^{-1} r^{-3} W \|_{\mathcal{W}} + \| f^{-1} r^{-2} \partial W \|_{\mathcal{W}} + \| f^{-1} r^{-6} B \|_{\mathcal{W}} + \| f^{-1} r^{-5} \partial B \|_{\mathcal{W}} \\ & \qquad + \| f^{-1} r^{-4} P \|_{\mathcal{W}} + \| f^{-1} r^{-5} E \|_{\mathcal{W}} + \| f^{-1} r^{-4} \partial E \|_{\mathcal{W}} + \|\nabla\chi \textbf{M} \|_{\mathcal{W}} . \end{align*} We add these inequalities to (\ref{CarlemanW2}) and observe that the $W$, $B$, $\partial B$, $P$ and $\partial P$ terms can be absorbed into the left hand side, thus obtaining, \begin{align} \label{CarlemanW3} \lambda^3 \| &f^{\delta} W \|_{\mathcal{W}} + \lambda \| f^{-\frac12} \Psi^{\frac12} \partial W \|_{\mathcal{W}} + \lambda\| f^{-1} r^{-5} B \|_{\mathcal{W}} + \lambda\| f^{-1} r^{-4} \partial B \|_{\mathcal{W}} \nonumber \\ & + \lambda\| f^{-1} r^{-4} P \|_{\mathcal{W}} + \lambda\| f^{-1} r^{-3} \partial P \|_{\mathcal{W}} \lesssim \| f^{-1} r^{-5} E \|_{\mathcal{W}} + \| \nabla\chi \textbf{M} \|_{\mathcal{W}} . \end{align} for sufficiently large $\lambda$ and $0<\delta<\frac12$. An analogous argument gives (note the different Carleman parameter), \begin{align} \label{CarlemanE} \lambda'^3 \| &f^{\delta} E \|_{\mathcal{W}'} + \lambda' \| f^{-\frac12} \Psi^{\frac12} \partial E \|_{\mathcal{W}'} + \lambda' \| f^{-1} r^{-5} B \|_{\mathcal{W}'} + \lambda' \| f^{-1} r^{-4} \partial B \|_{\mathcal{W}'} \nonumber \\ & + \lambda' \| f^{-1} r^{-3} P \|_{\mathcal{W}'} + \lambda' \| f^{-1} r^{-2} \partial P \|_{\mathcal{W}'} \lesssim \| f^{-1} r^{-2} W \|_{\mathcal{W}'} + \| \nabla\chi \textbf{M} \|_{\mathcal{W}'} . \end{align} We would like to add these last two inequalities and absorb the $W$ term on the left hand side to obtain the desired bound. However, for $\lambda =\lambda'$, this is not possible as $f^{-1} r^{-2}=\mathcal{O}(1)$ does not decay fast enough. To remedy this, we make the observation that the norms depend on $\lambda$ and by taking slightly different weights we can perform the procedure just described. More precisely, we want to find $\lambda'$ such that \begin{align} \label{inequalities} \begin{aligned} (e^{-\lambda F}f^{\frac12})f^{-1} r^{-5} & \lesssim e^{-\lambda' F} f^{\frac12} f^{\delta}, \\ (e^{-\lambda' F}f^{\frac12})f^{-1} r^{-2} & \lesssim e^{-\lambda F} f^{\frac12} f^{\delta}. \end{aligned} \end{align} Indeed, we will show that the choice $\lambda' := \lambda - \delta$ achieves the previous inequalities. Firstly, note that \begin{align*} r^{-5} & \lesssim f^{1+2\delta}, \\ r^{-2} & \lesssim f, \end{align*} these are a consequence of estimates (\ref{estimates f}) and (\ref{estimates f2}). They imply that \begin{align*} f^{-1} r^{-5} & \lesssim e^{\delta F} f^{\delta}, \\ e^{\delta F} f^{-1} r^{-2} & \lesssim f^{\delta}, \end{align*} since $F=\mathcal{O}(\log f) $. Finally, it is easy to see that these last inequalities are equivalent to (\ref{inequalities}), with $\lambda' := \lambda - \delta$. \\ \\ \textbf{Remark.} It is worth noticing that the previous argument did not make any special use of the power $r^{-5}$ accompanying $E$. The procedure will work for any $r^{-q}$ with $q>2$ by choosing $\delta>0$ small enough. \\ Now we are in position to close the argument. We add inequalities (\ref{CarlemanW3}) and (\ref{CarlemanE}), with $\lambda'= \lambda-\delta$. Inequalities (\ref{inequalities}) ensure that the terms $\| f^{-1} r^{-5} E \|_{\mathcal{W}}$ and $\| f^{-1} r^{-2} W \|_{\mathcal{W}'}$ can be absorbed into the left hand side. Moreover, the Carleman weights $e^{-\lambda F}$ and $e^{-\lambda' F}$ are monotonic increasing functions, so we can substitute its minimum value on left hand side and its maximum value on the right hand side (since these terms are supported in the cut-off region). Thus, after dropping the weight factors from the inequality we obtain the desired $L^2$-bound, \begin{align*} \lambda^3 \| f^{\frac12} f^{\delta} W \|_2 &+ \lambda \| \Psi^{\frac12} \partial W \|_2 + \lambda\| f^{\frac12} f^{-1} r^{-5} B \|_2 + \lambda\| f^{\frac12} f^{-1} r^{-4} \partial B \|_2 \\ & + \lambda\| f^{\frac12} f^{-1} r^{-4} P \|_2 + \lambda\| f^{\frac12} f^{-1} r^{-3} \partial P \|_2 \\ & + \lambda'^3 \| f^{\frac12} f^{\delta} E \|_2 + \lambda' \| \Psi^{\frac12} \partial E \|_2 + \lambda' \| f^{\frac12} f^{-1} r^{-5} B \|_2 + \lambda' \| f^{\frac12} f^{-1} r^{-4} \partial B \|_2 \\ & + \lambda' \| f^{\frac12} f^{-1} r^{-3} P \|_2 + \lambda' \| f^{\frac12} f^{-1} r^{-2} \partial P \|_2 \\ & \qquad \lesssim \| \nabla\chi \textbf{M} \|_2 + \| \nabla\chi \textbf{M} \|_2. \end{align*} Finally, the left hand side can be interpreted as integrated over the smaller domain $\mathcal{D}_{\omega_0}$ where $\chi=1$ and by taking $\lambda \rightarrow \infty$ we conclude that $B \equiv 0$, $P \equiv 0$, $W\equiv 0$ and $E\equiv 0$ on $\mathcal{D}_{\omega_0}$. In particular $$ \mathcal{L}_T g \equiv 0 \qquad \textrm{and} \qquad \mathcal{L}_T F \equiv 0 \qquad \textrm{on } \mathcal{D}_{\omega_0} .$$ This finishes the proof of Proposition \ref{prop:unique continuation2} and Theorem \ref{thm1}. $_{\blacksquare}$
c88c3bed778ebeda0f8f4a74ad2eb71085d86475
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Cycle graphs}\label{sec:cycles} We consider the family of graphs consisting of one cycle of length $\leq n$ (and no other edges or vertices). We discuss both of the cases where the decoder is aware and oblivious of the value of $n$, as discussed in \Cref{pro:aware}. In particular we show that oblivious decoding requires a larger induced universal graph for this problem. Our new bounds leave small gaps which are interesting open problems to tighten. \input{lowerbounds} \input{upperbounds} \section{General $D$} \newcommand{\mathcal{G}}{\mathcal{G}} In this section we present two upper bounds on $g_v(\mathcal{G}_D)$, the number of nodes in the smallest induced universal graph for graphs on $n$ nodes with bounded degree $D$. In \Cref{thmDetUpper} we give a deterministic construction of an induced universal graph for $\mathcal{G}_D$ that relies on the induced universal graph constructed in \Cref{sec:max2upper}. In \Cref{thmRandUpper} we give a randomized construction of an induced universal graph for $\mathcal{G}_D$ that with probability $\frac{1}{2}$ has a small number of nodes. Combining the two results shows the existence of an adjacency labeling scheme for $\mathcal{G}_D$ of size $\log \binom{\floor{n/2}|}{\floor{D/2}} + O\!\left(\min\set{D+\log n,\sqrt{D\log n}\log(n/d)}\right)$. In \Cref{corLowerBoundLabelSizeBoundedDeg} and \Cref{corRandLowerBound} we give lower bounds on $g_v(\mathcal{G}_D)$. These lower bounds imply that any adjacency labeling scheme for $\mathcal{G}_D$ must have labels of size at least $\log \binom{\floor{n/2}|}{\floor{D/2}} - O\!\left(\min\set{D,\sqrt{D\log n}\log(n/d)}\right)$, which means that the upper bounds are tight up to an additive term of size $O\!\left(\min\set{D+\log n,\sqrt{D\log n}\log(n/d)}\right)$, which is at most $O(\sqrt{n \log n})$. Previous labeling schemes use labels that are larger by an additive term of size $\Omega(D)$, which is $\Omega(n)$ when $D = \Omega(n)$, so this is the first adjacency labeling scheme for $\mathcal{G}_D$ where the dominating term is optimal. \subsection{Upper bounds on $g_v(\mathcal{G}_D)$} We show the following deterministic bound. \begin{theorem} \label{thmDetUpper} For the family $\mathcal{G}_D$ of graphs with bounded degree $D$ on $n$ nodes \[ g_v(\mathcal{G}_D) \le 2^{k+1} \cdot \frac{n^k}{(k-1)!} , \ \ \text{where} \ k = \ceil{D/2} \] \end{theorem} \begin{proof} For a set $S$ we let $S^{\le k}$ denote the set of all subsets of $S$ of size $\le k$. We note that $\abs{S^{\le k}} \le 2\frac{\abs{S}^k}{k!}$ whenever $S$ is finite. We will show that $g_v(\mathcal{G}_D) \le 2\frac{(2n-1)^k}{(k-1)!}$. Fix $n,D$, let $k = \ceil{D/2}$ and let $U_n$ be the induced universal graph for $\mathcal{G}_2$ defined in \Cref{sec:max2upper}. We note that $V[U_n] = [2n-1]$. We define the graph $G$ to have vertex set $[2n-1] \times [2n-1]^{\le k-1}$ and such that there is an edge between $(x,A)$ and $(y,B)$ iff $x \in B$, $y \in A$ or $x$ and $y$ are adjacent in $U_n$. Since $G$ has the desired number of nodes we proceed to show that $G$ is an induced universal graph for $\mathcal{G}_D$. Let $H$ be a graph in $\mathcal{G}_D$. By \Cref{Butlersplit} we know that we can decompose the edges of $H$ into $H_0$ and $H_1$ such that $\Delta(H_0) \le 2, \Delta(H_1) \le 2(k-1)$. We can find an embedding function $f : V[H] \to V[U_n]$ of $H_0$ in $U_n$ by the universality of $U_n$. By the same argument as in the first part of the proof we can orient the edges of $H_1$ such that any node has at most $k-1$ outgoing edges in $H_1$. For $u \in V[H]$ let $S_u$ be the set of nodes $v$ such that there exists an edge between $u$ and $v$ in $H_1$ oriented from $u$ to $v$. We see that $u$ and $v$ are adjacent iff $f(u)$ and $f(v)$ are adjacent in $U_n$ or it holds that $u \in S_v$ or $v \in S_u$. Therefore $\lambda : V[H] \to V[G]$ defined by $u \to (f(u), f(S_u))$ is an embedding function of $H$ in $G$. Hence $G$ must be an induced universal graph for $\mathcal{G}_D$. \end{proof} The intuition behind the randomized bound below is the following. Consider placing all $n$ vertices on a circle in a randomly chosen order and rename the vertices with indices $[n]$ following the order on the circle. Now, a vertex $v \in [n]$ remembers its neighbours in the next half of the circle, i.e., $v$ stores all the adjacant vertices among $\{v+1, \ldots, v+\lceil n/2 \rceil\}$ (where indices are taken modulu $n$). If two vertices $u,v$ are adjacant, then clearly either $u$ stores the index of $v$ or conversely, hence an adjacancy query can be answered. A Chernoff bound implies that vertex $v$ with high probability stores at most $D/2 + O(\sqrt{D \log n})$ indices. It follows that there exists an order of the points on the circle where every vertex stores that many neighbours and the theorem follows. \begin{theorem}\footnote{In the full version we improve this to $\binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{O \!\left (\sqrt{D\log D} \cdot \log(n/D) \right )}$} \label{thmRandUpper} For the family $\mathcal{G}_D$ of graphs with bounded degree $D$ on $n \ge 2D$ nodes \[ g_v(\mathcal{G}_D) \le \binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{O \!\left (\sqrt{D\log n} \cdot \log(n/D) \right )} \] \end{theorem} \begin{proof} Fix $n,D$ and wlog assume that $n$ is odd. For $D \le \log n$ the result follows from \Cref{thmDetUpper} so assume that $D \ge \log n$. Let $G$ be a graph in $\mathcal{G}_D$, and wlog assume that $V[G] = [n]$. Let $\pi : [n] \to [n]$ be a permutation of $[n]$ chosen uniformly at random. For each $u \in V[G]$ let $S_u$ be the set of differences $\pi(v) - \pi(u) \bmod n$ where $v$ is a neighbour of $u$ and $\pi(v) - \pi(u) \bmod n$ is at most $\floor{\frac{n}{2}}+1$. That is: \[ S_u = \set{(\pi(v) - \pi(u)) \bmod n \mid (u,v) \in E[G], \ (\pi(v) - \pi(u)) \bmod n \in \set{1,2,\ldots,\floor{\frac{n}{2}}}} \] Given two nodes $u,v$ we can determine whether $u$ are adjacent from $\pi(u),\pi(v)$ and $S_u,S_v$ in the following way. If $(\pi(u)-\pi(v)) \bmod n \le \floor{\frac{n}{2}}+1$ they are adjacent iff $(\pi(u)-\pi(v)) \bmod n \in S_v$. Otherwise they are adjacent iff $(\pi(v)-\pi(u)) \bmod n \in S_u$. We note that $\mathbb{E}(\abs{S_u}) = \frac{\deg_G(u)}{2} \le \frac{D}{2}$. By a standard Chernoff bound without replacement we see that \begin{align} \label{eqNeighbourBound} \abs{S_u} \le D', \ \ \text{where} \ D' = \floor{\frac{D}{2} + O \!\left ( \sqrt{D\log n} \right )} \end{align} with probability $\ge 1 - \frac{1}{2n}$ for a given vertex $u \in V[G]$. So with probability at least $\frac{1}{2}$ we have that \eqref{eqNeighbourBound} holds for every $u \in V[G]$. In particular there exists $\pi$ such that \eqref{eqNeighbourBound} holds for every $u \in V[G]$. Fix such a $\pi$. Let $D'' = \min\set{\floor{\frac{n}{2}},D'}$. Then for any node $u$ we can encode $\pi(u)$ and $S_u$ using at most $O(\log n) + \log \binom{\floor{n/2}}{D''}$ bits. Hence we conclude that: \[ g_v(\mathcal{G}_D) \le \binom{\floor{n/2}}{D''} n^{O(1)} \] The conclusion now follows from the following estimate \[ \binom{\floor{n/2}}{D''} \le \binom{\floor{n/2}}{\floor{D/2}} \cdot \left ( \frac{\floor{n/2}}{\floor{D/2}} \right )^{D''-\floor{D/2}} \le \binom{\floor{n/2}}{\floor{D/2}} \cdot \left ( \frac{n}{D} \right )^{O\!\left(\sqrt{D\log n}\right )} \] \end{proof} \subsection{Lower bounds on $g_v(\mathcal{G}_D)$} We now show lower bounds on $g_v(\mathcal{G}_D)$. Our first lower bound follows from counting perfect matchings. \begin{lemma} \label[lemma]{lemBipartiteBoundedDegLowerBound} Let $n, D$ be positive integers where $n$ is even. Let $V = [n]$. The number of graphs $G$ with $\Delta(G) \le D$ and vertex set $V$ is at least $\frac{\left((n/2)!\right)^{D}}{D^{Dn/2}}$. \end{lemma} \begin{proof} Let $V_0 = [n/2], V_1 = [n] \setminus [n/2]$. Let $M_0,M_1,\ldots,M_{r-1}$ be all perfect matchings of $V_0$ and $V_1$ where $r = (n/2)!$. Now consider the following family of graphs being the union of $D$ such perfect matchings: \[ \mathcal{F} = \set{G \mid V[G] = V, E[G] = M_{i_0} \cup \ldots M_{i_{D-1}}, i_0,\ldots,i_{D-1} \in [r]} \] Every graph in $\mathcal{F}$ is the union of $D$ perfect matchings and therefore has max degree $\le D$. Now fix $G \in \mathcal{F}$ and let $M$ be a perfect matching $G$. We can write $M = \set{(u,f(u)) \mid u \in V_0}$ for some bijective function $f : V_0 \to V_1$. There are at most $D$ ways to choose $f(u)$ for every $u \in V_0$ since $(u,f(u))$ must be an edge of $G$. Hence there are at most $D^{n/2}$ ways to choose a perfect matching of $G$, and $G$ can be written as a union of $D$ perfect matchings in at most $D^{Dn/2}$ ways. Since $G$ was arbitrarily chosen this must hold for any $G \in \mathcal{F}$. Since there are $r^D$ ways to choose $D$ perfect matchings we conclude that $\mathcal{F}$ consists of at least $\frac{r^D}{D^{nD/2}}$ graphs as desired. \end{proof} As an immediate corollary of \Cref{lemBipartiteBoundedDegLowerBound} we get a lower bound on the number of nodes in an induced universal graph, shown below in \cref{corLowerBoundLabelSizeBoundedDeg}. \begin{corollary} \label[corollary]{corLowerBoundLabelSizeBoundedDeg} The induced universal graph for the family $\mathcal{G}_D$ of graphs with bounded degree $D$ and $n$ nodes has at least $\Omega \!\left ( \left ( \frac{n}{2eD} \right )^{D/2} \right )$ nodes. \end{corollary} \begin{proof} Let $G$ be the induced universal graph for the family $\mathcal{G}_D$. Let $V = [n]$. Any graph $H$ from $\mathcal{G}_D$ on the vertex set $V$ is uniquely defined by the embedding function $f$ of $H$ in $G$. Since there are no more than $\abs{V[G]}^n$ ways to choose $f$, \Cref{lemBipartiteBoundedDegLowerBound} gives that $\abs{V[G]}^n \ge \frac{\left(\floor{n/2}!\right)^{D}}{D^{D\floor{n/2}}}$. The result now follows from Stirling's formula. \[ \abs{V[G]} \ge \left ( \frac{\left(\floor{n/2}!\right)^{2/n}}{D^{\floor{n/2}/(n/2)}} \right )^{D/2} \] We note that $\floor{n/2}/(n/2) = 1$ when $n$ is even. When $n$ is odd we have $\floor{n/2} = \frac{n-1}{2}$. Hence $\floor{n/2}/(n/2) = 1 - \frac{1}{n}$. Since $D \le n$ we have $D^{1-\frac{1}{n}} = \Theta(D)$. \end{proof} Our second lower bound comes from bounding the probability that a random graph on $n$ vertices, where each edge exists with probability around $D/n$, has max degree $D$. \begin{lemma} \label{lemRandLowerBound} Let $n, D$ be positive integers where $n \ge 2D$. Let $V = [n]$. The number of graphs $G$ with $\Delta(G) \le D$ and vertex set $V$ is at least $\binom{\binom{n}{2}}{\floor{nD/2}} \cdot 2^{-O \!\left (n\sqrt{D\log n} \cdot \log(n/D) \right )}$. \end{lemma} \begin{proof} Fix $n, D$. For $D \le \log^2 n$ the result follows from \Cref{lemBipartiteBoundedDegLowerBound}, so assume that $D \ge \log^2 n$. Let $D' = D - O\!\left(\sqrt{D\log n}\right)$ be an integer. Let $G$ be a random $G(n,p)$ graph where $p = \frac{D'}{n-1}$ and $V[G] = [n]$. That is, $G$ is a random graph on $n$ nodes and for every pair $u,v \in V[G]$ there is an edge between $u$ and $v$ with probability $p$. We say that $G$ is \emph{good} if it satisfies the following two properties: \begin{enumerate} \item[1] $\Delta(G) \le D$. \item[2] $\abs{E[G]} \ge nD''$ where $D''$ is an even integer satisfying $\frac{nD''}{2} = \frac{nD'}{2} - O(\sqrt{nD})$. \end{enumerate} We note that $D'' = D - O(\left(\sqrt{D\log n}\right)$. We will argue that $G$ satisfies Property 1 with probability at least $\frac{1}{3}$. By a Chernoff bound the probability that $u \in V[G]$ has more than $D$ neighbours is at most $\frac{1}{3n}$ if $D'$ is chosen sufficiently small. So with probability at least $\frac{2}{3}$ we have $\Delta(G) \le D$. Similarly, with probability at least $\frac{2}{3}$ we have $\abs{E[G]} \ge \frac{nD'}{2} - O(\sqrt{nD})$ if we choose the constant in the $O$-notation large enough. So with probability at least $\frac{1}{3}$ $G$ is good. Let $r$ be the number of good graphs and enumerate them $G_1,G_2,\ldots,G_r$. The probability that $G = G_i$ is $p^{\abs{E[G_i]}}(1-p)^{\binom{n}{2}-\abs{E[G_i]}}$. Since $G_i$ is good we know that $\abs{E[G_i]} \ge \frac{nD''}{2}$. Hence the probability is at most: \[ p^{nD''/2}(1-p)^{\binom{n}{2}-nD''/2} \le \binom{\binom{n}{2}}{nD''/2}^{-1} \] Where the inequality follows from the binomial expansion of $(p+(1-p))^{\binom{n}{2}}$. Hence we see that: \[ \frac{1}{3} \le \sum_{i=1}^r \Pr(G=G_i) \le r\binom{\binom{n}{2}}{\frac{nD''}{2}}^{-1} \] And hence there are at least $\frac{1}{3}\binom{\binom{n}{2}}{nD''/2}$ graphs with vertex set $[n]$ and maximum degree $\le D$. Now the result follows from the following estimate: \[ \binom{\binom{n}{2}}{\frac{nD''}{2}} \ge \binom{\binom{n}{2}}{\floor{\frac{nD}{2}}} \left ( \frac{\binom{n}{2}}{nD''/2} \right )^{nD''-nD} \ge \binom{\binom{n}{2}}{\floor{\frac{nD}{2}}} \left ( \frac{n}{D} \right )^{-O\!\left(n\sqrt{D \log n}\right)} \] \end{proof} As previously we get a bound on $g_v(\mathcal{G}_D)$. \begin{corollary} \label[corollary]{corRandLowerBound} For the family $\mathcal{G}_D$ of graphs with bounded degree $D$ on $n \ge 2D$ nodes \[ g_v(\mathcal{G}_D) \ge \binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{-O \!\left (\sqrt{D\log n} \cdot \log(n/D) \right )} \] \end{corollary} \begin{proof} By the same argument as for \Cref{corLowerBoundLabelSizeBoundedDeg} we get that \Cref{lemRandLowerBound} implies: \[ g_v(\mathcal{G}_D) \ge \binom{\binom{n}{2}}{\floor{nD/2}}^{1/n} \cdot 2^{-O \!\left (\sqrt{D\log n} \cdot \log(n/D) \right )} = \binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{-O \!\left (\sqrt{D\log n} \cdot \log(n/D) \right )} \] \end{proof} \section{Introduction} A graph $G=(V, E)$ is said to be an \emph{induced universal graph} for a family $\cal F$ of graphs if it contains each graph in $\cal F$ as a vertex-induced subgraph. A graph $H=(V',E')$ is contained in $G$ as a \emph{vertex-induced subgraph} if $V' \subseteq V$ and $E'=\{vw\mid v,w \in V' \wedge vw \in E\}$. Induced universal graphs have been studied since the 1960s~\cite{moon1965minimal,Rado64}, and bounds on the sizes of induced universal graphs have been given for many families of graphs, including general, bipartite~\cite{AlstrupKTZ14}, and bounded arboricity graphs~\cite{adjacencytrees2015}. We later define the classic distributed data structure \emph{adjacency labeling scheme} and describe how it is directly related to induced universal graphs. In Table~\ref{tab:adjacency2} in \Cref{sec:overview} below we give an overview of previous results and results in this paper. \subsection{Overview of new and existing results}\label{sec:overview} We give an overview in Table~\ref{tab:adjacency2} of dominating existing and new results. All bounds are on sizes of induced universal graphs. In the table, ``P'' refers to a result in this paper, $k=\ceil{D/2}$, $L=(\sqrt{D\log n} \cdot \log(n/D))$ and $U=(\sqrt{D\log n} \cdot \log(n/D))$. The ``A'' and ``B'' case below represent two different constructions. The upper bound in ``B'' is a randomized construction, whereas both lower bounds hold for both upper bounds. \begin{table*}[ht] \renewcommand{\arraystretch}{1.3} \small \centering \makebox[0pt][c]{ \begin{tabular}{|c|c|c|c|} \hline \hline \bf Graph family & \bf Lower bound & \bf Upper bound & \bf Lower/Upper\\ \hline \noalign{\vskip 2mm} \hline General& $2^{\frac{n-1}{2}}$ & $O( 2^{\frac{n}{2}})$ & \cite{moon1965minimal}/ \cite{AlstrupKTZ14} \\ \hline Tournaments & $2^{\frac{n-1}{2}}$ & $O( 2^{\frac{n}{2}})$ & \cite{moon1968topics}/\cite{AlstrupKTZ14} \\ \hline Bipartite& $\Omega(2^{\frac{n}{4}})$ & $O( 2^{\frac{n}{4}})$ & \cite{Lozin2007}/\cite{AlstrupKTZ14} \\ \hline \noalign{\vskip 2mm} \hline A: Max degree $D$ & $\Omega((\frac{n}{2eD})^{D/2})$ & $O\!\left(\frac{\min (n,k2^k)}{k!}n^k \right)$ & P/P and ~\cite{icalpnoy14} \\ \hline B: Max degree $D$ & $\binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{-O(L)}$ & $ \binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{O(U)}$ & P/P \\ \hline Max degree $D = o\!\left(\sqrt{n}\right)$ or $D \ge \frac{(2/3+\Omega(1))n}{\ln n}$ & $\binom{\floor{n/2}}{\floor{D/2}}\cdot n^{-O(1)}$ & & \cite{mckay1990asymptotic,mckay1991asymptotic} \\ \hline Constant odd degree $D$ & $\Omega(n^{\frac{D}{2}})$ & $O(n^{k-\frac{1}{D}})$ & \cite{Butler_induced-universalgraphs}/\cite{privatealon,AlonCapalbo2008,Esperet2008} \\ \hline Max degree 2& $11 \floor{n/6}$ & $2n-1$& \cite{Esperet2008}/P \\ \hline Acyclic, max degree 2& $\floor{3/2n}$ & $\floor{3/2n}$& P/P \\ \hline A cycle aware of $n$& $n+ \Omega(\log \log n)$ & $n+\log n + O(1)$ & P/P \\ \hline A cycle not aware of $n$& $n + \Omega\!\left(\sqrt[3]{n}\right)$ & $n+ O(\sqrt{n})$ & P/P \\ \hline \noalign{\vskip 2mm} \hline Excluding a fixed minor & $\Omega(n)$ & $n^2 (\log {n})^{O(1)} $ & \cite{gavoille2007shorter} \\ \hline Planar& $\Omega(n)$ &$n^2(\log n)^{O(1)}$ & \cite{gavoille2007shorter} \\ \hline Planar, constant degree & $\Omega(n)$ & $O(n^2)$& \cite{Chung90} \\ \hline Outerplanar & $\Omega(n)$ & $n (\log n)^{O(1)}$ & \cite{gavoille2007shorter} \\ \hline Outerplanar, constant degree & $\Omega(n)$& $O(n)$& \cite{Chung90}\\ \hline \noalign{\vskip 2mm} \hline Treewidth $l$ & $n2^{\Omega(l)}$& $n (\log \frac{n}{l})^{O(l)} $& \cite{gavoille2007shorter}\\ \hline Constant arboricity $l$ & $\Omega(n^l)$ & $O(n^l)$ & \cite{alstruprauhe}/\cite{adjacencytrees2015} \\ \hline \end{tabular} } \caption{Induced-universal graphs for various families of graphs. ``P'' is results in this paper. For the max degree results $k=\ceil{D/2}$. In the result for families of graphs with an excluded minor, the~$O(1)$ term in the exponent depends on the fixed minor excluded.} \label[table]{tab:adjacency2} \end{table*} \subsection{Maximum degree $2$ and maximum degree $D$} Let $g_v(\mathcal{F})$ be the smallest number of vertices in any induced universal graph for a family of graphs $\mathcal{F}$. Let $\mathcal{G}_D$ be the family of graphs with $n$ vertices and maximum degree $D$. In the families of graphs we study in this paper, a graph always has $n$ vertices, unless explicitly stated otherwise. {\bf {\large Maximum degree $2$.}} Butler~\cite{Butler_induced-universalgraphs} shows that $g_v(\mathcal{G}_2)\leq 6.5n$. That was subsequently improved to $g_v(\mathcal{G}_2)\leq 2.5n+O(1)$ by Esperet {\em et al.}~\cite{Esperet2008} who also show the lower bound $g_v(\mathcal{G}_2)\geq 11 \floor{n/6}$. Esperet {\em et al.}~\cite{Esperet2008} conjecture that $g_v(\mathcal{G}_2)\leq 2n+o(n)$ and raise as an open problem to prove or disprove this. We show the correctness of the conjecture by proving $g_v(\mathcal{G}_2)\leq 2n-1$. The $11 \floor{n/6}$ lower bound is based on a family of graphs whose largest component has $3$ vertices. We show matching $\frac{11}{6}n+\Oo(1)$ upper bounds for the family of graphs in $\mathcal{G}_2$ whose largest component is sufficiently small ($\leq6$ vertices), and for the family of graphs in $\mathcal{G}_2$ whose smallest component is sufficiently large ($\geq10$ vertices). {\bf {\large Maximum degree $D$.}} Let $k=\ceil{D/2}$. To give an upper bound for any value of $D$, Butler~\cite{Butler_induced-universalgraphs} first establishes: \begin{corollary}[\cite{Butler_induced-universalgraphs}]\label[corollary]{Butlersplit} Let $G \in \mathcal{G}_D$ be a graph on $n$ vertices with maximum degree $D$. Then $G$ can be decomposed into $k$ edge disjoint subgraphs where the maximum degree of each subgraph is at most $2$. \end{corollary} To achieve an upper bound this can be combined with: \begin{theorem}[\cite{Chung90}] \label[theorem]{ChungSplit} Let $\mathcal{F}$ and $\mathcal{Q}$ be two families of graphs and let $G$ be an induced universal graph for $\mathcal{F}$. Suppose that every graph in the family $\mathcal{Q}$ can be edge-partitioned into $\ell$ parts, each of which forms a graph in $\mathcal{F}$. Then $g_v(\mathcal{Q}) \leq |V[G]|^\ell$. \end{theorem} Butler~\cite{Butler_induced-universalgraphs} concludes $g_v(\mathcal{G}_D)\leq (6.5n)^k $. Similarly Esperet {\em et al.}~\cite{Esperet2008} achieve $g_v(\mathcal{G}_D)\leq (2.5n+O(1))^k$, and we achieve $g_v(\mathcal{G}_D)\leq (2n-1)^k=O(2^k n^k)$. For constant maximum degree $D$, Butler~\cite{Butler_induced-universalgraphs} also shows $g_v(\mathcal{G}_D)=\Omega(n^{D/2})$. When $D$ is even and constant, the bounds are hence very tight: $g_v(\mathcal{G}_D)=\Theta(n^{D/2})$. However, for non-constant $D$ we can, using another approach but still building on top of our maximum degree $2$ solution, beat Butler's lower bound for constant degree: For any value of $D$, we prove the upper bound $g_v(\mathcal{G}_D)=O\!\left(\frac{k2^k}{k!}n^k \right)$. We also give a lower bound for any value of $D$: $\Omega\!\left((\frac{n}{2eD})^{D/2}\right)$. {\bf {\large Constant odd degree.}} A \emph{universal} graph for a family of graphs ${\cal F}$ is a graph that contains each graph from ${\cal F}$ as a subgraph (not necessarily vertex induced). The challenge is to construct universal graphs with as few edges as possible. A graph has \emph{arboricity} $k$ if the edges of the graph can be partitioned into at most $k$ forests. Graphs with maximum degree $D$ have arboricity bounded by $\floor{\frac{D}{2}}+1$~\cite{chartrand68,Lovasz66}. When $D$ is odd and constant, some improvement have been achieved \cite{AlonCapalbo2008,Esperet2008} on the above bounds on $g_v(\mathcal{G}_D)$ by arguments involving universal graphs and graphs with bounded arboricity. Let $\mathcal{A}_k$ denote a family of graphs with arboricity at most $k$. \begin{theorem}[\cite{Chung90}] \label[theorem]{Arboricity} Let $G$ be a universal graph for $\mathcal{A}_k$ and $d_i$ the degree of vertex $i$ in $G$. Then $g_v(\mathcal{A}_k) \leq \sum_{i}(d_i+1)^k$. \end{theorem} Alon and Capalbo~\cite{alon2007sparse} describes a universal graph with $n$ vertices of maximum degree $c(D)n^{1-2/D}\log^{4/D}n$ for the family $\mathcal{G}_D$, where $D\geq 3$ and $c(D)$ is a constant. Using this bound in Theorem~\ref{Arboricity}, Esperet {\em et al.}~\cite{Esperet2008} note that for odd $D$ (and hence arboricity $k=\ceil{\frac{D}{2}}$), we get $g_v(\mathcal{G}_D)\leq c_1(D)n^{k-\frac{1}{D}}\log^{2+\frac{2}{D}}n$, for a constant $c_1(D)$.\footnote{In~\cite{Esperet2008} a typo states that the maximum degree for the universal graph in~\cite{alon2007sparse} is $c(D)n^{2-2/D}\log^{4/D}n$. The theorem in~\cite{alon2007sparse} only states the total number of edges being $c(D)n^{2-2/D}\log^{4/D}n$, however the maximum degree is $c(D)n^{1-2/D}\log^{4/D}n$~\cite{privatealon}.} Using the slightly better universal graphs from~\cite{AlonCapalbo2008} the maximum degree is reduced to $c(D)n^{1-2/D}$~\cite{privatealon}, giving $g_v(\mathcal{G}_D)\leq c_2(D)n^{k-\frac{1}{D}}$, for a constant $c_2(D)$. Note that using this technique for even values of $D$ would give $g_v(\mathcal{G}_D)\leq c_3(D)n^{\frac{D}{2}+1-\frac{2}{D}}$, for a constant $c_3(D)$, which is asymptotically worse even for constant values of $D$, compared to any of the new upper bounds presented in this paper. In~\cite{alon2010universality} it is stated that the methods in~\cite{AlonCapalbo2008} can be used to achieve $g_v(\mathcal{G}_D)=O(n^{D/2})$ for constant odd values of $D>1$, however according to~\cite{privatealon} this still has to be checked more carefully, and the hidden constant in the $O$-notation is not small. \subsection{Adjacency labeling schemes and induced universal graphs} An \emph{adjacency labeling scheme} for a given family $\mathcal{F}$ of graphs assigns \emph{labels} to the vertices of each graph in $\mathcal{F}$ such that a \emph{decoder} given the labels of two vertices from a graph, and no other information, can determine whether or not the vertices are adjacent in the graph. The labels are assumed to be bit strings, and the goal is to minimize the maximum label size. A $b$-bit labeling scheme uses at most $b$ bits per label. Information theoretical studies of adjacency labeling schemes go back to the 1960s~\cite{Breuer66,BF67}, and efficient labeling schemes were introduced in~\cite{KNR92,muller}. For graphs with bounded degree $D$, it was shown in~\cite{BF67} that labels of size $2nD$ can be constructed such that two vertices are adjacent whenever the Hamming distance~\cite{hamming} of their labels is at most $4D-4$. A labelling scheme for $\mathcal{F}$ is said to have \emph{unique labels} if no two vertices in the same graph from $\mathcal{F}$ are given the same label. \begin{theorem}[\cite{KNR92}] \label[theorem]{KNRreduction} A family $\mathcal{F}$ of graphs has a $b$-bit adjacency labeling scheme with unique labels iff $g_v(\mathcal{F}) \leq 2^b$. \end{theorem} From a labeling perspective the above new upper and lower bounds are at most an additive $O(D+\log n)$ term from optimality. \subsection{Better bounds for larger $D$, $D=\Omega(\log^3 n)$} We have another approach which for large $D$, $D=\Omega(\log^3 n)$, gives better bounds than the ones presented above for constant $D$. The previous best upper bound for such large $D$ was $\binom{n}{\ceil{D/2}} n^{O(1)}$ due to Adjiashvili and Rotbart~\cite{icalpnoy14}. For any $D$ we prove the lower and upper bounds \[\binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{-O \!\left (\sqrt{D\log n} \cdot \log(n/D) \right )} \text{ and } \binom{\floor{n/2}}{\floor{D/2}} \cdot 2^{O \!\left (\sqrt{D\log n} \cdot \log(n/D) \right )},\] where the upper bound is a randomized construction. From a labeling perspective our bounds are the first to give labels for any value of $D$ that are at most $o(n)$ bits longer than the shortest possible labels. An asymptotic enumeration of the number of $D$-regular graphs due to McKay and Wormald \cite{mckay1990asymptotic,mckay1991asymptotic} combined with Stirling's formula gives a stronger lower bound of $\binom{\floor{n/2}}{\floor{D/2}} \cdot n^{-O(1)}$ whenever $D = o\left(\sqrt{n}\right)$ or $D > \frac{cn}{\ln n}$ for a constant $c > \frac{2}{3}$. \subsection{Acyclic graphs and cycle graphs} On our way to understand the family $\mathcal{G}_2$ better, we first examine two other basic families of graphs. For the family $\mathcal{AC}$ of acyclic graphs on $n$ vertices with maximum degree $2$ we show an upper bound matching exactly the lower bound in~\cite{Esperet2008}, which makes us conclude $g_v(\mathcal{AC})=\floor{3/2n}$. This lower bound is not explicitly stated in~\cite{Esperet2008}, but follows directly from the construction of the lower bound for $g_v(\mathcal{G}_2)$. We also study the family $\mathcal{C}_n$ of graphs consisting of one cycle of length $\leq n$ (and no other edges or vertices). For this family we show $n+ \Omega(\log \log n)\leq g_v(\mathcal{C}_n) \leq n+\log n + O(1)$. \subsection{Oblivious decoding} From a labeling perspective one can assume all labels have the same length~\cite{AlstrupKTZ14}. Hence, if the decoder does not know what $n$ is, it will always be able to compute $n$ approximately. However, we show that the label size in an optimum labeling scheme can be smaller if the decoder knows $n$ precisely. To be more specific, let $\mathcal{F}_n$ be a family of graphs for each $n=1,2,\ldots$. We show that there is a labeling scheme of $\mathcal F_n$ using $f(n)$ labels that enables a decoder not aware of $n$ to answer adjacency queries iff there is a family of graphs $G_1,G_2,\ldots$ such that $G_n$ is an induced universal graph for $\mathcal F_n$, $|G_n|=f(n)$, and $G_n$ is an induced subgraph of $G_{n+1}$ for every $n$. Next we show that with this extra requirement to the induced universal graph we have $n + \Omega\!\left(\sqrt[3]{n}\right) \leq g_v(\mathcal{C}_n) \leq n+ O(\sqrt{n})$. The lower bound is true for infinitely many $n$, but for specific $n$ it might not hold. For the other problems studied in this paper, the decoder does not need to know $n$, but the lower bounds hold even if it does. To the best of our knowledge this is the first time this relationship between labeling schemes and induced universal graphs has been described and examples have been given where the complexities differ. \subsection{Related results} For the family of general, undirected graphs on $n$ vertices, Alstrup {\em et al.}~\cite{AlstrupKTZ14} give an induced universal graph with $O(2^{n/2})$ vertices, which matches a lower bound by Moon \cite{moon1965minimal}. More recently Alon~\cite{Alonconstant2016} shows the existents of a construction having a better constant factor than the one in~\cite{AlstrupKTZ14}. It follows from~\cite{adjacencytrees2015,alstruprauhe} that $g_v(\mathcal{A}_k)=\theta(n^k)$ for the family $\mathcal{A}_k$ of graphs with constant arboricity $k$ and $n$ vertices. Using universal graphs constructed by Babai {\em et al.}~\cite{BCEGS82}, Bhatt {\em et al.}~\cite{BCLR89}, and Chung {\em et al.}~\cite{CG78,CG79,CG83,CGP76}, Chung~\cite{Chung90} obtains the best currently known bounds for e.g.~induced universal graphs for planar and outerplanar bounded degree graphs. Labeling schemes are being widely used and well-studied in the theory community: Chung~\cite{Chung90} gives labels of size $\log n+O(\log \log n)$ for adjacency labeling in trees, which was improved to $\log n + O(\log^* n)$~\cite{alstruprauhe} and in~\cite{bonichon2006short,Chung90,Fraigniaud2009randomized,fraigniaudkorman2,KMS02} to $\log n + \Theta(1)$ for various special cases of trees. Finally it was improved to $\log n + \Theta(1)$ for general trees~\cite{adjacencytrees2015}. Using labeling schemes, it is possible to avoid costly access to large global tables and instead only perform local and distributed computations. Such properties are used in applications such as XML search engines~\cite{AKM01}, network routing and distributed algorithms~\cite{Cowen01,EilamGP03,Gavoille01,ThZw05}, dynamic and parallel settings ~\cite{CohenKaplan2010,dynamicKormanP07}, and various other applications~\cite{Korman2010,peleg2,SK85}. A survey on induced universal graphs and adjacency labeling can be found in~\cite{AlstrupKTZ14}. See~\cite{gavoillepeleg} for a survey on labeling schemes for various queries. \section{Maximum degree $2$ upper bounds}\label{sec:max2upper} In this section we prove upper bounds on $g_v(\mathcal{G}_2)$, and on $g_v(\mathcal{F})$ for several special families $\mathcal{F}\subseteq\mathcal{G}_2$. \subsection{$2n-1$ upper bound for $g_v(\mathcal{G}_2)$} \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} \newenvironment{tikzpicture-Un}[1][1]{% \begin{tikzpicture}[scale=0.5] \pgfmathtruncatemacro{\n}{#1}; \pgfmathtruncatemacro{\vmax}{2*\n-2}; \begin{scope}[ vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \foreach \i in {0,...,\vmax}{% \pgfmathtruncatemacro{\x}{div(\i+1,2)}; \pgfmathtruncatemacro{\y}{mod(\i+1,2)*(2*mod(\x,2)-1)}; \node[vertex style] (v\i) at (\x,\y) {\tiny $\i$}; }; }{% \foreach \i in {0,...,\vmax}{% \pgfmathtruncatemacro{\nexti}{\i+1}; \ifthenelse{\i=2 \OR \nexti>\vmax \OR \nexti=2}{% }{% \draw[thick,color=red] (v\i) -- (v\nexti); } \pgfmathtruncatemacro{\nexti}{4-mod(\i,2)+\i} \ifthenelse{\i=2 \OR \nexti>\vmax \OR \nexti=2}{% }{% \pgfmathtruncatemacro{\imod}{mod(\i,2)} \ifthenelse{\imod=1}{% \draw[thick,color=ForestGreen] (v\i) -- (v\nexti); }{% \draw[thick,color=blue] (v\i) -- (v\nexti); } } }; \end{scope} \end{tikzpicture} }% \NewEnviron{tikzpicture-Un-induced}[2]{ \begin{tikzpicture-Un}[#1] \BODY \begin{pgfonlayer}{background} \foreach \i in {#2}{% \node[selected vertex style] at (v\i) {\tiny $\i$}; } \foreach \i in {#2}{% \ifthenelse{\i=2}{% }{% \foreach \j in {#2}{% \ifthenelse{\i<\j \AND \NOT \j=2}{% \pgfmathtruncatemacro{\nexti}{\i+1}; \ifthenelse{\nexti=\j}{% \draw[selected edge style] (v\i.center) -- (v\j.center); }{} \pgfmathtruncatemacro{\nexti}{4-mod(\i,2)+\i} \ifthenelse{\nexti=\j}{% \draw[selected edge style] (v\i.center) -- (v\j.center); }{} }{ } } } } \end{pgfonlayer} \end{tikzpicture-Un} }% \newcommand{\tikzUnInduced}[2][]{ \begin{tikzpicture-Un-induced}{#2}{#1} \end{tikzpicture-Un-induced} } Here we prove that there exists an induced universal graph with $2n-1$ vertices and $4n-9$ edges for the family $\mathcal{G}_2$ of all graphs with $n$ vertices and maximum degree $2$. \begin{definition}\label[definition]{def:Un} Let \begin{align*} s(x)&:= \begin{cases} x+4&\text{ if }x\equiv0\pmod{2} \\ x+3&\text{ otherwise} \end{cases} \end{align*} and for any $n\in\mathbb{N}_0$ let $U_n$ be the graph with vertex set $[2n-1]$ and an edge $(u,v)$ iff $u,v\neq2$ and either $\abs{u-v}=1$ or $u=s(v)$ or $v=s(u)$. (See~\cref{fig:Un-example}). \end{definition} \begin{figure}[h!] \begin{center} \tikzUnInduced{1}~ \tikzUnInduced{2}~ \tikzUnInduced{3}~ \tikzUnInduced{4}~ \tikzUnInduced{5}~ \tikzUnInduced{6} \\ ~\\ \tikzUnInduced{26} \caption{$U_1,\ldots,U_6$, and $U_{26}$ with each $(u,v)$ colored \textcolor{red}{red}/\textcolor{ForestGreen}{green}/\textcolor{blue}{blue} if $\abs{u-v}=1/3/4$.} \label{fig:Un-example} \end{center} \end{figure} \begin{theorem} The graph family $U_0, U_1, \ldots $ has the property that for $n\in\mathbb{N}_0$ \begin{enumerate}[label=(\alph*)] \item $U_n$ has $\max\set{0,2n-1}$ vertices, and $\max\set{0,n-1,3n-5,4n-9}$ edges. \item $U_n$ is an induced subgraph of $U_{n+1}$. \item $U_n$ is an induced universal graph for the family of graphs with $n$ vertices and maximum degree $2$ \end{enumerate} \end{theorem} \begin{proof} If $n=0$, $U_n$ has $0$ vertices. Otherwise the number of vertices in $U_n$ is trivially $2n-1$ from the definition. It is also clear that $U_0$ and $U_1$ each have $0$ edges, $U_2$ has $1$ edge, $U_3$ has $4$ edges, and $U_4$ has $7$ edges. Finally for $n>4$, $U_{n}$ has exactly $4$ edges more than $U_{n-1}$, and therefore has $4(n-4)+7 = 4n-9$ edges, as desired. Since the existence of an edge $(u,v)$ does not depend on $n$, the subgraph of $U_{n+1}$ induced by all vertices with label $\leq2n-2$ is exactly $U_n$. For the final part, consider a graph $G\in\mathcal{G}_2$. We need to show that $G$ is an induced subgraph of $U_n$. The proof is by induction on the number of $P_{\set{\geq3}}$ and $C_{\set{\geq4}}$ components in $G$. If there are no such components, all components are either $P_1$, $P_2$, or $C_3$. Suppose therefore that $G \simeq k_1\times{}P_1 + k_2\times{}P_2 + k_3\times{}C_3$ for some $k_1,k_2,k_3\in\mathbb{N}_0$, and let $n_1=k_1$, $n_2=2k_2$, and $n_3=3k_3$ (so $n=n_1+n_2+n_3$). Further, assume that $n>1$ since otherwise it is trivial. Informally, we will show that assigning labels greedily, smallest label first, in the order $C_3$, $P_2$, $P_1$ is sufficient. Formally, let $\set{I_3,I_2,I_1}$ be the partition of $\set{-1,\ldots,2n-2}$ into parts of size $2n_3$, $2n_2$, and $2n_1$ such that $i_3<i_2<i_1$ for all $(i_3,i_2,i_1)\in I_3\times{}I_2\times{}I_1$, and let $A_3:=I_3\setminus\set{-1,2}$, $A_2:=I_2\setminus\set{-1,2}$, and $A_1:=(I_1\cup\set{2})\setminus\set{-1}$. Then $\set{A_3,A_2,A_1}$ is a partition of $V[U_n]$. Now let \begin{itemize} \item $V_3:=\set{i\in A_3\mathrel{}\middle\vert\mathrel{} i\in\set{0,1,4}\pmod{6}}$ \item $V_2:=\set{i\in A_2\mathrel{}\middle\vert\mathrel{} i-(6n_3-1)\in\set{1,2,4,7}\pmod{8}}$ \item $V_1:= \begin{cases} \emptyset&\text{if }n_1=0\\ \set{2}&\text{if }n_1=1\\ \set{2,2(n-n_1)}\cup\set{i\in A_1\mathrel{}\middle\vert\mathrel{} i\equiv1\pmod{2}\wedge i\geq2(n-n_1)+3}&\text{otherwise} \end{cases}$ \end{itemize} Let $V=V_1\cup{}V_2\cup{}V_3$ and let $G'$ be the subgraph of $U_n$ induced by $V$. We claim that $G\simeq{}G'$ (see~\cref{fig:V1V2V3}). \begin{figure}[h!] \begin{center} \begin{tikzpicture-Un-induced}{24}{ 0,1,4, 6,7,10, 12,13,16, 18,19,22, 24,25, 27,30, 32,33, 35,38, 2,40,43,45} \draw[very thick,dotted,red] ($(v0.north west)+(-0.15,1.15)$) -- ($(v3.north west)+(-0.15,0.15)$) -- ($(v3.north west)+(-0.15,1.15)$) -- ($(v22.north east)+(+0.15,0.15)$) -- ($(v21.south east)+(+0.15,-1.15)$) -- ($(v0.south west)+(-0.15,-0.15)$) -- cycle; \draw[very thick,dotted,red] ($(v24.south west)+(-0.15,-0.15)$) rectangle ($(v38.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v2.north west)+(-0.15,0.75)$) -- ($(v46.north east)+(0.15,0.75)$) -- ($(v45.south east)+(0.15,-1.15)$) -- ($(v40.south west)+(-0.15,-0.15)$) -- ($(v39.north west)+(-0.15,1.5)$) -- ($(v2.north east)+(0.15,0.5)$) -- ($(v2.south east)+(0.15,-0.15)$) -- ($(v2.south west)+(-0.15,-0.15)$) -- cycle; \end{tikzpicture-Un-induced} \caption{$U_{24}$ with $4\times{}P_1+4\times{}P_2+4\times{}C_3$ embedded. The dotted red boxes represent $A_1$, $A_2$, and $A_3$ respectively. $V_1$, $V_2$, and $V_3$ consist of the marked vertices in each of the dotted red boxes.} \label{fig:V1V2V3} \end{center} \end{figure} \noindent Now it follows from~\cref{def:Un} that for $v\in V$ the neighbors of $v$ in $G'$ are: \begin{align*} N(v) &:= (V\setminus\set{2})\cap \begin{cases} \emptyset &\text{ if }v=2\\ \set{v-4,v-3,v-1,v+1,v+4} &\text{ if }v\neq2 \wedge v\equiv0\pmod{2}\\ \set{v-1,v+1,v+3} &\text{ otherwise} \end{cases} \end{align*} To see that $G\simeq G'$, first note that $\abs{V_1}=n_1$ and $N(v_1)=\emptyset$ for all $v_1\in{}V_1$. Thus the component of $v_1$ in $G'$ is a $P_1$. Second, note that $\abs{V_2}=n_2$ and that each vertex $v_2\in{}V_2$ has the form $v_2=b_2+k$ with $b_2=(6n_3-1)+8j\equiv1\pmod{2}$ for some $j\in[\floor{\frac{n_2}{4}}], k\in\set{1,2,4,7}$. Now $N(b_2+1)=\set{b_2+2}\subseteq V_2$, $N(b_2+2)=\set{b_2+1}\subseteq V_2$, $N(b_2+4)=\set{b_2+7}\subseteq V_2$, and $N(b_2+7)=\set{b_2+4}\subseteq V_2$, so $v_2$ has exactly one neighbor in $V$, and this neighbor is also in $V_2$. Thus the component of $v_2$ in $G'$ is a $P_2$. Third, note that $\abs{V_3}=n_3$, and that each vertex $v_3\in{}V_3$ has the form $v_3=b_3+k$ with $b_3=6j\equiv0\pmod{2}$ for some $j\in[\frac{n_3}{3}], k\in\set{0,1,4}$. Now $N(b_3+0)=\set{b_3+1,b_3+4}\subseteq V_3$, $N(b_3+1)=\set{b_3+0,b+4}\subseteq V_3$, and $N(b_3+4)=\set{b_3+0,b+1}\subseteq V_3$, so the component of $v_3$ in $G'$ consists of the $3$ vertices $\set{b_3,b_3+1,b_3+4}$, which form a $C_3$. Thus $G'\simeq n_1\times{}P_1 + \frac{n_2}{2}\times{}P_2 + \frac{n_3}{3}\times{}C_3 \simeq k_1\times{}P_1 + k_2\times{}P_2 + k_3\times{}C_3 \simeq G$. For the induction case, suppose $G$ has a component $X$ which is either a $P_k$ for some $k\geq3$, or a $C_k$ for some $k\geq4$. In either case, let $I^-:=[2(n-k)-1]$ and $I^+:=[2n-1]\setminus{}I^-$. Then $I^-$ induces an $U_{n-k}$ subgraph in $U_n$, which by induction has $G-X$ as induced subgraph. Thus all we need to show is that we can extend this to an embedding of $G$ by using using only vertices from $I^+$ to embed $X$ (see~\cref{fig:VP-VC}). \begin{figure}[h!]% \begin{center}% \begin{tikzpicture-Un-induced}{6}{1,4,6,8,9,10} \draw[very thick,dotted,red] ($(v0.south west)+(-1.15,-0.15)$) rectangle ($(v0.north east)+(-0.85,2.15)$); \draw[very thick,dotted,green] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v10.north east)+(0.15,0.15)$); \end{tikzpicture-Un-induced} ~~~ \begin{tikzpicture-Un-induced}{7}{3,6,8,10,11,12} \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v0.north east)+(0.15,2.15)$); \draw[very thick,dotted,green] ($(v2.north west)+(-0.15,0.15)$) rectangle ($(v12.south east)+(0.15,-0.15)$); \end{tikzpicture-Un-induced} ~~~ \begin{tikzpicture-Un-induced}{10}{9,12,14,16,17,18} \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v6.north east)+(0.15,0.15)$); \draw[very thick,dotted,green] ($(v8.south west)+(-0.15,-0.15)$) rectangle ($(v18.north east)+(0.15,0.15)$); \end{tikzpicture-Un-induced}% \\ ~ \\ \begin{tikzpicture-Un-induced}{6}{3,4,6,8,9,10} \draw[very thick,dotted,red] ($(v0.south west)+(-1.15,-0.15)$) rectangle ($(v0.north east)+(-0.85,2.15)$); \draw[very thick,dotted,green] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v10.north east)+(0.15,0.15)$); \end{tikzpicture-Un-induced} ~~~ \begin{tikzpicture-Un-induced}{7}{5,6,8,10,11,12} \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v0.north east)+(0.15,2.15)$); \draw[very thick,dotted,green] ($(v2.north west)+(-0.15,0.15)$) rectangle ($(v12.south east)+(0.15,-0.15)$); \end{tikzpicture-Un-induced} ~~~ \begin{tikzpicture-Un-induced}{10}{11,12,14,16,17,18} \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v6.north east)+(0.15,0.15)$); \draw[very thick,dotted,green] ($(v8.south west)+(-0.15,-0.15)$) rectangle ($(v18.north east)+(0.15,0.15)$); \end{tikzpicture-Un-induced}% \\ \caption{$P_6$ (top) and $C_6$ (bottom) embedded in $U_{6}$ (left), $U_{7}$ (middle), and $U_{10}$ (right). In each case the dotted red and green boxes represent $I^-$ and $I^+$ respectively, and $V^P$ or $V^C$ is the marked vertices in the green box. Notice that the contents of the red boxes is $U_0$ (left), $U_1$ (middle), and $U_4$ (right).} \label{fig:VP-VC} \end{center} \end{figure} \noindent Now let \begin{itemize} \item $V^P:=\set{2(n-k)+1, 2n-3}\cup\set{i\in I^+\mathrel{}\middle\vert\mathrel{} i\geq 2(n-k)+4 \wedge i\equiv0\pmod{2}}$ \item $V^C:=\set{2(n-k)+3, 2n-3}\cup\set{i\in I^+\mathrel{}\middle\vert\mathrel{} i\geq 2(n-k)+4 \wedge i\equiv0\pmod{2}}$ \end{itemize} Now for $k\geq3$ the subgraph induced by $V^P$ in $U_n$ is $P_k$, and for all $v\in{}V^P$ all neighbors to $v$ are in $I^+$. Similarly, for $k\geq4$ the subgraph induced by $V^P$ in $U_n$ is $C_k$, and for all $v\in{}V^C$ all neighbors to $v$ are in $I^+$. Thus, either $V^P$ or $V^C$ can be used to extend the embedding of $G-X$ in $U_{n-k}$ to an embedding of $G$ in $U_n$. \end{proof} \subsection{$\frac{11}{6}n+\Oo(1)$ upper bound when all components are small} Esperet {\em et al.}~\cite{Esperet2008} showed that $g_v(\mathcal{G}_2)\geq 11 \floor{n/6}$ by considering a specific family of graphs whose largest component had $3$ vertices. We used the same idea in our proof of \Cref{thm:pathlower}. A natural attempt to improve the lower bound would be to include larger components. However, as the following shows, considering components with $4$, $5$, or $6$ vertices is not sufficient. \begin{figure}[h!] \begin{tikzpicture*}{\textwidth} \begin{scope}[every node/.append style={ draw, circle, minimum size=2mm, inner sep=0pt, outer sep=0pt% }] \node (v0) at (1,1) {}; \node (v1) at (2,1) {}; \node (v2) at (3,1) {}; \node (v3) at (1,2) {}; \node (v4) at (2,2) {}; \node (v5) at (3,2) {}; \node (v6) at (1,3) {}; \node (v7) at (2,3) {}; \node (v8) at (3,3) {}; \node (v9) at (1,4) {}; \node (v10) at (3,4) {}; \node (v11) at (1,5) {}; \node (v12) at (2,5) {}; \node (v13) at (3,5) {}; \node (v14) at (1,6) {}; \node (v15) at (2,6) {}; \node (v16) at (3,6) {}; \node (v17) at (1,7) {}; \node (v18) at (2,7) {}; \node (v19) at (3,7) {}; \draw (v0) -- (v1); \draw (v1) -- (v2); \draw (v0) -- (v3); \draw (v1) -- (v3); \draw (v2) -- (v4); \draw (v2) -- (v5); \draw (v3) -- (v4); \draw (v3) -- (v6); \draw (v4) -- (v6); \draw (v5) -- (v7); \draw (v5) -- (v8); \draw (v6) -- (v7); \draw (v7) -- (v8); \draw (v7) -- (v9); \draw (v7) -- (v10); \draw (v9) -- (v12); \draw (v10) -- (v12); \draw (v11) -- (v12); \draw (v11) -- (v14); \draw (v12) -- (v13); \draw (v12) -- (v14); \draw (v13) -- (v15); \draw (v13) -- (v16); \draw (v14) -- (v17); \draw (v15) -- (v16); \draw (v15) -- (v17); \draw (v16) -- (v18); \draw (v16) -- (v19); \draw (v17) -- (v18); \draw (v18) -- (v19); \node (v20) at (1+4,1) {}; \node (v21) at (2+4,1) {}; \node (v22) at (3+4,1) {}; \node (v23) at (1+4,2) {}; \node (v24) at (2+4,2) {}; \node (v25) at (3+4,2) {}; \node (v26) at (1+4,3) {}; \node (v27) at (2+4,3) {}; \node (v28) at (3+4,3) {}; \node (v29) at (1+4,4) {}; \node (v30) at (3+4,4) {}; \node (v31) at (1+4,5) {}; \node (v32) at (2+4,5) {}; \node (v33) at (3+4,5) {}; \node (v34) at (1+4,6) {}; \node (v35) at (2+4,6) {}; \node (v36) at (3+4,6) {}; \node (v37) at (1+4,7) {}; \node (v38) at (2+4,7) {}; \node (v39) at (3+4,7) {}; \draw (v20) -- (v21); \draw (v21) -- (v22); \draw (v20) -- (v23); \draw (v21) -- (v23); \draw (v22) -- (v24); \draw (v22) -- (v25); \draw (v23) -- (v24); \draw (v23) -- (v26); \draw (v24) -- (v26); \draw (v25) -- (v27); \draw (v25) -- (v28); \draw (v26) -- (v27); \draw (v27) -- (v28); \draw (v27) -- (v29); \draw (v27) -- (v30); \draw (v29) -- (v32); \draw (v30) -- (v32); \draw (v31) -- (v32); \draw (v31) -- (v34); \draw (v32) -- (v33); \draw (v32) -- (v34); \draw (v33) -- (v35); \draw (v33) -- (v36); \draw (v34) -- (v37); \draw (v35) -- (v36); \draw (v35) -- (v37); \draw (v36) -- (v38); \draw (v36) -- (v39); \draw (v37) -- (v38); \draw (v38) -- (v39); \node (v40) at (1+14,1) {}; \node (v41) at (2+14,1) {}; \node (v42) at (3+14,1) {}; \node (v43) at (1+14,2) {}; \node (v44) at (2+14,2) {}; \node (v45) at (3+14,2) {}; \node (v46) at (1+14,3) {}; \node (v47) at (2+14,3) {}; \node (v48) at (3+14,3) {}; \node (v49) at (1+14,4) {}; \node (v50) at (3+14,4) {}; \node (v51) at (1+14,5) {}; \node (v52) at (2+14,5) {}; \node (v53) at (3+14,5) {}; \node (v54) at (1+14,6) {}; \node (v55) at (2+14,6) {}; \node (v56) at (3+14,6) {}; \node (v57) at (1+14,7) {}; \node (v58) at (2+14,7) {}; \node (v59) at (3+14,7) {}; \draw (v40) -- (v41); \draw (v41) -- (v42); \draw (v40) -- (v43); \draw (v41) -- (v43); \draw (v42) -- (v44); \draw (v42) -- (v45); \draw (v43) -- (v44); \draw (v43) -- (v46); \draw (v44) -- (v46); \draw (v45) -- (v47); \draw (v45) -- (v48); \draw (v46) -- (v47); \draw (v47) -- (v48); \draw (v47) -- (v49); \draw (v47) -- (v50); \draw (v49) -- (v52); \draw (v50) -- (v52); \draw (v51) -- (v52); \draw (v51) -- (v54); \draw (v52) -- (v53); \draw (v52) -- (v54); \draw (v53) -- (v55); \draw (v53) -- (v56); \draw (v54) -- (v57); \draw (v55) -- (v56); \draw (v55) -- (v57); \draw (v56) -- (v58); \draw (v56) -- (v59); \draw (v57) -- (v58); \draw (v58) -- (v59); \node (v60) at (11,1) {}; \node (v61) at (11,2) {}; \node (v62) at (10,3) {}; \node (v63) at (11,3) {}; \node (v64) at (12,3) {}; \node (v65) at (10,4) {}; \node (v66) at (12,4) {}; \node (v67) at ( 9,5) {}; \node (v68) at (13,5) {}; \node (v69) at (10,6) {}; \node (v70) at (12,6) {}; \node (v71) at (10,7) {}; \node (v72) at (11,7) {}; \node (v73) at (12,7) {}; \node (v74) at (11,8) {}; \node (v75) at (11,9) {}; \draw (v60) -- (v61); \draw (v61) -- (v62); \draw (v61) -- (v63); \draw (v61) -- (v64); \draw (v62) -- (v65); \draw (v63) -- (v65); \draw (v63) -- (v66); \draw (v64) -- (v66); \draw (v65) -- (v66); \draw (v65) -- (v67); \draw (v66) -- (v68); \draw (v67) -- (v69); \draw (v68) -- (v70); \draw (v69) -- (v70); \draw (v69) -- (v71); \draw (v69) -- (v72); \draw (v70) -- (v72); \draw (v70) -- (v73); \draw (v71) -- (v74); \draw (v72) -- (v74); \draw (v73) -- (v74); \draw (v74) -- (v75); \node (v76) at (11,13) {}; \node (v77) at ( 8,10) {}; \node (v78) at ( 9,10) {}; \node (v79) at ( 7,11) {}; \node (v80) at ( 8,11) {}; \node (v81) at (10,11) {}; \node (v82) at ( 8,12) {}; \node (v83) at ( 9,12) {}; \draw (v77) -- (v78); \draw (v77) -- (v79); \draw (v78) -- (v80); \draw (v78) -- (v81); \draw (v78) -- (v83); \draw (v79) -- (v80); \draw (v79) -- (v82); \draw (v81) -- (v83); \draw (v82) -- (v83); \node (v84) at (22- 8,10) {}; \node (v85) at (22- 9,10) {}; \node (v86) at (22- 7,11) {}; \node (v87) at (22- 8,11) {}; \node (v88) at (22-10,11) {}; \node (v89) at (22- 8,12) {}; \node (v90) at (22- 9,12) {}; \draw (v84) -- (v85); \draw (v84) -- (v86); \draw (v85) -- (v87); \draw (v85) -- (v88); \draw (v85) -- (v90); \draw (v86) -- (v87); \draw (v86) -- (v89); \draw (v88) -- (v90); \draw (v89) -- (v90); \draw (v75) -- (v78); \draw (v75) -- (v85); \draw (v76) -- (v83); \draw (v76) -- (v90); \node (v91) at ( 2, 9) {}; \node (v92) at ( 6, 9) {}; \node (v93) at ( 4,10) {}; \node (v94) at ( 3,11) {}; \node (v95) at ( 5,11) {}; \node (v96) at ( 4,12) {}; \node (v97) at ( 2,13) {}; \node (v98) at ( 6,13) {}; \node (v99) at ( 1,11) {}; \draw (v91) -- (v92); \draw (v91) -- (v93); \draw (v91) -- (v97); \draw (v92) -- (v93); \draw (v93) -- (v94); \draw (v93) -- (v95); \draw (v94) -- (v96); \draw (v95) -- (v96); \draw (v96) -- (v97); \draw (v96) -- (v98); \draw (v97) -- (v98); \draw (v97) -- (v99); \node (v100) at (22- 2, 9) {}; \node (v101) at (22- 6, 9) {}; \node (v102) at (22- 4,10) {}; \node (v103) at (22- 3,11) {}; \node (v104) at (22- 5,11) {}; \node (v105) at (22- 4,12) {}; \node (v106) at (22- 2,13) {}; \node (v107) at (22- 6,13) {}; \node (v108) at (22- 1,11) {}; \draw (v100) -- (v101); \draw (v100) -- (v102); \draw (v100) -- (v106); \draw (v101) -- (v102); \draw (v102) -- (v103); \draw (v102) -- (v104); \draw (v103) -- (v105); \draw (v104) -- (v105); \draw (v105) -- (v106); \draw (v105) -- (v107); \draw (v106) -- (v107); \draw (v106) -- (v108); \draw (v79) -- (v92); \draw (v79) -- (v98); \draw (v86) -- (v101); \draw (v86) -- (v107); \node (v109) at (19,4) {}; \end{scope} \end{tikzpicture*} \caption{This graph on $110$ nodes embeds all graphs on $60$ nodes whose maximum degree is $2$ and whose largest component has size at most $6$.} \label{fig:116-small} \end{figure} \begin{theorem} For any $n\in\mathbb{N}$ there exists a graph with $\frac{11}{6}n+\mathcal{O}(1)$ vertices, that contains as induced subgraphs all graphs on $n$ vertices whose maximum degree is $2$ and whose largest component has at most $6$ vertices. \end{theorem} \begin{proof} Let $G$ be a graph with $n$ vertices and maximum degree $2$ whose largest component has at most $6$ vertices. Let $H$ be the graph on $110$ vertices depicted in Figure~\ref{fig:116-small}. It is easy to see by inspection that this graph has each of $60\times P_1$, $30\times P_2$, $20\times P_3$, $20\times C_3$, $15\times P_4$, $15\times C_4$, $12\times P_5$, $12\times C_5$, $10\times P_6$, and $10\times C_6$ as induced subgraphs. Now, any graph with the desired properties whose components do not include all the components of one of these graphs has at most $(60-1)+(60-2)+(60-3)+(60-3)+(60-4)+(60-4)+(60-5)+(60-5)+(60-6)+(60-6)=561$ vertices, and can be embedded in at most $10$ copies of $H$. The whole of $G$ can therefore be embedded in at most $c=\left\lfloor\frac{n}{60}\right\rfloor+10$ copies of $H$. The total number of vertices in $c\times H$ is $110c\leq\frac{11}{6}n+1100$, which is $\frac{11}{6}n+\mathcal{O}(1)$ as desired. \end{proof} A more careful analysis shows that $\left\lceil\frac{n}{60}\right\rceil$ copies of $H$ is always sufficient to embed any $G\in\mathcal{G}_2$ with $n$ vertices, which in particular means that the $11\floor{n/6}$ bound is achievable for any such graph with $n$ divisible by $60$. We are not sure if the above construction can be extended to handle components of size $7$ or more. It would involve constructing a graph with $770$ vertices. An interesting open question is: Is there a function $f$, such that for any $n,s\in\mathbb{N}$ the family of graphs with $n$ vertices, maximum degree $2$, and maximum component size $s$ has an induced universal graph with at most $\frac{11}{6}n+f(s)$ vertices? \subsection{$\frac{11}{6}n+\Oo(1)$ upper bound when all components are large} Since it appears difficult to improve the lower bound by considering only small components, the next natural thing might be to consider just large components. As the following upper bound shows, this is also unlikely to succeed. \begin{figure}[h] \begin{tikzpicture*}{\textwidth} \pgfmathtruncatemacro{\cols}{13}; \pgfmathtruncatemacro{\p}{1}; \begin{scope}[every node/.append style={ draw, circle, minimum size=2mm, inner sep=0pt, outer sep=0pt% }, vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[vertex style] (v01) at (0,2) {}; \foreach \i in {1,...,\cols}{% \node[vertex style] (v\i0) at (-1+2*\i,0) {}; \node[vertex style] (v\i1) at ( 0+2*\i,2) {}; \node[vertex style] (v\i2) at (-1+2*\i,4) {}; }; \foreach \i in {1,...,\cols}{% \pgfmathtruncatemacro{\previ}{\i-1}; \draw[thick,color=blue] (v\i0) -- (v\previ1) -- (v\i2); \ifthenelse{\i>1}{% \draw[thick,color=blue] (v\previ0) -- (v\i0); \draw[thick,color=blue] (v\previ2) -- (v\i2); }{% } \draw[thick,color=blue] (v\i0) -- (v\i1) -- (v\i2); \pgfmathtruncatemacro{\imod}{mod(\i,\p)}; \ifthenelse{\imod=0}{% \draw[dashed] ($(v\i0.south east)+(0.25,-1.25)$) -- ($(v\i2.north east)+(0.25,1.25)$); }{% } }; \end{scope} \end{tikzpicture*} \caption{The graph with this repeated pattern embeds in the first $\frac{3}{2}n-2$ vertices all graphs on $n$ vertices whose components are all even cycles.} \label{fig:32-even} \end{figure} \begin{figure}[h] \begin{tikzpicture*}{\textwidth} \pgfmathtruncatemacro{\cols}{13}; \pgfmathtruncatemacro{\p}{3}; \begin{scope}[every node/.append style={ draw, circle, minimum size=2mm, inner sep=0pt, outer sep=0pt% }, vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[vertex style] (v01) at (0,2) {}; \foreach \i in {1,...,\cols}{% \node[vertex style] (v\i0) at (-1+2*\i,0) {}; \node[vertex style] (v\i1) at ( 0+2*\i,2) {}; \node[vertex style] (v\i2) at (-1+2*\i,4) {}; \pgfmathtruncatemacro{\imod}{mod(\i,\p)}; \pgfmathtruncatemacro{\itop}{\i+1}; \ifthenelse{\imod=1 \AND \itop<\cols}{% \node[vertex style,color=red] (v\i3) at (2+2*\i,3.25) {}; }{% } }; \foreach \i in {1,...,\cols}{% \pgfmathtruncatemacro{\previ}{\i-1}; \draw[thick,color=blue] (v\i0) -- (v\previ1) -- (v\i2); \ifthenelse{\i>1}{% \draw[thick,color=blue] (v\previ0) -- (v\i0); \draw[thick,color=blue] (v\previ2) -- (v\i2); }{% } \draw[thick,color=blue] (v\i0) -- (v\i1) -- (v\i2); \pgfmathtruncatemacro{\imod}{mod(\i,\p)}; \pgfmathtruncatemacro{\itop}{\i+2}; \ifthenelse{\imod=1 \AND \itop<\cols}{% \draw[thick,color=red] (v\i2) -- (v\i3); \draw[thick,color=red] (v\i3) -- (v\itop1); \pgfmathtruncatemacro{\inext}{\i+3}; \draw[thick,color=red] (v\i3) -- (v\inext2); \draw[thick,color=red] (v\i3) -- (v\i1); }{% } \ifthenelse{\imod=0}{% \draw[dashed] ($(v\i0.south east)+(0.25,-1.25)$) -- ($(v\i2.north east)+(0.25,1.25)$); }{% } }; \end{scope} \end{tikzpicture*} \caption{The graph with this repeated pattern embeds in the first $\frac{11}{6}n+\mathcal{O}(1)$ vertices all graphs on $n$ vertices whose maximum degree is $2$ and whose smallest component has size at least $10$.} \label{fig:116-large} \end{figure} \begin{theorem} For any $n\in\mathbb{N}$ there exists a graph with at most $\frac{11}{6}n+\mathcal{O}(1)$ vertices, that contain as induced subgraphs all graphs on $n$ vertices whose maximum degree is $2$ and whose smallest component has at least $10$ vertices. \end{theorem} \begin{proof} Consider the graph family $A$ implied by the pattern in Figure~\ref{fig:32-even}. Observe that the leftmost $\frac{3}{2}n-2$ vertices of this pattern embeds all graphs on $n$ vertices whose components are all even cycles. In particular, each (even) cycle of length $s\geq10$ will be embedded in such a way that it uses exactly $\frac{s-4}{2}\geq3$ consecutive edges from the top or bottom rows. Now consider the modified graph family $B$ in Figure~\ref{fig:116-large}. Every third edge along the top row is associated with one of the added red vertices. Now consider an odd cycle $C$ of length $s>10$. By splitting one vertex in the cycle, we can extend it to an even cycle $C'$ of length $\geq12$. When we embed this $C'$, it will contain an edge associated with one of the red vertices. We can therefore arrange that the split vertices are the endpoints of such an edge, and we can obtain an embedding of $C$ by replacing them with the associated red vertex. Similarly, since the minimum component size is assumed to be $10$, we can turn any number of paths of total length $s$ into a cycle of even length by adding at most $\frac{1}{10}n+\mathcal{O}(1)$ vertices. Thus any graph $G$ with $n$ vertices, maximum degree $2$ and minimum component size $10$ can be converted into a graph $G'$ with at most $n'=\frac{11}{10}n+\mathcal{O}(1)$ vertices whose components are all even cycles. The graph $G'$ can then be embedded in a graph from family $A$ with $\frac{3}{2}n'+\mathcal{O}(1)$ vertices. This graph can then be converted to a graph in family $B$ with at most $\frac{10}{9}(\frac{3}{2}n'+\mathcal{O}(1))=\frac{11}{6}n+\mathcal{O}(1)$ vertices, and the embedding of $G'$ in $A$ can be converted to an embedding of $G$ in $B$ as previsously described. \end{proof} \section{Paths}\label{sec:paths} \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} We show that there exists an induced universal graph with $\lfloor 3n/2 \rfloor$ vertices and at most $\lfloor 3n/2 \rfloor - 1$ edges for the family $\mathcal{AC}$ of graphs consisting of a set of paths with a total of $n$ vertices. Our new induced universal graph matches the following lower bound. \begin{theorem}[Claim 1, \cite{Esperet2008}]\label{thm:pathlower} Every induced universal graph that embeds any acyclic graph $G$ on $n$ vertices with $\Delta(G)\leq 2$ has at least $\lfloor 3n/2 \rfloor$ vertices. \end{theorem} \begin{proof} The theorem can be extracted from the proof of Claim $1$ in \cite{Esperet2008}. Consider the two graphs $G'$ and $G''$, with $G'$ consisting of $n$ $P_1$ and $G''$ consising of $\floor{n/2}$ $P_2$ plus at most one $P_1$. An induced universal subgraph must contain $n$ disjoint $P_1$ to embed $G'$, and $\floor{n/2}$ disjoint $P_2$ to embed $G''$. Each $P_2$ in the embedding of $G''$ can overlap with at most one $P_1$ from the embedding of $G'$ in the induced universal graph, hence we need at least $n + \lfloor n/2 \rfloor = \lfloor 3n/2 \rfloor$ vertices in the induced universal graph. \end{proof} Our induced universal graph $U^p_n$ that matches \Cref{thm:pathlower} is defined as below. \begin{definition}\label{def:upn} Let \begin{align*} s(x)&:= \begin{cases} x+2&\text{ if }x\equiv2\pmod{3} \\ x+1&\text{ otherwise} \end{cases} \end{align*} For any $n \in {\mathbb{N}}$ let $U^p_n$ be the graph over vertex set $\left[ \lfloor 3n/2 \rfloor \right]$, and for all $u < v \in [\lfloor 3n/2 \rfloor]$ let there be an edge $(u,v)$ iff $v = s(u)$ (See \Cref{fig:Up_n}). \end{definition} \begin{figure}[h] \centering \begin{tikzpicture}[scale=0.5] \pgfmathtruncatemacro{\n}{6}; \pgfmathtruncatemacro{\vmax}{2*\n}; \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} \begin{scope}[ vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[vertex style] (v0) at (0,-1) {\tiny $0$}; \node[vertex style] (v1) at (0,0) {\tiny $1$}; \node[vertex style] (v2) at (1,0) {\tiny $2$}; \node[vertex style] (v3) at (2,-1) {\tiny $3$}; \node[vertex style] (v4) at (2,0) {\tiny $4$}; \node[vertex style] (v5) at (3,0) {\tiny $5$}; \node[vertex style] (v6) at (4,-1) {\tiny $6$}; \node[vertex style] (v7) at (4,0) {\tiny $7$}; \node[vertex style] (v8) at (5,0) {\tiny $8$}; \node[vertex style] (v9) at (6,-1) {\tiny $9$}; \node[vertex style] (v10) at (6,0) {\tiny $10$}; \node[vertex style] (v11) at (7,0) {\tiny $11$}; \node[vertex style] (v12) at (8,-1) {\tiny $12$}; \node[vertex style] (v13) at (8,0) {\tiny $13$}; \node[vertex style] (v14) at (9,0) {\tiny $14$}; \node[vertex style] (v15) at (10,-1) {\tiny $15$}; \draw[thick,color=blue] (v0) -- (v1); \draw[thick,color=blue] (v3) -- (v4); \draw[thick,color=blue] (v6) -- (v7); \draw[thick,color=blue] (v9) -- (v10); \draw[thick,color=blue] (v12) -- (v13); \draw[thick,color=blue] (v1) -- (v2); \draw[thick,color=red] (v2) -- (v4); \draw[thick,color=blue] (v4) -- (v5); \draw[thick,color=red] (v5) -- (v7); \draw[thick,color=blue] (v7) -- (v8); \draw[thick,color=red] (v8) -- (v10); \draw[thick,color=blue] (v10) -- (v11); \draw[thick,color=red] (v11) -- (v13); \draw[thick,color=blue] (v13) -- (v14); \end{scope} \end{tikzpicture} \caption{$U^p_{11}$, with each $(u,v)$ colored \textcolor{blue}{blue} if $v=u+1$ and \textcolor{red}{red} if $v=u+2$, i.e., the edges come from cases $1$ and $2$ of \Cref{def:upn}, respectively.} \label[figure]{fig:Up_n} \end{figure} When $n$ is even $U^p_n $ is a simple caterpillar graph, where every second vertex of the main path has an extra vertex connected to it. When $n$ is odd $U^p_n$ is a caterpillar and a single isolated vertex (see \Cref{fig:Up_n}). In the following, we show that $U^p_n$ embeds the family of acyclic graphs over $n$ vertices with maximum degree $2$, and that the corresponding decoder is size oblivious. We proceed by showing that a few particular paths can be embedded in the graph, and finally that any graph in $\mathcal{AC}$ can be embedded using the same embedding techniques. \begin{definition}\label{def:bb} Let $G$ and $U$ be graphs and let $\lambda$ be an embedding function of $G$ into $U$. We say that a vertex $v \in V[U]$ is \emph{used} if there exists $v' \in V[G]$ such that $\lambda(v') = v$. A node $w\in V[U]$ is \emph{allocated} if $w$ is used or adjacent to a used vertex. A subset $X \subseteq V[U]$ is \emph{allocated} when every vertex $x \in X$ is allocated. \end{definition} This definition allows us to argue that, given a specific graph $G\in\mathcal{AC}$ with $n$ vertices, we can allocate a part of our induced universal graph $U^p_{n}$ for embedding a part of $G$, and then the remaining unallocated part of $U^p_{n}$ is available for embedding for the remaining part of $G$. We divide $G$ into several parts and show that each part can be embedded in $U^p_{n}$. For each part we embed, the number of allocated vertices divded by the number of used vertices is at most $\lfloor 3n/2 \rfloor$. \Cref{thm:paths} implies that these embedding strategies can be combined, i.e., applied consecutively one after the other, implying that all of $G$ can be embedded in $U^p_{n}$. Our argument relies heavily on allocation of blocks as defined below. \begin{definition}\label{def:block} Let \emph{block} $B_j$ of $U^p_n$, where $j\in[\floor{\frac{n}{2}}]$, be the vertices $\{3j, 3j+1, 3j+2 \}$. \end{definition} In \Cref{lem:combined}, we show how to embed single paths of any length and simple families of paths in $U^p_{n}$ efficiently. \Cref{fig:p2k1,fig:p3p1} shows how the cases are handled. \begin{figure}[h] \centering \begin{tikzpicture}[scale=0.5] \pgfmathtruncatemacro{\n}{6}; \pgfmathtruncatemacro{\vmax}{2*\n}; \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} \begin{scope}[ vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[selected vertex style] (v0) at (0,-1) {\tiny $0$}; \node[selected vertex style] (v1) at (0,0) {\tiny $1$}; \node[selected vertex style] (v2) at (1,0) {\tiny $2$}; \node[vertex style] (v3) at (2,-1) {\tiny $3$}; \node[selected vertex style] (v4) at (2,0) {\tiny $4$}; \node[selected vertex style] (v5) at (3,0) {\tiny $5$}; \node[vertex style] (v6) at (4,-1) {\tiny $6$}; \node[selected vertex style] (v7) at (4,0) {\tiny $7$}; \node[selected vertex style] (v8) at (5,0) {\tiny $8$}; \node[vertex style] (v9) at (6,-1) {\tiny $9$}; \node[selected vertex style] (v10) at (6,0) {\tiny $10$}; \node[vertex style] (v11) at (7,0) {\tiny $11$}; \node[vertex style] (v12) at (8,-1) {\tiny $12$}; \node[vertex style] (v13) at (8,0) {\tiny $13$}; \node[vertex style] (v14) at (9,0) {\tiny $14$}; \draw[thick,color=blue] (v0) -- (v1); \draw[thick,color=blue] (v3) -- (v4); \draw[thick,color=blue] (v6) -- (v7); \draw[thick,color=blue] (v9) -- (v10); \draw[thick,color=blue] (v12) -- (v13); \draw[thick,color=blue] (v1) -- (v2); \draw[thick,color=red] (v2) -- (v4); \draw[thick,color=blue] (v4) -- (v5); \draw[thick,color=red] (v5) -- (v7); \draw[thick,color=blue] (v7) -- (v8); \draw[thick,color=red] (v8) -- (v10); \draw[thick,color=blue] (v10) -- (v11); \draw[thick,color=red] (v11) -- (v13); \draw[thick,color=blue] (v13) -- (v14); \begin{pgfonlayer}{background} \draw[selected edge style] (v0.center) -- (v1.center) -- (v2.center) -- (v4.center) -- (v5.center) -- (v7.center) -- (v8.center) -- (v10.center); \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v2.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v3.south west)+(-0.15,-0.15)$) rectangle ($(v5.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v6.south west)+(-0.15,-0.15)$) rectangle ($(v8.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v9.south west)+(-0.15,-0.15)$) rectangle ($(v11.north east)+(0.15,0.15)$); \end{pgfonlayer} \end{scope} \end{tikzpicture} ~~~~ \begin{tikzpicture}[scale=0.5] \pgfmathtruncatemacro{\n}{6}; \pgfmathtruncatemacro{\vmax}{2*\n}; \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} \begin{scope}[ vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[selected vertex style] (v0) at (0,-1) {\tiny $0$}; \node[selected vertex style] (v1) at (0,0) {\tiny $1$}; \node[selected vertex style] (v2) at (1,0) {\tiny $2$}; \node[vertex style] (v3) at (2,-1) {\tiny $3$}; \node[selected vertex style] (v4) at (2,0) {\tiny $4$}; \node[selected vertex style] (v5) at (3,0) {\tiny $5$}; \node[vertex style] (v6) at (4,-1) {\tiny $6$}; \node[selected vertex style] (v7) at (4,0) {\tiny $7$}; \node[selected vertex style] (v8) at (5,0) {\tiny $8$}; \node[selected vertex style] (v9) at (6,-1) {\tiny $9$}; \node[selected vertex style] (v10) at (6,0) {\tiny $10$}; \node[vertex style] (v11) at (7,0) {\tiny $11$}; \node[vertex style] (v12) at (8,-1) {\tiny $12$}; \node[vertex style] (v13) at (8,0) {\tiny $13$}; \node[vertex style] (v14) at (9,0) {\tiny $14$}; \draw[thick,color=blue] (v0) -- (v1); \draw[thick,color=blue] (v3) -- (v4); \draw[thick,color=blue] (v6) -- (v7); \draw[thick,color=blue] (v9) -- (v10); \draw[thick,color=blue] (v12) -- (v13); \draw[thick,color=blue] (v1) -- (v2); \draw[thick,color=red] (v2) -- (v4); \draw[thick,color=blue] (v4) -- (v5); \draw[thick,color=red] (v5) -- (v7); \draw[thick,color=blue] (v7) -- (v8); \draw[thick,color=red] (v8) -- (v10); \draw[thick,color=blue] (v10) -- (v11); \draw[thick,color=red] (v11) -- (v13); \draw[thick,color=blue] (v13) -- (v14); \begin{pgfonlayer}{background} \draw[selected edge style] (v0.center) -- (v1.center) -- (v2.center) -- (v4.center) -- (v5.center) -- (v7.center) -- (v8.center) -- (v10.center) -- (v9.center); \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v2.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v3.south west)+(-0.15,-0.15)$) rectangle ($(v5.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v6.south west)+(-0.15,-0.15)$) rectangle ($(v8.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v9.south west)+(-0.15,-0.15)$) rectangle ($(v11.north east)+(0.15,0.15)$); \end{pgfonlayer} \end{scope} \end{tikzpicture} \caption{(left) $U^p_{10}$, with a $P_{8}$ embedded, and (right) $U^p_{10}$, with a $P_{9}$ embedded. The embedded paths are shown embedded in \textcolor{blue!40}{blue}, and the allocated blocks in \textcolor{red}{red}. Both cases use at most $\lfloor 3n/2 \rfloor $ vertices.}\label[figure]{fig:p2k1} \end{figure} \begin{figure}[h] \centering \begin{tikzpicture}[scale=0.5] \pgfmathtruncatemacro{\n}{6}; \pgfmathtruncatemacro{\vmax}{2*\n}; \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} \begin{scope}[ vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[selected vertex style] (v0) at (0,-1) {\tiny $0$}; \node[selected vertex style] (v1) at (0,0) {\tiny $1$}; \node[selected vertex style] (v2) at (1,0) {\tiny $2$}; \node[vertex style] (v3) at (2,-1) {\tiny $3$}; \node[vertex style] (v4) at (2,0) {\tiny $4$}; \node[selected vertex style] (v5) at (3,0) {\tiny $5$}; \node[selected vertex style] (v6) at (4,-1) {\tiny $6$}; \node[selected vertex style] (v7) at (4,0) {\tiny $7$}; \node[vertex style] (v8) at (5,0) {\tiny $8$}; \node[vertex style] (v9) at (6,-1) {\tiny $9$}; \node[vertex style] (v10) at (6,0) {\tiny $10$}; \node[vertex style] (v11) at (7,0) {\tiny $11$}; \node[vertex style] (v12) at (8,-1) {\tiny $12$}; \node[vertex style] (v13) at (8,0) {\tiny $13$}; \node[vertex style] (v14) at (9,0) {\tiny $14$}; \draw[thick,color=blue] (v0) -- (v1); \draw[thick,color=blue] (v3) -- (v4); \draw[thick,color=blue] (v6) -- (v7); \draw[thick,color=blue] (v9) -- (v10); \draw[thick,color=blue] (v12) -- (v13); \draw[thick,color=blue] (v1) -- (v2); \draw[thick,color=red] (v2) -- (v4); \draw[thick,color=blue] (v4) -- (v5); \draw[thick,color=red] (v5) -- (v7); \draw[thick,color=blue] (v7) -- (v8); \draw[thick,color=red] (v8) -- (v10); \draw[thick,color=blue] (v10) -- (v11); \draw[thick,color=red] (v11) -- (v13); \draw[thick,color=blue] (v13) -- (v14); \begin{pgfonlayer}{background} \draw[selected edge style] (v0.center) -- (v1.center) -- (v2.center); \draw[selected edge style] (v5.center) -- (v7.center) -- (v6.center); \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) rectangle ($(v2.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v3.south west)+(-0.15,-0.15)$) rectangle ($(v5.north east)+(0.15,0.15)$); \draw[very thick,dotted,red] ($(v6.south west)+(-0.15,-0.15)$) rectangle ($(v8.north east)+(0.15,0.15)$); \end{pgfonlayer} \end{scope} \end{tikzpicture} ~~~~ \begin{tikzpicture}[scale=0.5] \pgfmathtruncatemacro{\n}{6}; \pgfmathtruncatemacro{\vmax}{2*\n}; \pgfdeclarelayer{background} \pgfdeclarelayer{foreground} \pgfsetlayers{background,main,foreground} \begin{scope}[ vertex style/.style={ draw, circle, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected vertex style/.style={ draw, circle, fill=blue!20, minimum size=3mm, inner sep=0pt, outer sep=0pt% }, selected edge style/.style={ rounded corners,line width=1.5mm,blue!20,cap=round% } ] \node[selected vertex style] (v0) at (0,-1) {\tiny $0$}; \node[selected vertex style] (v1) at (0,0) {\tiny $1$}; \node[selected vertex style] (v2) at (1,0) {\tiny $2$}; \node[selected vertex style] (v3) at (2,-1) {\tiny $3$}; \node[vertex style] (v4) at (2,0) {\tiny $4$}; \node[selected vertex style] (v5) at (3,0) {\tiny $5$}; \node[selected vertex style] (v6) at (4,-1) {\tiny $6$}; \node[vertex style] (v7) at (4,0) {\tiny $7$}; \node[selected vertex style] (v8) at (5,0) {\tiny $8$}; \node[selected vertex style] (v9) at (6,-1) {\tiny $9$}; \node[vertex style] (v10) at (6,0) {\tiny $10$}; \node[selected vertex style] (v11) at (7,0) {\tiny $11$}; \node[selected vertex style] (v12) at (8,-1) {\tiny $12$}; \node[vertex style] (v13) at (8,0) {\tiny $13$}; \node[vertex style] (v14) at (9,0) {\tiny $14$}; \draw[thick,color=blue] (v0) -- (v1); \draw[thick,color=blue] (v3) -- (v4); \draw[thick,color=blue] (v6) -- (v7); \draw[thick,color=blue] (v9) -- (v10); \draw[thick,color=blue] (v12) -- (v13); \draw[thick,color=blue] (v1) -- (v2); \draw[thick,color=red] (v2) -- (v4); \draw[thick,color=blue] (v4) -- (v5); \draw[thick,color=red] (v5) -- (v7); \draw[thick,color=blue] (v7) -- (v8); \draw[thick,color=red] (v8) -- (v10); \draw[thick,color=blue] (v10) -- (v11); \draw[thick,color=red] (v11) -- (v13); \draw[thick,color=blue] (v13) -- (v14); \begin{pgfonlayer}{background} \draw[selected edge style] (v0.center) -- (v1.center) -- (v2.center); \draw[selected edge style] (v5.center) -- (v5.center) ; \draw[selected edge style] (v6.center) -- (v6.center) ; \draw[selected edge style] (v8.center) -- (v8.center) ; \draw[selected edge style] (v9.center) -- (v9.center) ; \draw[selected edge style] (v11.center) -- (v11.center) ; \draw[selected edge style] (v12.center) -- (v12.center) ; \draw[very thick,dotted,red] ($(v0.south west)+(-0.15,-0.15)$) -- ($(v1.north west)+(-0.15,0.15)$) -- ($(v2.north east)+(0.15,0.15)$) -- ($(v2.south east)+(0.15,-0.15)$) -- ($(v1.south east)+(0.15,-0.15)$) -- ($(v0.south east)+(0.15,-0.15)$) -- ($(v0.south west)+(-0.15,-0.15)$) ; \draw[very thick,dotted,green] ($(v3.south west)+(-0.15,-0.15)$) rectangle ($(v3.north east)+(0.15,0.15)$); \draw[very thick,dotted,black] ($(v5.south west)+(-0.15,-1.15)$) rectangle ($(v13.north east)+(0.15,0.15)$); \end{pgfonlayer} \end{scope} \end{tikzpicture} \caption{(left) $U^p_{10}$ with two $P_{3}$ embedded, and (right) $U^p_{10}$ with one $P_3$ and a number of $P_1$ embedded. Embedded paths are shown in \textcolor{blue!40}{blue}, and allocated blocks in \textcolor{red}{red}. The extra allocation to handle one $P_3$ and one $P_1$ is shown in dotted \textcolor{green}{green}. The trivial extension of $6$ additional $P_1$ is shown in dotted black. All cases use at most $\lfloor 3n/2 \rfloor $ vertices.} \label[figure]{fig:p3p1} \end{figure} \begin{lemma}\label{lem:combined} Let $k > 0$ unless otherwise specified. For the induced universal graph $U^p_{n}$ it holds that \begin{enumerate}[label=(\alph*)] \item $U^p_{2k}$ embeds $P_{2k}$. \label{case:p2k} \item $U^p_{2k}$ embeds $P_{2k+1}$ when $k > 1$. \label{case:p2k1} \item $U^p_{k}$ embeds $k$ $P_1$. \label{case:p1} \item $U^p_{6}$ embeds two $P_3$. \label{case:2p3} \item $U^p_{3+l}$ embeds one $P_3$ and $l \geq 0$ $P_1$. \label{case:p3p1} \end{enumerate} \end{lemma} \begin{proof} We show that the cases in \Cref{lem:combined} can be embedded in our induced universal graph. The general strategy is to allocate blocks as in \Cref{def:block} and embed the input in these blocks, where at least two out of the three vertices in a block are used, yielding the desired ratio between allocated and used vertices. Consider first case \ref{case:p2k}, where we are to embed an even path. See \Cref{fig:p2k1}. We allocate $k$ block $B_0, \ldots, B_{k-1}$ and embed the path by using three vertices in $B_0$, one vertex in $B_{k-1}$, and two vertices in $B_1, \ldots, B_{k-2}$. Say $n=2k$, then the size of the induced universal graph is thus $\lfloor 3n/2 \rfloor $. In the second case \ref{case:p2k1} we are to embed an odd path. See \Cref{fig:p2k1}. We use the same strategy as above, the difference being that two vertices are used in $B_{k-1}$, namely the two with the smallest label. As we embed one more vertex in the same number of blocks we are still within $\lfloor 3n/2 \rfloor $. For \ref{case:p1} we can embed a number $k$ of $P_1$ by allocating $\lfloor k/2 \rfloor$ blocks and embedding two $P_1$ per block. If $k$ is odd we embed the last $P_1$ by allocating the isolated vertex labelled $\floor{3k/2}-1$. When the input is \ref{case:2p3} we embed the two $P_3$ by allocating three blocks as seen in \Cref{fig:p3p1}. As we allocate $9$ vertices and use $6$, this strategy is within the $\lfloor 3n/2 \rfloor $ bound. In the final case \ref{case:p3p1} we allocate the first block $B_0$ and embed the $P_3$ in the block. Next we apply case \ref{case:p1} on the remaining $P_1$. See \Cref{fig:p3p1}. In total we then use $3 + \lfloor 3l/2 \rfloor$ vertices. \end{proof} We are now ready to state the main theorem of this section. \Cref{case:main} follows from consecutive applications of \Cref{lem:combined} and a careful order in which we embed the vertices of some given graph $G\in\mathcal{AC}$ with $n$ vertices in $U^p_n$. In particular, it is crucial that the $P_3$'s and $P_1$'s in $G$ are embedded second to last and last, respectively, since after embedding one of these two parts of $G$, the next unallocated vertex is not necessarily the first vertex of a new block. \begin{theorem}\label{thm:generalupperbound}\label{thm:paths} The graph family $U^p_0, U^p_1, \ldots$ has the property that for $n \in {\mathbb{N}}_1$ \begin{enumerate}[label=(\alph*)] \item $U^p_n$ has $\lfloor 3n/2 \rfloor$ vertices and $\max\set{0, 3\floor{\frac{n}{2}}-1}$ edges.\label{case:size} \item $U^p_n$ is an induced subgraph of $U^p_{n+1}$.\label{case:induced} \item $U^p_n$ is an induced universal graph for the family of acyclic graphs $G$ with $\Delta(G)\leq2$ and $n$ vertices.\label{case:main} \end{enumerate} \end{theorem} \begin{proof} Cases \ref{case:size} and \ref{case:induced} follow immediately from \Cref{def:upn}. Say the input graph $G$ consists of set $S_{\mathrm{even}}$ of $P_{2k}$ for $k \geq 1$, set $S_{\mathrm{odd}}$ of $P_{2k+1}$ for $k \geq 2$, set $S_{3}$ of $P_3$, and set $S_{1}$ of $P_1$. Let the cardinality of each set be the number of paths in the set and let the total number of vertices in each set be denoted $l_{\mathrm{even}}$, $l_{\mathrm{odd}}$, $l_{3}$, and $l_1$ respectively. Observe that every acyclic graph with $\Delta(G)\leq2$ has such a partition. The proof direction is to embed the different parts of the input in the right order, such that our embedding strategy from \Cref{lem:combined} can be applied. We start by embedding $S_{\mathrm{even}}$ by allocating the at most $3l_{\mathrm{even}}/2$ first vertices of $U^p_{n}$ by $|S_{\mathrm{even}}|$ invocations of \Cref{lem:combined} case \ref{case:p2k}. It follows that the next unallocated vertex is the first vertex of a new block, i.e., after embedding the even length paths we have spent at most $l_{\mathrm{even}}/2$ of blocks of size $3$. We allocate the next $\lfloor 3l_{\mathrm{odd}}/2 \rfloor$ vertices of $U^p_{n}$ by performing $|S_{\mathrm{odd}}|$ invocations of \Cref{lem:combined} case \ref{case:p2k1}. Again, the next unallocated vertex is the first vertex of a new block. The next step is embedding $S_3$ and $S_1$. If $|S_3|$ is even then proceed by invoking \Cref{lem:combined} case \ref{case:2p3} $|S_3|/2$ times to embed all pairs of $P_3$. This allocation uses $3 l_3 /2$ vertices from $U^p_{n}$. We can then embed $S_1$ trivially as in \Cref{lem:combined} case \ref{case:p1}, allocating additionally $\lfloor 3 l_1 / 2 \rfloor$ vertices of $U^p_{n}$. If $|S_3|$ is odd then allocate all the pairs using $|S_3|-1$ invocations of \Cref{lem:combined} case \ref{case:2p3}, which uses $3(l_3 - 3)/2$ vertices of $U^p_{n}$. We proceed by invoking \Cref{lem:combined} case \ref{case:p3p1} to embed the remaining $P_3$ along with $S_1$. In total we need to allocate at most $\lfloor 3(l_{\mathrm{even}} + l_{\mathrm{odd}} + l_3 + l_1 )/2 \rfloor = \lfloor 3n/2 \rfloor$ vertices, and hence $U^p_{n}$ induces any acyclic graph $G$ on $n$ vertices where $\Delta(G)\leq2$. \end{proof} \subsection{Preliminaries}\label{sec:prelim} Let $[n]=\{0, \ldots, n-1\}$, ${\mathbb{N}}_0 = \set{0,1,2,\ldots}$, ${\mathbb{N}}={\mathbb{N}}_1=\set{1,2,\ldots}$, and let $\log n$ refer to $\log_2 n$. For a graph $G$, let $V[G]$ be the set of vertices and $E[G]$ be the set of edges of $G$, and let $\abs{G}=\abs{V[G]}$ be the number of vertices. We denote the maximum degree of graph $G$ as $\Delta(G)$. For $i \in {\mathbb{N}}$, let $P_{i}$ denote a path with $i$ vertices, and for $i>2$, let $C_i$ denote a simple cycle with $i$ vertices. Let $G$ and $U$ be two graphs and let $\lambda\colon V[G] \to V[U]$ be an injective function. If $\lambda$ has the property that $uv\in E[G]$ if and only if $\lambda(u)\lambda(v)\in E[U]$, we say that $\lambda$ is an \emph{embedding function} of $G$ into $U$. $G$ is an \emph{induced subgraph} of $U$ if there exists an embedding function of $G$ into $U$, and in that case, we say that $G$ is \emph{embedded} in $U$ and that $U$ \emph{embeds} $G$. Let $\mathcal F$ be a family of graphs. $U$ is an \emph{induced universal graph} of $\mathcal F$ if $G$ is an induced subgraph of $U$ for each $G\in\mathcal F$. \newcommand{\mathcal{F}}{\mathcal{F}} Let $\mathcal{F}$ be a family of graphs and for each positive integer $n$, let $\mathcal{F}_n$ be the graphs in $\mathcal{F}$ on $n$ vertices. For $f : {\mathbb{N}}_0 \to {\mathbb{N}}_0$ we say that $\mathcal{F}$ can be labeled using $f(n)$ labels with a \emph{size aware decoder} if there exists a decoder $d_n : {\mathbb{N}}_0 \times {\mathbb{N}}_0 \to \{0,1\}$ for $n\in{\mathbb{N}}$, that satisfies the following: For every graph $G \in \mathcal{F}_n$ there is an encoding function $e : G \to [f(n)]$ such that $u,v \in G$ are adjacent iff $d_n(e(u),e(v)) = 1$. We say that $\mathcal{F}$ can be labeled using $f(n)$ labels with a \emph{size oblivious decoder} if there exists $d : {\mathbb{N}}_0 \times {\mathbb{N}}_0 \to \{0,1\}$ with the following property: For every $n\in{\mathbb{N}}$ and every $G\in\mathcal{F}_n$, there is an encoding function $e : G \to [f(n)]$ such that $u,v \in G$ are adjacent iff $d(e(u),e(v)) = 1$. The following theorem explains the relation between size aware and size oblivious decoders and induced universal graphs. \begin{proposition}\label[proposition]{pro:aware} Given a family of graphs $\mathcal{F}$ and a function $f : {\mathbb{N}} \to {\mathbb{N}}$ the following holds. 1) $\mathcal{F}$ can be labeled using $f(n)$ labels with a size aware decoder iff there exists a family of graphs $G_1, G_2, \ldots$ such that $G_n$ is an induced universal graph for $\mathcal{F}_n$ and $\abs{G_n} = f(n)$ for every $n$. 2) $\mathcal{F}$ can be labeled using $f(n)$ labels with a size oblivious decoder iff there exists a family of graphs $G_1, G_2, \ldots$ such that $G_n$ is an induced universal graph for $\mathcal{F}_n$, $\abs{G_n} = f(n)$, and $G_n$ is an induced subgraph of $G_{n+1}$ for every $n$. \end{proposition} \begin{proof} 1) First assume that $\mathcal{F}$ can be labeled using $f(n)$ labels with a size aware decoder. Let $d_n : {\mathbb{N}}_0 \times {\mathbb{N}}_0 \to \{0,1\}$ be the corresponding family of decoders. Let $G_n$ be defined on the vertex set $[f(n)]$ such that $i \neq j$ are adjacent iff $d_n(i,j) = 1$. Obviously, $G_n$ contains exactly $f(n)$ nodes. Furthermore $G_n$ is an induced universal graph of $\mathcal{F}_n$. This shows the first direction. For the other direction let such a family $G_1, G_2, \ldots$ be given. Wlog assume that the vertex set of $G_n$ is $[f(n)]$. We let $d_n : {\mathbb{N}}_0 \times {\mathbb{N}}_0 \to \{0,1\}$ be defined by $d_n(u,v) = 1$ iff $u,v < f(n)$ and $u,v$ are adjacent in $G_n$. For $G \in \mathcal{F}_n$ let $\lambda : G \to G_n$ be an embedding function of $G$ in $G_n$. For a node $u \in V[G]$ we assign it the label $\lambda(u)$. Then for nodes $u,v \in V[G]$ we have $d_n(\lambda(u),\lambda(v)) = 1$ iff $\lambda(u)$ and $\lambda(v)$ are adjacent. This happens iff $u$ and $v$ are adjacent as desired. 2) The first direction is as in the first part of the theorem. We just need to note that $\lambda : [f(n)] \to [f(n+1)]$ is an embedding function of $G_n$ in $G_{n+1}$. Now consider the other direction and let such a family $G_1, G_2, \ldots$ be given and wlog assume that the vertex set of $G_n$ is $[f(n)]$. Furthermore wlog assume that $\lambda_n : G_n \to G_{n+1}$ given by $k \to k$ is an embedding function of $G_n$ in $G_{n+1}$. Let $d : {\mathbb{N}}_0 \times {\mathbb{N}}_0 \to \{0,1\}$ be defined in the following way. For $u,v \in {\mathbb{N}}_0$ let $d(u,v) = 1$ iff there exists $n$ such that $f(n) > u,v$ and $u$ and $v$ are adjacent in $G_n$. We note that if one such $n$ exists it must hold for all $n$ with $f(n) > u,v$ that $u$ and $v$ are adjacent in $G_n$. Hence we can assign labels in the same way as the first part of the theorem. \end{proof}
fb1c3050dbb51572f2beb82b39cdc35a4990c1e4
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} For the first decade of microlens planet detections, beginning with OGLE-2003-BLG-235Lb \citep{ob03235}, the great majority of detections required a combination of survey and followup data. This is a consequence of two effects. First, the survey coverage was generally too sparse to characterize the planetary anomalies in the detected events \citep{gouldloeb}. Second, thanks to aggressive alert capability, pioneered by the Optical Gravitational Lensing Experiment (OGLE) Early Warning System (EWS, \citealt{ews1,ews2}), it became possible to organize intensive followup of planet-sensitive events -- or even ongoing planetary anomalies -- and so obtain sufficient time resolution to detect and characterize planets. However, as surveys have become more powerful over the past decade, they have become increasingly capable of detecting planets without followup observations. That is, making use of larger cameras, the surveys are able to monitor fairly wide areas at cadences of up to several times per hour. While still substantially lower than followup observations of the handful of events that were monitored by followup groups, this is still adequate to detect most planets (provided that the anomalies occur when the survey is observing). Very simple reasoning given below, which is supported by detailed simulations \citep{Zhu:2014}, leads one to expect that the transition from survey+followup to survey-only mode implies a corresponding transition from planets detected primarily in high-magnification events via central and resonant caustics to planets primarily detected in lower magnification events via planetary caustics. High-magnification events are intrinsically sensitive to planets because they probe the so-called ``central caustic'' that lies close to (or overlays) the position of the host \citep{griest98}. Planets that are separated from the hosts by substantially more (less) than the Einstein radius generate one (two) other caustics that are typically much larger than the central caustic and thus have a higher cross section for anomalous deviations from a point-lens light curve due to a random source trajectory. However, for high-magnification events, the source is by definition passing close to the host and hence close to or over the central caustic. For planet-host separations that are comparable to the Einstein radius, the two sets of caustics merge into a single (and larger) ``resonant caustic'', which is even more likely to generate anomalous deviations of a high-magnification event. For many years, the Microlensing Follow Up Network ($\mu$FUN) employed a strategy based on this high planet sensitivity of high-magnification events. They made detailed analyses of alerts of ongoing events from the OGLE and the Microlensing Observations in Astrophysics (MOA) teams to predict high-magnification events and then mobilized followup observations over the predicted peak. \citet{gould10} showed that $\mu$FUN was able to get substantial data over peak for about 50\% of all identified events with maximum magnification $A_{\rm max}>200$, but that its success rate dropped off dramatically at lower magnification, i.e., even for $100<A_{\rm max}<200$. The reason for this drop off was fundamentally limited observing resources: there are twice as many events $A_{\rm max}>100$ compared to $A_{\rm max}>200$, and monitoring the full-width half-maximum requires twice as much observing time. Hence, observations grow quadratically with effective magnification cutoff. By contrast, because planetary caustics are typically much larger than central caustics, most planets detected in survey-only mode are expected to be from anomalies generated by the former, which occur primarily in garden-variety (rather than high-mag) events \citep{Zhu:2014}. For example, \citet{ob120406} detected a large planetary caustic in OGLE-2012-BLG-0406 based purely upon OGLE data, while \citet{moabin1} detected one in MOA-bin-1 based mostly on MOA data. In the latter case it would have been completely impossible to discover the planet by survey+followup mode because the ``primary event'' (due to the host) was so weak that it was never detected in the data. Nevertheless, there has been a steady stream of survey-only detections of planets in high-magnification events as well. The first of these was MOA-2007-BLG-192Lb, a magnification $A_{\rm max}> 200$ event, which required a combination of MOA and OGLE data \citep{mb07192}. The first planet detected by combining three surveys (MOA, OGLE, Wise), MOA-2011-BLG-322Lb, was also via a central caustic, although in this case the caustic was very large so that the magnification did not have to be extremely large $(A_{\rm max}\sim 20)$ \citep{mb11322}. Similarly, \citet{ob150954} detected a large central caustic due to the large planet OGLE-2015-BLG-0954Lb despite modest peak magnification of the underlying event $A_{\rm max}\sim 20$. This case was notable because high-cadence data from the Korea Microlensing Telescope Network (KMTNet) captured the caustic entrance despite the extremely short source self-crossing time, $t_*=16\,$min. There also exist 2 planets MOA-2008-BLG-379Lb \citep{Suzuki2014} and OGLE-2012-BLG-0724Lb \citep{Hirao2016} that were detected by the OGLE+MOA surveys through the high-magnification channel. KMTNet is still in the process of testing its reduction pipeline. Motivated by the above experience, the KMTNet team focused its tests on high-magnification events identified as such on the OGLE web page. In addition to exposing the reduction algorithms to a wide range of brightnesses, this testing has the added advantage that there is a high probability to find planets. Here we report on the first planet found by these tests from among the first seven high-mag events that were examined: OGLE-2016-BLG-(0261,0353,0471,0528,0572,0596,0612). These events were chosen to have model point-lens magnifications $A>20$ and modeled peak times $2457439<t_0<2457492$. The lower limit was set by the beginning of the KMTNet observing season and the upper limit was the time of the last OGLE update when the seven events were selected. \section{Observations} On 2016 April 8 UT 12:15 (HJD$^\prime=$ HJD$-2450000=7487.0$), OGLE alerted the community to a new microlensing event OGLE-2016-BLG-0596 based on observations with the 1.4 deg$^2$ camera on its 1.3m Warsaw Telescope at the Las Campanas Observatory in Chile \citep{ogleiv} using its EWS real-time event detection software \citep{ews1,ews2}. Most observations were in $I$ band, but with some $V$ band observations that are, in general, taken for source characterization. These $V$-band data are not used in the modeling. At equatorial coordinates $(17^{\rm h} 51^{\rm m} 12^{\rm s}\hskip-2pt .81, -30^\circ 50' 59''\hskip-2pt.4)$ and Galactic coordinates $(-1.01^\circ,-2.03^\circ)$, this event lies in OGLE field BLG534, with an observing cadence during the period of the anomaly of roughly 0.4 per hour\footnote{OGLE cadences were significantly adjusted at the time of the peak and planetary anomaly of this event, due to the {\it Kepler} K2 Campaign 9 microlensing campaign. The five fields covering the K2 field were observed 3 times per hour, while other fields (including BLG534) were observed somewhat less frequently (very roughly 2/3) compared to their usual rates.}. KMTNet employs three $4.0\,\rm deg^2$ cameras mounted on 1.6m telescopes at CTIO/Chile, SAAO/South Africa, and SSO/Australia \citep{kmtnet}. In 2015 KMTNet had concentrated observations on 4 fields. However, in 2016, this strategy was radically modified to cover (12, 40, 80) $\rm deg^2$ at cadences of $(4,\ \geq 1,\ \geq 0.4)\,\rm hr^{-1}$. For the three highest-cadence fields, KMTNet observations are alternately offset by about $6^\prime$ in order to ensure coverage of events in gaps between chips. As a result, OGLE-2016-BLG-0596 lies in two slightly offset fields BLG01 and BLG41, which are each observed at a cadence 2 per hour\footnote{Like OGLE, KMTNet also adjusted its schedule for the K2 campaign, but in a different way. First, CTIO observations were not adjusted. Second, KMTNet only began ``K2 mode'' on 2016 April 23. This was after the event peak and caustic entrance but before the exit. Therefore, in particular, the caustic exit observations from SAAO were at the lower cadence (reduced by a factor 0.75)}. KMTNet observes primarily in $I$ band, but 1/11 observations from CTIO and 1/21 observations from SAAO are in $V$-band. Reductions of the primary data were made using difference image analysis (DIA) \citep{alard98}. However, due to issues discussed in Section 3, special variants of DIA were developed specifically for this event. See below. KMT CTIO $V$ and $I$ images were, in addition, reduced using DoPHOT \citep{dophot}, solely for the purpose of determining the source color. \begin{figure} \includegraphics[width=\columnwidth]{fig1.eps} \caption{ Correction of light curve variability. Top panel shows OGLE online reductions. Variability is roughly periodic, $P\simeq 126.5$ days, and semi-amplitude of $\sim 8.7\%$ of baseline flux, but 57\% of source flux. Second panel shows result from simultaneous fit to (1) microlensed source, (2) bright variable at $1.5^{\prime\prime}$, (3) nearby blended star. Periodic variability is removed but annual trend remains. Mean flux drops by 0.5 mag due to fitting out third star. Bottom panel shows measured flux (from second panel) but as a function of sidereal time. This is well fit by a straight line. Note that the full range of this fit to the variation is a factor $\sim 1.5$ larger than the source flux. Final photometry (third panel) is obtained by subtracting this straight-line fit from all flux measurements (not just at baseline). } \label{fig:one} \end{figure} \section{Light Curve Variability} \subsection{Evidence of Variability} The OGLE-2016-BLG-0596 light curve shows clear variability over the course of 6 seasons of OGLE-IV data prior to 2016. This variability is roughly consistent with being sinusoidal at period $P=126.5\,$ days. See Figure~\ref{fig:one}. While the semi-amplitude of the variability is only $8.7\%$ of the baseline flux $(I_{\rm base}\sim 19)$, this semi-amplitude turns out to be roughly equal to the source flux derived from the model. \subsection{Importance of Variability} Assuming (as proves to be the case) that it is not the source itself that is variable, such low level variability cannot significantly impact characterization of the anomalous features of the lightcurve because they are at relatively high magnification and take place on much shorter timescales. However, if not properly accounted for, such variability can seriously impact the estimate of the source flux and, as a direct consequence of this, the Einstein timescale \begin{equation} t_{\rm E} = {\theta_{\rm E}\over\mu}; \qquad \theta_{\rm E}^2\equiv \kappa M \pi_{\rm rel}; \qquad \kappa\equiv {4 G\over c^2\au}\simeq 8.14{{\rm mas}\over M_\odot}. \label{eqn:tedef} \end{equation} Here $\theta_{\rm E}$ is the angular Einstein radius, $\pi_{\rm rel} = \au(D_L^{-1}-D_S^{-1})$ is the lens-source relative parallax, and $\mu$ is the lens-source relative proper motion in the Earth frame. Errors in these quantities would propagate into the estimates of the planet-star mass ratio $q$, the Einstein radius $\theta_{\rm E}$, and the proper motion $\mu$, all of which are important for assessing the physical implications of the detection. The reason that $t_{\rm E}$ is potentially impacted by unmodeled variability is that it is determined primarily from the wings of the light curve where the amplitude of the amplification of flux is comparable to that of the variability. Hence, it is important to track down the source of this variability and correct for it to the extent possible. See, e.g., \citet{mb11293}. \subsection{Removal of Variability I: Variations of Neighbor} In principle, tracking down such a low level of variability in such a crowded field could have been very difficult. However, in the present case, it turns out to be due to a star $1.5^{\prime\prime}$ to the southeast, which is quite bright $I\sim 14.5$ and shows variability with the same period and phase. Within the framework of standard DIA, it is natural that this variable should impact the microlensing light curve because the difference image contains residuals from the variable that overlap the point spread function (PSF) of the microlensed star. Hence, when the difference image is dot-multiplied by the PSF to estimate the flux, it includes a contribution from the residual flux of the variable. It is straightforward to simultaneously fit for two (or $n$) variables with possibly overlapping PSFs. After constructing difference images in the standard way, one simply generalizes the normal procedure by calculating the $n$ flux-difference values \begin{equation} F_i = \sum_{j=1}^n c_{ij} d_j; \qquad c\equiv b^{-1}, \label{eqn:multifit1} \end{equation} where \begin{equation} b_{ij} = \sum_k {P_{i,k}P_{j,k}\over \sigma_k^2}; \qquad d_i = \sum_k {P_{i,k}f_k\over \sigma_k^2}, \label{eqn:multifit2} \end{equation} $P_{i,k}$ is the (unit normalized: $\sum_k P_{i,k}=1$) amplitude of the $i$th PSF in the $k$th pixel and $(f_k,\sigma_k)$ are the value and error of the difference flux in the $k$th pixel. We use a variant of this formalism to reduce the OGLE data with $n=3$ stars, including the microlensed source, the bright variable, and one other very nearby (but non-variable) blended star. The result is shown in the second panel of Figure~\ref{fig:one}. First note that the fluxes have decreased by about 0.5 mag because the non-variable neighboring blend (third star in the fit) has been removed from the baseline flux. The semi-periodic variations are removed. However there remains an annual trend. \subsection{Removal of Variability II: Annual Variations} The bottom panel of Figure~\ref{fig:one} shows that this annual trend is due to variations with sidereal time, almost certainly due to the impact of the bright red neighbor (even if it were constant) via differential refraction. We fit this variation to polynomials of order $n$, but find that there is no significant improvement beyond $n=1$, for which $f(t) = 0.2415 + 0.1490(t-0.5)$, where $f$ is flux in units of $I=18$. Note that the variation from 0.3 to 0.7 (on the figure) is $0.4\times 0.149\sim 0.06$ which is 1.5 times larger than the source flux derived below. The third panel shows the results of applying this sidereal-time correction. As expected the annual trend is gone. We apply this flux correction to all data, not just the baseline data shown in this figure. We find (as expected) that this corrects the slope of the rising part of the light curve, which indeed impacts the estimate of $t_{\rm E}$, though by less than 10\%. Note that while, in most cases, it is possible to derive an accurate estimate of the photometric error bars from those reported by the photometric pipeline \citep{skowron16}, in this case we did not apply this simple prescription because the data were reduced using a special customized pipeline. \subsection{Correction of KMTNet Data} We apply the same formalism given by Equations~(\ref{eqn:multifit1}) and (\ref{eqn:multifit2}) to the KMTNet data, but with only two stars, i.e., the microlensed source and the neighboring variable. Note that this difference between OGLE and KMTNet reductions plays no role in the final result because the third star incorporated into the OGLE fit is not variable, and the KMTNet flux scale is ultimately aligned to OGLE through the microlens fit. Thus, in particular, we retain the advantage of resolving out this blend, thus placing better limits on flux from the lens. We note that it is difficult to correct for the annual variation in the KMTNet data. Reliable measurement of the annual variation would require baseline data, which do not exist because the photometry system was not the same for the whole of 2015 compared to 2016. In principle, we could have applied the OGLE-based correction to KMTNet data, but this type of correction is observatory-specific and this would not have been a reliable approach and could easily cause more problem than what was being corrected for. From checking the impact of only correcting the OGLE data for variable-contamination (but not annual variation), however, we found essentially no change. The scatter (hence renormalized error bars) are very slightly smaller, but no change in parameters. The same would be case for KMTNet data. \begin{figure} \includegraphics[width=\columnwidth]{fig2.eps} \caption{ Light curve and geometry of OGLE-2016-BLG-0596. The event is primarily characterized by a strong caustic entrance at HJD$^\prime\sim 7486$ superposed on an otherwise slightly asymmetric point-lens-like light curve. There is a weak caustic exit at HJD$^\prime\sim 7502.5$ which is well covered by KMT SAAO data. This morphology, together with the $\sim 16$ day interval from caustic entrance to exit, is indicative of a resonant caustic (top panel) due to a high-mass planet or low-mass brown dwarf. } \label{fig:two} \end{figure} \subsection{Guideline for Assessing the Need of Multi-star Fitting} The formalism introduced in Section 3.3 can also be used to gain intuition about the impact of uncorrected variability, which can then be used to assess whether such corrections are necessary in specific cases. First note that for $n=1$, Equations~(\ref{eqn:multifit1}) and (\ref{eqn:multifit2}) reduce to the standard formula: \begin{equation} F_1 = {d_1\over b_{11}}; \qquad b_{11} = \sum_k {P_{1,k}^2\over \sigma_k^2}; \qquad d_1 = \sum_k {P_{1,k}f_k\over \sigma_k^2} \label{eqn:singlefit} \end{equation} Equation~(\ref{eqn:singlefit}) then allows us to express the properly corrected ``true'' photometry in terms of the ``naive'' single-source photometry that ignores neighbors. We first ``infer'' the value, $d_{i,\rm inferred} = b_{ii}*F_{i,\rm naive}$, which then yields, \begin{equation} F_{i,\rm true} = \sum_{j=1}^2 c_{ij}d_{j,\rm inferred}= {F_{i,\rm naive} - (b_{12}/b_{ii})F_{(3-i),\rm naive}\over 1 - b_{12}^2/(b_{11}b_{22})}. \label{eqn:backout} \end{equation} Hence, the correction is governed by the ratio of the PSF overlap integral $b_{12}$ to the integral of the PSF squared, $b_{11}$. We can evaluate this explicitly for the special case of a Gaussian PSF and below-sky sources ($\sigma_k =$const) \begin{equation} {b_{12}\over b_{11}} = 4^{-(\Delta \theta/{\rm FWHM})^2} \qquad (\rm below-sky\ Gaussian), \label{eqn:gaussbelow} \end{equation} where $\Delta\theta$ is the separation between the two sources. If the two sources are reasonably well separated, $\Delta\theta\ga$FWHM, and (as in the present case) the target (1) is below sky while the contaminating variable (2) is well above sky, then the effect is roughly half of that given by Equation~(\ref{eqn:gaussbelow}). This is because the squared PSF integral is basically unaffected while the half of the contribution to the overlap integral that is closer to the contaminant is heavily suppressed by the higher flux errors per pixel. We close by re-emphasizing that these order-or-magnitude estimates are not used in the present analysis but are intended as guidance for future cases. \begin{figure} \includegraphics[width=\columnwidth]{fig3.eps} \caption{ $\Delta\chi^2$ map of the MCMC chain in the $\log s$--$\log q$ parameter space obtained from the preliminary grid search. The lower panel shows the entire range where the grid search is conducted. The upper panel show the enlarged view around the best-fit solution. Color coding represents MCMC points with $\leq 1n\sigma$ (red), $2n\sigma$ (yellow), $3n\sigma$ (green), $4n\sigma$ (cyan), $5n\sigma$ (blue), and $6n\sigma$ (magenta) of the best-fit and $n=10$. } \label{fig:three} \end{figure} \section{Light Curve Analysis} The lightcurve, presented in Figure~\ref{fig:two}, has two principal features: a strong caustic entrance near peak at HJD$^\prime = 7486.4$ and a weak caustic exit at HJD$^\prime = 7502.6$. Apart from these caustic crossings, the morphology is that of a slightly distorted point-lens event. This morphology points to a binary lens with very unequal mass ratio $q\ll 1$, i.e., in the brown-dwarf or planetary regime. The long duration of the caustic (16 days) then points to a resonant caustic, and so projected separation (in units of $\theta_{\rm E}$) of $s\sim 1$. A thorough search of the parameter space spanning $-1.0\leq \log s \leq 1.0$ and $-5.0\leq \log q \leq 1.0$ leads to only one viable solution, which confirms the above naive reasoning. The uniqueness of the solution is shown in Figure~\ref{fig:three}, where we present the $\Delta\chi^2$ map of the MCMC chain in the $\log s$--$\log q$ parameter space obtained from the preliminary grid search. In fact, initial modeling based on data taken up through HJD$^\prime = 7500.8$ (so, before the caustic exit) already led to essentially this same solution (although the predicted caustic exit was 2.6 days later than the one subsequently observed). The model is described by seven parameters. These include three that are analogous to a point-lens event $(t_0,u_0,t_{\rm E})$, i.e., the time of closest approach to the center of magnification, the impact parameter normalized to $\theta_{\rm E}$ and the Einstein crossing time; three to describe the binary companion $(s,q,\alpha)$ where $\alpha$ is the angle of binary axis relative to the source trajectory; and $\rho\equiv \theta_*/\theta_{\rm E}$, where $\theta_*$ is the angular radius of the source. \begin{figure} \includegraphics[width=\columnwidth]{fig4.eps} \caption{ Caustic geometry of OGLE-2016-BLG-0596. Top panel shows the causic (red) with the positions of the host (left) and planet (right) represented as blue circles. Zoom in lower panel shows that the source passed close to, but did not cross (because $u_0>\rho$), the small central cusp. In this region, the caustic is very strong, accounting for the sharp jump at HJD$^\prime\sim 7486.4$. On the other hand the caustic exit to the right is very weak, which accounts for the smallness of the corresponding bump in the light curve at HJD$^\prime\sim 7502.6$. } \label{fig:four} \end{figure} \begin{deluxetable}{ccrr} \tablecaption{Microlensing parameters for OGLE-2016-BLG-0596\label{table:one}} \tablewidth{0pt} \tablehead{ \multicolumn{1}{c}{Parameter} & \multicolumn{1}{c}{Unit} & \multicolumn{1}{c}{best} & \multicolumn{1}{c}{error} } \startdata $t_0$ & day & 7486.464 & 0.010 \\ $u_0$ & $10^{-2}$ & 1.112 & 0.031 \\ $t_{\rm E}$ & day & 81.694 & 2.195 \\ $s$ & & 1.075 & 0.003 \\ $q$ & $10^{-2}$ & 1.168 & 0.040 \\ $\alpha$ & radian & 5.886 & 0.009 \\ $\rho$ & $10^{-2}$ & 0.060 & 0.008 \\ $I_s$ & & 21.510 & 0.028 \\ $I_b$ & & 19.739 & 0.028 \enddata \end{deluxetable} The best fit parameters and errors (determined from a Markov Chain) are given in Table~\ref{table:one}. We present the model light curve superposed on the data points in Figure~\ref{fig:two} and the lens geometry is shown in Figure~\ref{fig:four}. We also fit the lightcurve for the microlens parallax effect, but found no improvement. We note that compared to other planetary and binary events with well-covered caustic crossings, the parameter $\rho=(6.0\pm 0.8)\times 10^{-4}$ (and the parameter combination $t_* = \rho t_{\rm E} = 0.049 \pm 0.007$) have relatively large errors. These parameters are usually better measured because caustic crossings tend to be bright (since the caustic itself is a contour of formally infinite magnification), which means that the photometry over the caustic crossing is relatively precise. Since $t_*$ depends almost entirely on the duration of this crossing, with only weak dependence on other model parameters, it can then be determined quite precisely. In the present case, however, the first crossing was entirely missed simply because it was not visible from any of the five survey telescopes currently in operation (OGLE, MOA, and three from KMTNet: CTIO, SAAO, SSO). The caustic exit was captured by KMTNet SAAO, with 12 points taken over 3.63 hours (i.e., 20 minute cadence). However, since this caustic was quite weak, peaking at $I\sim 18.1$, the photometry has much larger errors than the SAAO photometry near peak. See upper two panels of Figure~\ref{fig:two}. Adopting a more glass-half-full orientation, we should assess the prior probability that either of the two caustic crossings would have been adequately observed to measure $\rho$. Considering the 20 days between HJD$^\prime$ 7485 and 7505, the three KMTNet observatories each took at least two points on 13 nights, with total durations, (2.11,2.76,2.52) days for SSO, SAAO, and CTIO, i.e., a total 7.39 days. Essentially all of these 39 intervals had approximately continuous coverage. We estimate that the probability that $\rho$ can be measured is the same as the probability that the caustic peak is covered, which may be slightly too conservative. Under this assumption, the probability that the caustics would be observed are each 37\%, so that the probability that at least one would be observed is $1-(1-0.37)^2=46\%$. Of course, since the midpoint of the two caustic crossings was 16 April, this probability is adversely affected by the shortness of the bulge observing window relative to microlensing ``high season'' (21 May -- 21 July). At that time the observing window is roughly 2.5 hours longer, and so (assuming comparable weather conditions), the probability for each crossing would be 52\% and the probability for at least one would be 77\%. Nevertheless, the mid-April values may be considered as a proxy for the microlensing season as a whole. \begin{figure} \includegraphics[width=\columnwidth]{fig5.eps} \caption{ Instrumental CMD for $100^{\prime\prime}$ square around OGLE-2016-BLG-0596 using KMTNet CTIO data. The instrumental source color is measured from model-independent regression and the instrumental magnitude is measured from the fit of the $I$-band data to the model light curve. By measuring the offset of this source position from that of the red clump (red), one can determine the angular source radius $\theta_*$, using standard techniques \citep{ob03262}, as described in the text. } \label{fig:five} \end{figure} \section{Physical Parameters} We use KMTNet CTIO DoPHOT reductions to construct an instrumental color magnitude diagram (CMD) that is presented in Figure~\ref{fig:five}.\footnote{ Correction of the DoPHOT data for variation is not done, but this would have little effect on the result. The variable is extraordinarily red, $\sim 1.2$ magnitude redder than the clump, whereas the source is $\sim 0.2$ magnitude bluer than the clump. Hence, by a naive estimate, the variations would be fractionally smaller by a factor 4. The full amplitude of these variations in $I$ band is of order the source flux, whereas the color measurement is made when the source is magnified 60 to 100 times. The color measurement is differential over short timescales of a few days, whereas the period is a large fraction of a year. Combining these very small factors, we expect the color measurement to be impacted at the level $(1/4)\times (1/80)\times (3/(126/\pi)) \sim 2\times 10^{-4}$. It is general practice to ignore such small errors, which in this case are more than hundred times smaller than the measurement error. We also note that the dependence of the color measurement on the choice of the $V$-band data set (OGLE or KMTNet) is small considering that the offset from the clump has an accuracy of 0.03 magnitude whereas the precision of the color measurement is 0.05 magnitude. Furthermore, the SAAO $V$-band data are taken for redundancy, primarily in a case there is no CTIO data due to bad weather when the event is well magnified or for very short, highly magnified events that peak of South Africa. } We find the instrumental source color from model-independent regression and the instrumental source magnitude by fitting the $I$ band light curve to the model. We then find the offset from the clump $\Delta[(V-I),I]=(-0.23,4.06)\pm (0.03,0.10)$, where the error in the color offset is dominated by the regression measurement while the error in the magnitude offset is dominated by fitting for the clump centroid. We then adopt $[(V-I),I]_{0,\rm clump}=(1.06,14.49)$ \citep{bensby13,nataf13} to obtain $[(V-I),I]_{0,s}=(0.83,18.55)$. Then using standard techniques \citep{ob03262}, we convert from $V/I$ to $V/K$ using the \citet{bb88} $VIK$ color-color relations and then use the \citet{kervella04} color/surface-brightness relations to derive \begin{equation} \eqalign{ \theta_* = 0.690\pm 0.065\,\ \mu{\rm as}; \cr \theta_{\rm E} = {\theta_*\over\rho} = 1.15\pm 0.18\ {\rm mas}; \cr \mu = {\theta_{\rm E}\over t_{\rm E}} = 5.1\pm 0.8\,\ {\rm mas}\,{\rm yr}^{-1} .\cr } \label{eqn:thetastar} \end{equation} The error in $\theta_*$ is dominated by the uncertainties in transforming from color to surface brightness (8\%), with a significant contribution from the error in $I_s$ (5\%). The fractional errors in $\theta_{\rm E}$ and $\mu$ are substantially larger than in $\theta_*$ due to the relatively large error in $\rho$. See Section 4. The relatively large value of $\theta_{\rm E}$ almost certainly implies that the lens lies in the Galactic disk since the lens-source relative parallax is \begin{equation} \pi_{\rm rel} = {\theta_{\rm E}^2\over\kappa M} = (0.16\pm 0.05 {\rm mas}) \biggl({M\over M_\odot}\biggr)^{-1}. \label{eqn:mpirel} \end{equation} That is, only if the lens were substantially heavier than $1\,M_\odot$ could it be in the bulge ($\pi_{\rm rel}\la 0.03$). However, first, there are almost no such massive stars in the bulge and second, its light would then exceed the blended light $(I_b\sim 19.7)$, even allowing for the $A_I=2.96$ extinction toward this line of sight \citep{nataf13}. The only exception to this line of reasoning would be if the lens were a black hole. Although the model considering parallax effects does not have improvement compared to the non-parallax model, non-detection of $\pi_{\rm E}$ can give constraints on the mass and distance. In Figure~\ref{fig:six}, we present the $\Delta\chi^2$ map of the MCMC chain in the $\pi_{{\rm E},E}$--$\pi_{E,{\rm N}}$ parameter space obtained from the modeling considering both the lens orbital motion and the microlens parallax effect. The upper limit of the microlens parallax as measured 3$\sigma$ level is $\pi_{\rm E} \lesssim 0.4$. This gives the lower limits of the mass and distance of $M\gtrsim 0.35\ M_\odot$ and $D_{\rm L}\gtrsim 1.7$ kpc. \begin{figure} \includegraphics[width=\columnwidth]{fig6.eps} \caption{ $\Delta\chi^2$ map of the MCMC chain in the $\pi_{{\rm E},E}$--$\pi_{{\rm E},N}$ parameter space. Color coding represents points in the MCMC chain with $\leq 1\sigma$ (red), $2\sigma$ (yellow), $3\sigma$ (green), $4\sigma$ (cyan), and $5\sigma$ (blue) of the best fit. The dotted circles represent the boundaries of $\pi_{\rm E}=0.1$, 0.2, 0.3, 0.4, and 0.5. } \label{fig:six} \end{figure} \section{Resolving the Nature of the Planet} The most notable characteristic of OGLE-2016-BLG-0596 is its high mass ratio $q=0.0117\pm 0.0004$, implying that the mass is $m_p = 12.2\,M_{\rm jup} (M/M_\odot)$. Hence, if the host is one solar mass, this planet would be just below the the deuterium-burning limit (usually regarded as the planet/brown-dwarf boundary). While the host could in principle have arbitrarily low mass (and so, by Equation~(\ref{eqn:mpirel}), be arbitrarily close), distances closer than $D_L\la 1\,{\rm kpc}$ are strongly disfavored by the relatively low proper motion, the parallax constraint, and the paucity of nearby lenses. At this limiting distance, and so $M =\theta_{\rm E}^2/\kappa\pi_{\rm rel} \sim 0.18\,M_\odot$, the planet would still be $m_p\sim 2\,M_{\rm jup}$, i.e., quite massive for such a low-mass host. Hence, regardless of the host mass, this is a fairly extreme system. To distinguish among these interesting possibilities will require measuring (or strongly constraining) the host mass. This can be accomplished with high resolution imaging, either using the {\it Hubble Space Telescope (HST)} or ground-based adaptive optics (AO) imaging on an 8m class telescope. An advantage of {\it HST} is that it can observe in the $I$ band for which the source flux is directly measured from the event. Hence, the source light can be most reliably separated from the blended light in $I$. In contrast to many previous cases, there are no $H$ band observations during the event, so ground-based AO observations (which must be in the infrared) cannot be directly compared to an event-derived source flux. Nevertheless, it is probably possible to transform from $V/I$ light-curve measurement to $H_s$ source flux with a precision of 0.2 mag, using a $VIH$ color-color diagram. For definiteness, we will assume that the lens can be reliably detected from {\it HST} or AO observations provided that the flux is at least half that of the source, i.e., $I_L<22.3$. For example, if the lens were an $M=0.5\,M_\odot$ early M dwarf (so $D_L\sim 2.2\,{\rm kpc}$), then it would have $I_L \sim 18.8 + A_I\leq 21.8$ (and brighter if, as is almost certainly the case, a substantial fraction of the dust is beyond 2.2 kpc). Thus, there is a good chance that AO or {\it HST} observations could detect the lens, and even if this failed, the observations would strongly constrain the host to be of very low mass. \begin{figure} \includegraphics[width=\columnwidth]{fig7.eps} \caption{ Log-log plot of planet-star mass ratio $q$ versus separation (normalized to $\theta_{\rm E}$) $s$ for 44 previously published or submitted planets and OGLE-2016-BLG-0596Lb (green pentagon). Planets are colored by path to detection: detected and characterized by followup observations (blue), detected by survey but characterized by followup (magenta), and detected and characterized by surveys (green). Their shapes indicate the principal caustic feature giving rise the anomaly: planetary (circles), central (squares), and resonant (triangles). Planets suffering from the close/wide degeneracy are shown by two open symbols, whereas those for which this degeneracy is resolved are shown by a single solid symbol. By this system, OGLE-2016-BLG-0596Lb should be a solid green triangle. } \label{fig:seven} \end{figure} \section{Discussion} OGLE-2016-BLG-0596Lb is a very high mass-ratio $(q=0.0117)$ planet that lies projected very close to the Einstein ring $(s=1.075)$, which consequently generated a huge resonant caustic that required 16 days for the source to traverse. The underlying event was of quite high magnification $(A_{\rm point-lens}\sim 100)$, which led to pronounced features at peak. It therefore would seem to be extremely easy to discover. While the data set posted on the OGLE web site are adversely affected by the nearby variable, it is still the case that a free fit to these data leads to a solution qualitatively similar to the one presented here (except that it lacks a measurement of $\rho$). It is therefore striking that none of the automated programs nor active individual investigators that query this site noticed this event (or at least they did not alert the community to what they found as they do for a wide range of other events, many less interesting). This indicate the possibility that there may be many other planets ``hidden in plain sight'' in existing data. This is also supported by the planet discoveries MOA-2008-BLG-397Lb \citep{Suzuki2014}, OGLE-2008-BLG-355Lb \citep{Koshimoto2014}, and MOA-2010-BLG-353Lb \citep{Rattenbury2015}, for which the planetary signals were not noticed during the progress of events. These three characteristics, high-magnification (which is usually associated with survey+followup rather than survey-only mode), very high mass ratio, and apparent failure of both machine and by-eye recognition of the planetary perturbation, prompt us to address two questions. First, how do the real (as opposed to theoretical) planet sensitivities differ between survey-only and survey+followup modes. Second, why was this planet discovered only based on systematic analysis and what does this imply about the need for such systematic analysis of all events? \subsection{Summary of Microlens Planet Detections in the Observational $(s,q)$ Plane} Many papers contain figures that summarize microlensing planet detections in the physical plane of planet-mass versus projected separation (with the latter sometimes normalized by the snow line), e.g., Figure 1 of \citet{mb13605}. And there are many studies that show plots of {\it planet sensitivity} in the observational $(s,q)$ plane, (e.g., \citealt{gaudi02,gould10}). But to our knowledge, there are no published figures (or even figures shown at conferences) showing the census of microlensing planet discoveries on this plane. \begin{figure} \includegraphics[width=\columnwidth]{fig8.eps} \caption{ Cumulative microlensing planet detections by log mass ratio $\log(q)$, with top normalized and bottom unnormalized. Green shows the 18 planets that were detected and characterized by surveys, while magenta show the 27 planets that required significant followup observations for detection and/or characterization. Black is total. The green and magenta curves are not statistically distinguishable. } \label{fig:eight} \end{figure} Figure~\ref{fig:seven} illustrates the position of OGLE-2016-BLG-0596 (green pentagon) among the 44 previously published planets (or, to the extent we have such knowledge, submitted for publication). Discovered bodies are defined to be ``planets'' if their measured or best-estimated mass $m_p<13\,M_{\rm jup}$ and if they are known to orbit a more massive body\footnote{To facilitate comparison with future compilations, we list here the 45 planets used to construct this figure and those that follow. We compress, e.g., OGLE-2003-BLG-235Lb to OB03235 for compactness and only use ``b,c'' for multiple planets: OB03235, OB05071, OB05169, OB05390, MB06bin1, OB06109b, OB06109c, MB07192, MB07400, OB07349, OB07368, MB08310, MB08379, OB08092, OB08355, MB09266, MB09319, MB09387, MB10073, MB10328, MB10353, MB10477, MB11028, MB11262, MB11293, MB11322, OB110251, OB110265, OB120026b, OB120026c, OB120358, OB120406, OB120455, OB120563, OB120724, MB13220, MB13605, OB130102, OB130341, OB140124, OB141760, OB150051, OB150954, OB150966, OB160596.}. Planets are color-coded by discovery method: discovered by followup observations (blue), discovered (or discoverable) in survey-only observations but requiring followup for full-characterization (magenta), fully (or essentially fully) characterized by survey observations (green). The shapes of the symbols indicate the type of caustic that gave rise to the planetary perturbation: circles, squares, and triangles for planetary, central, and resonant caustics, respectively. In many cases, solutions with $(q,s)$ and $(q,1/s)$ yield almost equally good fits to the data \citep{griest98}. In these cases, the two solutions are shown as open symbols in order to diminish their individual visual ``weight'' relative to the filled symbols used when this degeneracy is broken. Hence, OGLE-2016-BLG-0596Lb would be a green filled triangle if it were not being singled out by making it larger pentagon. \begin{figure} \includegraphics[width=\columnwidth]{fig9.eps} \caption{ Cumulative microlensing planet detections by absolute value of the log projected separation (normalized to $\theta_{\rm E}$) $|\log(s)|$, with top normalized and bottom unnormalized. Colors are the same as in Figure~\ref{fig:eight}. The gap between the green (survey) and magenta (followup) curves has a 8.5\% probability of being random. If real, this indicates that followup observation have been relatively more sensitive to planets near the Einstein ring while surveys are more sensitive to those further from the Einstein ring. } \label{fig:nine} \end{figure} The most striking feature of this figure is that, in sharp contrast to the triangular appearance of high-magnification-event planet-sensitivity plots (e.g., \citealt{gould10}) and to ``double pronged'' low-magnification sensitivity plots (e.g., \citealt{gaudi02}), this detection plot looks basically like a cross, with a vertical band of detections near $\log s\sim 0$ and a horizontal band near $\log(q)\sim -2.5$. The part of this structure at high mass ratio $\log(q)>-2$ is easily explained: companions with high mass ratio are, a priori, most likely stars or brown dwarfs (BDs) and can only be claimed as ``planets'' if the host mass is known to be low. This in turn usually requires a measurement of the microlens parallax, which for ground based observations is much more likely if there is a large caustic and so $s\sim 1$. We note that there are 4 planet detections in the region $(\log(s)>+0.15, \log(q)<-3)$, while there is no detection in the opposite quadrant $(\log(s)<-0.15, \log(q)<-3)$. All the 4 planets derive from planetary caustics and 3 of them are pure survey detections: MOA-2011-BLG-028Lb\footnote{We note that this event's light curve does contain some followup data, but it is not essential for characterizing the planet.}, OGLE-2008-BLG-092Lb, and MOA-2013-BLG-605Lb \citep{mb11028,ob08092,mb13605}. The remaining planet, OGLE-2005-BLG-390Lb \citep{ob05390}, dates from an era when followup groups intensively monitored the wings of events, primarily due to the paucity of better targets. Thus we may expect that surveys will gradually fill in this quadrant. The difference in the detection rates between the quadrants with $\log(s)>+0.15$ and $\log(s)<-0.15$ can be explained by the difference in the size of the planetary caustics with $s<1$ and $s>1$. In the case of $s>1$, there exist a single planetary caustic. In the case of $s<1$, on the other hand, there exist two sets of planetary caustics and each one is smaller than the planetary caustic with $s>1$. As a result, the planetary caustic with $s>1$ has a larger cross section and thus higher sensitivity. Furthermore, smaller caustic size of planets with $s<1$ makes planetary signals tend to be heavily affected by finite-source effects, which diminish planetary signals, while signals of planets with $s>1$ can survive and show up in the wings of light curves. Actually, all 4 events with planet detections via planetary-caustic perturbations are involved with large source stars, i.e.\ giant and subgiant stars for which finite-source effects are important. \begin{figure} \includegraphics[width=\columnwidth]{fig10.eps} \caption{ Cumulative microlensing planet detections by year of discovery, with top normalized and bottom unnormalized. Colors are the same as in Figure~\ref{fig:eight}. Followup discoveries (magenta) have dropped off dramatically since 2013 } \label{fig:ten} \end{figure} Apart from this quadrant, it is not obvious that surveys are probing a different part of parameter space from the previously dominant survey+followup mode. To further investigate this, we show in Figures~\ref{fig:eight} and \ref{fig:nine} the cumulative distributions of planets by log mass ratio $\log(q)$ and (absolute value of) log separation $|\log(s)|$. In this case we distinguish only between events that could be fully characterized by survey observations (green) and those that required significant followup (including auto-followup by surveys). These distributions generally appear quite similar. For the mass ratio distribution, the greatest difference (0.259) is at $\log(q)=-2.319$, which is very typical (Kolmogorov-Smirnoff (KS) probability 40\%). The greatest difference for the separation distribution (0.334 at $|\log(s)|=0.124$) has a KS probability of 8.5\%. This may be indicative of a real difference. If so, the difference would be that pure-survey is relatively more efficient at finding widely separated lenses, which was already hinted at by inspection of the $(q,s)$ scatter plot. Finally, in Figure~\ref{fig:ten}, we show cumulative distributions by year of discovery. One might expect that with the massive ramp-up of surveys, survey-only discoveries would move strongly ahead of survey+followup. This expectation is confirmed in its sign but not its magnitude by Figure~\ref{fig:ten}. It shows that in (2014, 2015, 2016) there have been (2,2,1) and (0,1,0) discoveries by survey-only and survey+followup, respectively. This is certainly not a complete accounting, in part because 2016 has just begun and in part because historically there has been a considerable delay in microlensing planet publications for a variety of reasons. For example, of the 28 planets discovered prior to 2012, the number with delays (publication year minus discovery year) of $(0,1,\ldots,9)$ years was $N=(1,5,9,5,1,2,4,0,0,1)$. In the history of microlensing, there has been only one planet published during the discovery year, OGLE-2005-BLG-071Lb \citep{ob05071}. Hence, we will only get a full picture of this transition after a few years. \subsection{Challenges to the By-Eye and By-Machine discovery of OGLE-2016-BLG-0596} There are three interrelated reasons why OGLE-2016-BLG-0596 may have escaped notice as a potentially planetary event until the KMTNet data for this event were examined (for reasons unrelated to any apparent anomaly). First, it is relatively faint at peak. Second, it has a variable baseline. Third, it was not announced as a microlensing event until one day after the peak. As a general rule, high-magnification events are singled out for intensive followup observations only if they are still rising. When such intensive observations would have been conducted, they would have immediately revealed the anomalous nature of the event, probably triggering additional observations. This is how many of the planets discovered by $\mu$FUN were found. While $\mu$FUN itself is now semi-dormant, its protocols are directly relevant here because what is of interest is whether there is prima facie evidence for a population of missed planets during past years, during most of which $\mu$FUN was active. Now, in fact, OGLE-2016-BLG-0596 met the criteria for an OGLE alert 24 hours previously, but no alert was issued because of caution due to the variable baseline. Nevertheless, even if such an alert had been issued, it would not have triggered any followup observations because (due to the anomaly) the event would have appeared to have already peaked at that time. Finally, the variability of the baseline may have influenced modelers and followup groups to discount the evident irregularities in the light curve near peak as being due to data artifacts. This could have been exacerbated by the faintness of the event, which increases both the formal error bars and the probability of centroiding errors (hence irregular photometry) due to bright blends. Both of these effects reduce the confidence of modelers that apparent anomalies in online ``quick look'' photometry are due to physical effects. It is nevertheless a fact that when the original OGLE data are modeled, they show a clear signal for a massive planet or low-mass BD, which would trigger a re-reduction of the data, such as the one we report here. We therefore conclude that while OGLE-2016-BLG-0596 has some near-unique features that increased the difficulty of recognizing it as a planetary event, such recognition was clearly feasible. Hence, we do indeed regard this event as prima facie evidence for more such events in archival data, particularly OGLE-IV data 2010-2015. \acknowledgments The OGLE project has received funding from the National Science Centre, Poland, grant MAESTRO 2014/14/A/ST9/00121 to AU. Work by C.H.\ was supported by Creative Research Initiative Program (2009-0081561) of National Research Foundation of Korea. The OGLE Team thanks Profs. M.~Kubiak and G.~Pietrzy{\'n}ski, former members of the OGLE team, for their contribution to the collection of the OGLE photometric data over the past years. WZ and AG were supported by NSF grant AST-1516842. Work by JCY was performed under contract with the California Institute of Technology (Caltech)/Jet Propulsion Laboratory (JPL) funded by NASA through the Sagan Fellowship Program executed by the NASA Exoplanet Science Institute. This research has made the telescopes of KMTNet operated by the Korea Astronomy and Space Science Institute (KASI).
cba9a11ed04f8b223c371d569d015737d5c45e1f
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Supplementary materials} Here we elaborate on the symmetries of the studied models, derive the effective Hamiltonian, as well as motivate the chosen disordered interaction strength. \subsection{Symmetries} Even with spatially dependent interactions the Hamiltonian, \begin{equation} \hat{H}=-t\sum_{i\sigma}\left(\hat{c}_{i\sigma}^{\dagger}\hat{c}_{i+1,\sigma}+\hat{c}_{i+1,\sigma}^{\dagger}\hat{c}_{i,\sigma}\right)+\sum_{i}U_{i}\hat{n}_{\uparrow i}\hat{n}_{\downarrow i},\label{eq:nonsymm} \end{equation} has a $SU\left(2\right)$ symmetry with respect to the rotation of the spin, with the following generators, \begin{eqnarray} \hat{S}^{z} & = & \frac{1}{2}\sum_{i}\left(\hat{n}_{i\uparrow}-\hat{n}_{i\downarrow}\right)\nonumber \\ \hat{S}^{+} & = & \sum_{i}\hat{c}_{i\uparrow}^{\dagger}\hat{c}_{i\downarrow}\nonumber \\ \hat{S}^{-} & = & \sum_{i}\hat{c}_{i\downarrow}^{\dagger}\hat{c}_{i\uparrow}. \end{eqnarray} The modified Hamiltonian, \begin{eqnarray} \hat{H}' & = & -t\sum_{i\sigma}\left(\hat{c}_{i\sigma}^{\dagger}\hat{c}_{i+1,\sigma}+\hat{c}_{i+1,\sigma}^{\dagger}\hat{c}_{i,\sigma}\right)+\label{eq:symmetric}\\ & + & \sum_{i}U_{i}\left(\hat{n}_{\uparrow i}-\frac{1}{2}\right)\left(\hat{n}_{\downarrow i}-\frac{1}{2}\right),\nonumber \end{eqnarray} has an additional $SU\left(2\right)$ symmetry in the charge sector, which follows particle-hole symmetry, \begin{equation} \hat{c}_{i\sigma}\to\left(-1\right)^{i}\hat{c}_{i\sigma}^{\dagger}. \end{equation} Its symmetry group is therefore $SU\left(2\right)\times SU\left(2\right)$. An additional symmetry which exists for the (\ref{eq:symmetric}) Hamiltonian in the $\left(N=L,S_{z}=0\right)$ sector is that the eigenvalues satisfy, $E\longleftrightarrow-E$. \subsection{Selection of the disordered interaction strength} To delineate the phase diagram of the system we study the properties of its eigenstates using exact diagonalization. For this purpose we calculate the entanglement entropy of all the eigenstates. We first obtain the reduced density matrix, \begin{equation} \rho_{A}\left(E\right)=\text{Tr }_{\bar{A}}\left|\psi\left(E\right)\right\rangle \left\langle \psi\left(E\right)\right|, \end{equation} for every eigenstate $\left|\psi\left(E\right)\right\rangle $, where $A$ is a subsystem (we chose it to be the left half of the system) and $\bar{A}$ is its complimentary. The entanglement entropy is then defined as, \begin{equation} S\left(E\right)=-\text{Tr }\rho_{A}\left(E\right)\log_{2}\rho_{A}\left(E\right). \end{equation} We will designate by $s\left(\varepsilon\right)=S\left(\varepsilon\right)/L$ the entanglement entropy density, and take $\epsilon\equiv\left(E-E_{\min}\right)/\left(E_{\text{max}}-E_{\min}\right)-0.5$, to be the renormalized unitless energy density, which lies in the interval $\varepsilon\in\left[-0.5,0.5\right]$. For one-dimensional systems which we consider in this work, $s\left(\varepsilon\right)$ is expected to be independent of system size if the system is ergodic, and to be inversely proportional to system size for a non-ergodic system. The location of the transition could be therefore obtained by the calculation of $s\left(\varepsilon\right)$ for various disorder strengths and energy densities. This procedure is presented in Fig.~\ref{fig:ee}. On the left panel we obtain the location of the mobility edge $\Delta U_{c}$, for a fixed energy density, $\epsilon$. This is inferred from the intersection point of plots of $s\left(\varepsilon\right)$ for various system sizes. Each $s\left(\varepsilon\right)$ is obtained by calculating $S\left(\varepsilon\right)$ and averaging over a small interval of energy densities $\delta\varepsilon=0.04$, as also $1000$ disorder realization. In the left panel of Fig.~\ref{fig:ee} we demonstrate this procedure for the model (Eq.~\ref{eq:ham_non_sym}) for $\varepsilon=-0.3$. Due to the fast growth of the Hilbert space $\left(4^{L}\right)$ we have access to only two system sizes, $L=6$ and $8$ in the sector $N=L,\,S_{z}=0$. After repeating the outlined procedure for a number of energy densities we obtained the mobility edge, $\Delta U_{c}\left(\varepsilon\right)$, as a function of the energy density for both symmetric and non-symmetric versions of the model. As one can see from the right panel of Fig.~\ref{fig:ee}, it has a typical domed shape with states in the middle of the band less localized than the states with lower (higher) energy densities. \begin{figure} \begin{centering} \includegraphics[width=85mm]{s1_hubbard_entanglement_entropy} \par\end{centering} \caption{\label{fig:ee}\emph{Left panel}: entanglement entropy density $\left(s\left(\varepsilon\right)\right)$ is plotted for various values of interaction disorder for a fixed energy density, $\varepsilon=-0.3$, and the $SU\left(2\right)$ symmetric version of model, (\ref{eq:nonsymm}). \emph{Right panel}: the critical interaction disorder (mobility edge) is plotted versus the energy density $\varepsilon$ for both for models (\ref{eq:nonsymm}) {[}black{]} and (\ref{eq:symmetric}) {[}red{]}.} \end{figure} Since in this work we are interested in demonstrating localization we choose $\Delta U=30$, such that \emph{all} many-state are expected to be localized for \emph{both} versions of the model. \subsection{Derivation of the Effective Hamiltonian} We write the Hamiltonian as \begin{equation} \hat{H}=\hat{H}_{K}+\hat{H}_{U}, \end{equation} where $\hat{H}_{U}$ is the interaction term, \begin{equation} \hat{H}_{U}=\sum_{i}U_{i}\left(\hat{n}_{\uparrow i}-\frac{1}{2}\right)\left(\hat{n}_{\downarrow i}-\frac{1}{2}\right) \end{equation} and \begin{equation} \hat{H}_{K}\equiv\sum_{i\sigma}t_{ij}\hat{c}_{i\sigma}^{\dagger}\hat{c}_{j\sigma}, \end{equation} is the perturbation. For $U_{i}<0$ the ground state of $\hat{H}_{U}$ contains only doublons and is $2^{L}$ times degenerate, since a doublon and a holon have same energies. We define a projector on the zero and doubly occupied states as, $\hat{P}$ and its complimentary as, $\hat{Q}=1-\hat{P}$. Then the Hamiltonian can be written in a block form, \begin{equation} \hat{H}=\hat{P}\hat{H}\hat{P}+\hat{Q}\hat{H}\hat{Q}+\left(\hat{P}\hat{H}\hat{Q}+\hat{Q}\hat{H}\hat{P}\right). \end{equation} We further divide the Hamiltonian into unperturbed, $\hat{H}_{0}=\hat{P}\hat{H}\hat{P}+\hat{Q}\hat{H}\hat{Q}$ and perturbed parts, $\hat{H}_{1}=\hat{P}\hat{H}\hat{Q}+\hat{Q}\hat{H}\hat{P}$, and note that $\left[\hat{P},\hat{H}_{0}\right]=0$ but $\left[\hat{P},\hat{H}_{1}\right]\neq0$. By performing a unitary transformation we eliminate the off-diagonal term up to second order in $\hat{H}_{1}$. Using Baker-Cambell-Hausdorff formula we can write, \begin{eqnarray} e^{S}\hat{H}e^{-S} & = & \hat{H}_{0}+\left[S,\hat{H}_{0}\right]+\frac{1}{2!}\left[S,\left[S,\hat{H}_{0}\right]\right]+\cdots\\ & + & \hat{H}_{1}+\left[S,\hat{H}_{1}\right]+\frac{1}{2!}\left[S,\left[S,\hat{H}_{1}\right]\right]+\cdots.\nonumber \end{eqnarray} Setting, \begin{equation} \left[S,\hat{H}_{0}\right]=-\hat{H}_{1},\label{eq:sw_req} \end{equation} we can recast the series into the form, \begin{eqnarray} e^{S}\hat{H}e^{-S} & = & \hat{H}_{0}+\left(\frac{1}{1!}-\frac{1}{2!}\right)\left[S,\hat{H}_{1}\right]\\ & + & \left(\frac{1}{2!}-\frac{1}{3!}\right)\left[S,\left[S,\hat{H}_{1}\right]\right]+\cdots.\nonumber \end{eqnarray} The effective Hamiltonian to second order in the off-diagonal terms is therefore given by, \begin{equation} \hat{H}_{\text{eff}}=e^{S}\hat{H}e^{-S}=\hat{H}_{0}+\frac{1}{2}\left[S,\hat{H}_{1}\right]. \end{equation} To calculate $S$ we use (\ref{eq:sw_req}) in the unperturbed basis, \begin{equation} \left\langle \alpha\left|\left[S,\hat{H}_{0}\right]\right|\beta\right\rangle =S_{\alpha\beta}\left(E_{\beta}^{\left(0\right)}-E_{\alpha}^{\left(0\right)}\right)=-\left\langle \alpha\left|\hat{H}_{1}\right|\beta\right\rangle , \end{equation} and therefore, \begin{equation} S_{\alpha\beta}=\frac{\left\langle \alpha\left|\hat{P}\hat{H}\hat{Q}+\hat{Q}\hat{H}\hat{P}\right|\beta\right\rangle }{E_{\alpha}^{\left(0\right)}-E_{\beta}^{\left(0\right)}}. \end{equation} In case that $\hat{P}$ projects to a degenerate subspace, such that $\hat{H}_{0}\hat{P}\left|\alpha\right\rangle =E_{P}^{\left(0\right)}\hat{P}\left|\alpha\right\rangle $, we can write, \begin{equation} \hat{S}=\hat{P}\hat{H}\left(E_{P}^{\left(0\right)}-\hat{H}_{0}\right)^{-1}\hat{Q}-\hat{Q}\left(E_{P}^{\left(0\right)}-\hat{H}_{0}\right)^{-1}\hat{H}\hat{P}. \end{equation} After some algebra we obtain the effective Hamiltonian, \begin{equation} \hat{H}_{\text{eff}}=E_{P}^{\left(0\right)}+\sum_{ijkl,\sigma\sigma'}\frac{t_{ij}t_{kl}}{U_{ij}}\hat{P}\hat{c}_{k\sigma'}^{\dagger}\hat{c}_{l\sigma'}\hat{c}_{i\sigma}^{\dagger}\hat{c}_{j\sigma}\hat{P}, \end{equation} where \begin{equation} U_{ij}\equiv\frac{U_{i}+U_{j}}{2}. \end{equation} Finally defining the following pseudo-spin operators, \begin{equation} \hat{S}_{i}^{+}=\hat{c}_{i\uparrow}^{\dagger}\hat{c}_{i\downarrow}^{\dagger}\qquad\hat{S}_{i}^{-}=\hat{c}_{i\downarrow}\hat{c}_{i\uparrow},\qquad\hat{S}_{i}^{z}=\frac{1}{2}\left(\hat{n}_{i\downarrow}+\hat{n}_{i\uparrow}-1\right), \end{equation} and using the properties of the projectors we obtain the anti-ferromagnetic random Heisenberg model, \begin{equation} \hat{H}_{\text{eff}}=E_{P}^{\left(0\right)}+\sum_{ij}\frac{2\left|t_{ij}\right|^{2}}{\left|U_{ij}\right|}\left(\hat{\mathbf{S}}_{i}\cdot\hat{\mathbf{S}}_{j}-\frac{1}{4}\right). \end{equation} \subsection{Finite time and finite size effects} As in any numerical study, our results are limited to finite sizes and times. It is therefore pertinent to present evidence that the observed behavior persists also for larger systems and longer times. We would like to consider three possible objections to the observed localization, all related to the relatively high $U$ we use in our simulation: \begin{enumerate} \item For small systems and large interaction, $U/t_{h}\gg1$, the many-body spectrum includes gaps, which effectively makes the system non-ergodic \cite{Papic2015}. \item The system size is too small to either break apart the localized doublons, or to build a large, almost ``classical'' quasi-particle along the scenario of delocalization introduced in Ref.~\cite{Potter2016}. \item The time it takes to observe delocalization is beyond the reach of our simulations. \end{enumerate} \begin{figure} \begin{centering} \includegraphics{s2_DoS-L=7} \par\end{centering} \caption{\label{fig:spectrum}Eigenvalues of a single disorder realization normalized to lie in the interval $\left[0,1\right]$. \emph{Left panel}: $SU\left(2\right)$ symmetric model, (\ref{eq:nonsymm}), for $\Delta U=3$ (light orange) and $\Delta U=30$ (dark orange). \emph{Central panel}: clean system, $\Delta U=0$. \emph{Right panel}: same as left panel but for the $SU\left(2\right)\times SU\left(2\right)$ symmetric model (\ref{eq:symmetric}). For all models $L=7$. } \end{figure} We argue that the existence of gaps in the many-body spectrum does not \emph{necessarily} imply localization, moreover some of the gaps may very well persist in the thermodynamic limit \cite{Garrison2016}. In Fig.~\ref{fig:spectrum} we present the many-body eigenvalues of models (\ref{eq:nonsymm}) and (\ref{eq:symmetric}) for a single disorder realization, as well as the eigenvalues of the corresponding clean model $\left(\Delta U=0\right).$ Gaps exist for all three models. In fact, since for the disordered models we use $-\Delta U\leq U_{i}\leq0$, gaps are \emph{smaller} for larger $\Delta U$, since disorder enlarges the phase space of possible $\left(U_{1}+U_{2}+\cdots\right)$ combinations. Therefore the largest gaps appear for $\Delta U=0$, when only multiples of $U$ are possible. The transport, on the other hand, is very different in all three models. It varies from localized, delocalized to subdiffusive, as we present in the main text. This shows that having gaps in the many-body spectrum need not have direct implications for ergodicity. \begin{figure} \begin{centering} \includegraphics[width=86mm]{s3_su2-no-ones} \par\end{centering} \caption{\label{fig:su2-no-ones}Charge excitation dynamics for starting from a random charge configuration \emph{without} singlons calculated using tDMRG simulation for $L=20$ (orange) and ED for $L=7$ (black/gray). \emph{Left panel}: Excitation profile at the final times of the simulation $L=7,$ $t=10^{10}$ and $L=20,$ $t=30$. \emph{Right panel}: charge excitation width as a function of time on a log-log plot. Shades designate uncertainty due to averaging. The parameters used are $-30\leq U_{i}\leq0$. } \end{figure} The second point concerns the finite size of the system. For small systems, some excitations might never decay. For example, due to energy conservation, a doublon with an energy of $U$ would \emph{never} decay in a system of $N$ particles, if the maximal change of the single particles is $\delta_{\max}$, such that $U>\delta_{max}N$. For larger systems a collective (though very rare) excitation of the particles would be able to break the doublon apart. However since this process requires a delicate cooperation of the particles the decay rate of the doublon would be exponentially slow \cite{Sensarma2010}. Since in the model we study (\ref{eq:ham_non_sym}) a finite doublon density is necessary to localize the singlons, one may question: \textbf{a)} if localization might be lost due to doublon decay \textbf{b)} the small system size prevents the decay to occur. We argue that a \emph{finite} doublon density will never decay to zero for sufficiently strong interaction even for \emph{infinite} systems. This means that the mechanism of localization due to doublons may exist also in the thermodynamic limit, though with (slightly) renormalized disorder strength. To show that finite doublon density cannot decay, we repeat the above argument for a finite system with a minor modification. By duplicating the small system, and using energy conservation, it is easy to see that the maximal change in the doublon density is $\delta n_{d}=O\left(1/U\right),$ which means that for sufficiently large $U$ a \emph{finite} density of doublons cannot decay completely even in an \emph{infinite} system. Moreover a finite doublon density implies that the temperature is $T\approx U$, for which the processes required for the decay of the doublons are not rare anymore, and occur at the rate of $t_{h}.$ The change of the effective disordered potential as a result of the doublon decay is about $O\left(1/U\right)$. We thus claim that the effect of the decay of the doublon density on localization is negligible for large $U$. In Ref.~\cite{Potter2016} it is suggested that for $SU\left(2\right)$ symmetric models the local degrees of freedom will join into large ``quasi-classical'' objects which will induce delocalization. One might therefore suspect that either the system is too small to create such a large object, or alternatively the creation and motion of such objects might occur at later times. This study however does not specify what are the required length and time scales to observe delocalization in such models. \begin{figure} \begin{centering} \includegraphics[width=86mm]{s4_su2-all} \par\end{centering} \caption{\label{fig:su2-all}Charge excitation dynamics for starting from a random charge configuration \emph{(no constraints)} calculated using tDMRG simulation for $L=20$ and $50$ (light and dark orange) and ED for $L=9$ (black/gray). \emph{Left panel}: Excitation profile at the final times of the simulation $L=9,$ $t=100$ and $L=20,50$, $t=30$. \emph{Right panel}: charge excitation width as a function of time on a log-log plot. Shades designate uncertainty due to averaging. The parameters used are $-30\leq U_{i}\leq0$. } \end{figure} In our work we calculate the spread of a \emph{local} excitation which is described by the correlation function, \begin{equation} C_{i}\left(t\right)=\frac{1}{\text{Tr }\hat{P}}\text{Tr }\hat{P}\left(\hat{n}_{i}\left(t\right)-1\right)\left(\hat{n}_{0}-1\right). \end{equation} The locality of the operators which appear in the correlation function, as well as the locality of the Hamiltonian (\ref{eq:ham_non_sym}), guarantee \begin{equation} \left\Vert \left\langle \hat{O}_{i}\left(t\right)\hat{O}_{j}\right\rangle \right\Vert \leq\exp\left[-a\left(\left|i-j\right|-vt\right)\right], \end{equation} \begin{figure} \begin{centering} \includegraphics[width=86mm]{s5_domain-wall} \par\end{centering} \caption{\label{fig:domain_wall} On the left, the charge density perturbation, $\left\langle \hat{n}_{i}\left(t\right)-\hat{n}_{i}\left(0\right)\right\rangle $, is plotted vs time and space for models (\ref{eq:nonsymm}) {[}top{]} and (\ref{eq:symmetric}) {[}bottom{]}. On the right the total charge on the LHS, $M_{L}\left(t\right),$ is plotted versus time on the top panel and the entanglement entropy, $S\left(t\right)$ versus time is plotted on the bottom. The data has been obtained using tDMRG for $L=30$ and $\Delta U=30$, and averaged over at least 300 disorder realizations.} \end{figure} following from rigorous Lieb-Robinson bounds \cite{Lieb1972}, where $v$ is the maximal spread velocity, which is of order of the hopping rate (here $O\left(1\right)$). This means that for any \emph{finite} time, the excitation is bounded by an exponential in the distance $\left|i-j\right|$, namely it is effectively contained in a finite box. \emph{Finite size} effects could be therefore eliminated to any required precision \emph{up to any finite time}. For localized systems the bound is uniform in time, which signifies the absence of transport \cite{Friesdorf2015}. In our work, by looking at the profile of the excitation as well as its width, we show that for the $SU\left(2\right)$ symmetric model (\ref{eq:nonsymm}) this bound is indeed uniform at least up to some finite time. As can be seen from the left panel of Fig.~\ref{fig:su2-no-ones}, starting from an initial configuration of only doublons, the $L=7$ system and the $L=20$ system essentially give the same results, while compared at strikingly different times $t=30$ $\left(L=20\right)$ versus $t=10^{10}$ $\left(L=7\right)$. This signifies both absence of charge transport, and effective elimination of finite size effects. For \emph{unconstrained} random initial configuration, as can be seen from Fig.~\ref{fig:su2-all}, non-negligible residual tails of the excitation prevent elimination of finite size effects for a system size of $L=9$ (ED). Nevertheless, the excitation profile stays almost the same across increasing system sizes of $L=9,$ $20$ and $50$, and the width of the excitation appears to be converged already for $L=20$. The tDMRG simulation demands considerable resources in this regime due to the initial fast growth of the entanglement entropy, which results from ballistic motion of the singlons between the doublons. This results in poor averaging (100-300 realizations) and a bias towards more localized charge configurations. Notwithstanding, the appearance of localization plateau, while less convincing compared to ``doublons only'' case, is evident here as well. \subsection{Domain wall initial condition} We have also considered the dynamics starting from a domain wall initial conditions. This type of initial conditions is more amenable to experimental study in cold atoms \cite{Choi2016}. We have chosen an initial condition such that the right half of the system is completely full, while the left half is empty, and measured the total number of particles on the LHS, $M_{L}\left(t\right).$ In Fig.~\ref{fig:domain_wall} one can see the profile on the charge density perturbation,$\left\langle \hat{n}_{i}\left(t\right)-\hat{n}_{i}\left(0\right)\right\rangle ,$ on the left, as also the charge transfer, $M_{L}\left(t\right)$ and $S\left(t\right)$ on the right. For the $SU\left(2\right)$ symmetric model described by (\ref{eq:nonsymm}) the ballistic jet of the singlons is clearly visible, as is its reflections from the simulation boundaries as well. The core of the system however remains localized. The $SU\left(2\right)\times SU\left(2\right)$ symmetric model described by (\ref{eq:symmetric}) is clearly delocalized and on the accessible time scales exhibits transport very close to diffusive, \begin{equation} M_{L}\left(t\right)\propto t^{0.6}, \end{equation} with entanglement entropy which is almost linearly growing with time. \end{document}
aece2e1b113816758ed870a9ab9952e49b9175d8
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sxn:intro} Graphs, long popular in computer science and discrete mathematics, have received renewed interest recently in statistics, machine learning, data analysis, and related areas because they provide a useful way to model many types of relational data. In this way, graphs can be used to extract insight from data arising in many application domains. In biology, e.g., graphs are routinely used to generate hypotheses for experimental validation~\cite{Tuncbag-2016-glioblastoma}; and in neurscience, they are used to study the networks and circuits in the brain~\cite{Bassett-2006-small-world,Sporns-2002-network-analysis}. Given their ubiquity, graphs and graph-based data have been approached from several different perspectives. In computer science, it is common to develop algorithms, e.g., for connected component, minimum spanning tree, and maximum flow problems, to run on data that are modeled as a precisely-specified input graph. These algorithms are then characterized by the number of operations they use for the worst-possible input at any size. In statistics and machine learning, on the other hand, it is common to use graphs as models to perform inference about unseen data. In this case, one often hypothesizes an unseen graph with a particular structure, such as block structure, hierarchical structure, low-rank structure in the adjacency matrix, etc. Then one runs algorithms on the observed data in order to impute entries in the unobserved hypothesized graph. These methods may be characterized in terms of running time, but they are also characterized in terms of the amount of data needed to recover the hidden hypothesized structure. In many application areas where the end goal is to obtain some sort of domain-specific insight, e.g., such as in social network analysis, neuroscience, medical imaging, etc., one constructs graphs from primary data, and then one runs a computational procedure that does \emph{not} come with either of these traditional types of theoretical guarantees. As an example, consider the GeneRank method~\cite{morrison2005-generank}, where we have a set of genes related to an experimental condition in a microarray study. This set of genes is ``refined'' via a locally-biased graph algorithm closely related to those we will discuss. Importantly, this operational refinement procedure does \emph{not} come with the sort of theory traditional in statistics, machine learning, or computer science. As anther example, e.g., in social network applications, one might run a random walk process for a few steps from a starting node of interest, and if the walk ``gets stuck'' then one might interpret this as evidence that that region of the graph is meaningful in the application domain~\cite{EH09_TR}. These are examples of the types of heuristics commonly-used in applications. By heuristic, we mean an algorithm in the sense that it performs a sequence of well-defined steps, but one where precise theory is lacking (although usually heuristics come with strong intuitive motivation and are justified in terms of some downstream application). In particular, typically heuristics do not explicitly optimize a well-defined objective function and typically they do not come with well-defined inferential guarantees. Note that, in both of these examples, the goal is to find ``local'' or ``small scale'' structure in the data graph. Both examples also correspond to what practitioners interested in downstream applications actually do. Existing algorithmic and statistical theory, however, has challenges with these local or small-scale structures. For instance, a very ``good'' algorithmic runtime on a graph is traditionally one that is \emph{linear} in the number of vertices and edges. If the output of interest is only a vanishingly small fraction of a large graph, however, then this theory may not provide strong qualitative guidance on how these locally-biased methods behave in practice. Likewise, inferential methods often assume that the structures inferred constitute a substantial fraction of the graph, and many statistical techniques have challenges differentiating very small structure from random noise. In this overview, we describe a class of graph algorithms that has proven to be very useful for identifying and interpreting small-scale local structure in large-scale data. For this class of algorithms, however, strong algorithmic and statistical theory has been developed. In particular, these graph algorithms are \emph{locally-biased} in one of several precisely-quantified senses. We will describe what we mean by this in more detail below, but, informally, this means that the algorithms are most interested in only a small part of a large graph. As opposed to heuristic operational procedures, however, many of these algorithms do come with strong worst-case algorithmic guarantees, and many of these algorithms also do come with statistical guarantees that prove they have \emph{implicit} regularization properties. This complementary algorithmic-statistical theory helps explain their improved performance in many practical applications. While the approach of locally-biased graph algorithms is very general, it has been developed most extensively for the fundamental problem of finding locally-biased graph partitions, i.e., clusters or communities, and so we will focus on locally-biased approaches for the graph clustering problem. Of course, this partitioning question is of interest in many more application-driven areas, where one is interested in finding useful or meaningful clusters as part of a data analysis pipeline. \subsection{The rationale for local analysis in real-world data} As a quick example of why local graph analysis is frequently used in data and data science applications, we present in Figure~\ref{fig:small-cluster} the results of finding the best partition of both a random geometric graph and a more typical data graph. Standard graph partitioning algorithms must operate on, or ``touch'', each vertex and edge of the graph to identify these partitions. The best partition of the geometric graph is around half the data, where it is reasonable to run an algorithm that touches all the data. On the other hand, the best partition of the data graph is very small, and in this case touching the entire graph to find it can be too costly in terms of computation time. The local graph clustering techniques discussed in this paper can find this cluster touching only edges and nodes in the output cluster, greatly reducing the computation~time. Far from being a pathology or a peculiarity, the finding that optimal partitions of real-world networks are often extremely imbalanced, thus leading to very small optimal clusters, is endemic to many of the graphs arising in large-scale data analysis~\cite{LLDM08_communities_CONF,LLDM09_communities_IM,LLM10_communities_CONF,Jeub15}.\footnote{An important applied question has to do with the meaningfulness, usefulness, etc., of such small clusters. We do not consider those questions here, and instead we refer the interested reader to prior work~\cite{LLDM08_communities_CONF,LLDM09_communities_IM,LLM10_communities_CONF,Jeub15}. Here, we instead focus on the algorithmic and statistical properties of these locally-biased algorithms.} Let us now explain in more detail Figure~\ref{fig:small-cluster}. Figure~\ref{fig:small-cluster-a} shows a graph with 3000 nodes that is typical of many graphs that possess a strong underlying \emph{geometry}, e.g., those used in computer graphics, computer vision, logistics planning, road network analysis, and so on. This particular graph is produced by generating 3000 random points in the plane and connecting all points within a small radius, such that the final graph is connected. The geometric graph can be nearly bisected by optimizing a measure known as conductance (we shall define this shortly), which is designed to balance partition size and quality. In Figure~\ref{fig:small-cluster-b}, we show a more typical data graph of around 10680 nodes~\cite{Moguna-2004-models}, where this particular data graph is based on the trust relationships in a PGP (Pretty Good Privacy) key-chain. Optimizing the same conductance objective function results in a tiny set, Figure \ref{fig:small-cluster-d}, and not the near-bisection as in the geometric case, Figure \ref{fig:small-cluster-c}. (We were able to use integer optimization techniques to directly solve the NP-hard problems at the cost of months of computation.) Many other examples of this general phenomenon can be found in prior work~\cite{LLDM08_communities_CONF,LLDM09_communities_IM,LLM10_communities_CONF,Jeub15}. \begin{figure}[t] \subfigure[A random geometric graph\label{fig:small-cluster-a}]{\includegraphics[width=0.5\linewidth]{U3A-full}}% \subfigure[A typical data graph\label{fig:small-cluster-b}]{\includegraphics[width=0.5\linewidth]{pgp-full}} \subfigure[The optimal conductance solution for the geometric graph bisects the graph into two large well-balanced pieces\label{fig:small-cluster-c}]{\includegraphics[scale=0.32]{U3A-bestset}} \subfigure[The optimal conductance solution for a typical data graph. (Inset. A zoomed view of the subgraph where the two unfilled notes are the border with the rest of the graph.)\label{fig:small-cluster-d}]{\raisebox{0cm}{\includegraphics[width=0.5\linewidth]{pgp-bestset}}\llap{\colorbox{white}{\includegraphics[scale=0.32]{pgp-bestset-zoom}}}} \caption{At left, a geometric graph has a pleasing and intuitive layout in the two-dimensional plane. At right, a more typical data graph has a classic \emph{hairball} layout that shows little high level structure. Due to the evident lack of global structure in the data graph, locally-biased graph algorithms are often used in these contexts. The solution of the Minimum-Conductance problem in the geometric graph is a large set of nodes, and it has conductance value $0.00464$. The solution of the Minimum-Conductance problem in the more typical data graph is a small set of nodes, and it has conductance value $0.00589$. The inset figure shows that this small graph is very dense and has only 3 edges leaving the set.} \label{fig:small-cluster} \end{figure} \subsection{Partitioning as a model problem} The problem of finding good partitions or clusters is ubiquitous. Representative examples include biomedical applications \cite{ICOYHS01,GN02}, internet and world-wide-web \cite{FFF99,AJB99,KKRRT99}, social graphs \cite{LLDM08_communities_CONF,scott,Traud11,UKBM11}, human communication graphs \cite{OSHSLKKB07}, human mobility graphs \cite{EEBL10}, voting graphs in political science \cite{PMNW05,multiplex_Mucha,MMP12,TKMP11,GHKV07}, protein interaction graphs \cite{JPD10}, material science \cite{BODP12,RCHSSKMN11}, neuroscience \cite{BWPMCG11,WBMPG12,BWRPMG13}, collaboration graphs \cite{ELP11}. All of the features of locally-biased computations are present in this model partitioning problem. For example, while some of these algorithms read the entire graph as input but are engineered to return answers that are biased toward and meaningful for a smaller part of the input graph~\cite{MOV12_JMLR,andersen08soda,ZBLWS04,pan2004-cross-modal-discovery}, other algorithms can take as input a small ``seed set'' of nodes as well as an oracle with which to access neighbors of a node, and they return meaningful answers without even touching the entire graph~\cite{FCSRM16_TR,ACL06,KG14,ST13}. Similarly, while these algorithms are often formulated in the language of theoretical computer science as approximation algorithms, i.e., they come with running time guarantees and can be shown to \emph{approximate} to within some quality-of-approximation factor some objective function of interest, e.g., conductance, in other cases one can prove statistical results such as that they \emph{exactly} solve a regularized version of that objective~\cite{GM14_ICML,Veldt-preprint-simple-local-flow,MO11-implementing,FCSRM16_TR}. Importantly, this statistical regularization is implicit rather than explicit. Typically, regularization is explicitly used in statistics, when fitting models with a large number of parameters, in order to avoid overfitting to the given data. It is also used to select answers biased towards specific sets---for example, sparse solution sets by using the Lasso~\cite{T96}. In the case of locally-biased graph algorithms, one simply runs a \emph{faster} algorithm for a problem. In some cases, the nature of the approximation that this fast algorithm makes can be related back to a regularized variant of the objective function. \subsection{Overview} In this paper, we will consider partitioning from the perspective of conductance, and we will survey a recent class of results about a common localizing construction. These methods will let us find the optimal sets in Figure~\ref{fig:small-cluster} without resorting to integer optimization (but also losing the proof of optimality). Depending on the exact construction, they will also come with a variety of helpful statistical properties. We'll conclude with a variety of different perspectives on these problems and some open questions. In Section \ref{sec:prelim} we describe assumptions, notation and preliminary results which we will use in this paper. In Section \ref{sec:global_cond_spcut} we discuss two global graph partitioning problems and their spectral relaxation. In Section \ref{sec:localgraphpart} we describe the local graph clustering application. In Section \ref{sec:empirical} we provide empirical evaluations for the global and local graph clustering algorithms which are described in this paper. Finally, in Section \ref{sec:disc_conclusion} we give our conclusions. \section{Text that has been cut} \section{Intro} Motivated by these very large-scale data analysis applications where only small meaningful clusters exist and where the graphs are so large that iterating over all edges and nodes is measured in minutes or hours, recent work has focused on \emph{locally-biased graph algorithms}. This refers to a broad class of algorithms that are designed to take as input a large graph, e.g., a billion node social or information network, and to solve problems that are in some sense ``nearby'' an exogenously-specified ``seed set'' of nodes. While these algorithms are often very useful in practice, and while many of them have connections with heuristic procedures that practitioners often apply to their data, a distinguishing feature of them is their non-trivial algorithmic and/or statistical theory. In this overview paper, we will cover recent work of global and local graph algorithms, touching on each of these topics, reviewing complementary aspects, and focusing on optimization questions. We split this paper into three parts, optimization objectives, global and local algorithms. First we will present optimization objectives which by consensus quantify the notion of a cluster in a graph. The optimization problems that we present are NP-hard, thus cannot be optimized in polynomial time. Then we discuss state-of-the-art polynomial time relaxations that result in spectral or max-flow problems. Second, we present an overview of locally-biased graph partitioning methods in terms of \emph{flow-based methods} and \emph{spectral-based methods}. Finally, we illustrate their performance in practice and comment on the implicit or explicit regularization that is applied by each algorithm. \section{Regularization, Beyond $\ell_1$ and $\ell_2$} It is easy to see that the above problem is a continuous representation of $s$-$t$ Min-Cut problem. The constraints $x_s=1$ and $x_t=0$ represent regularization such that the trivial solution of a constant vector $x$ is obtained. At the same time these two constraints affect the separation of the graph in two parts such that the nodes $v_s$ and $v_t$ are in different parts of the partition. The approximation is based on the distortion error of embedding the semi-metric to the so called $\ell_1$-metrics set, see \cite{} (\textcolor{red}{Leighton Rao}) for details. \paragraph{Theoretical issues: combining spectral and flow.} \textcolor{red}{The material below is also covered in the SDP section, we need to combine them.} Although it isn't our main focus, we comment briefly on one set of well-known results and how it fits in with our discussion. A natural question in the above is whether one can combine the complementary ideas from spectral and flow to obtain improved results. The answer is yes, and this is the ARV algorithm. It (1) can be written as a tighter relaxation of the expansion/conductance problem, (2) can be written as an SDP, (3) provides improvement over spectral as well as LR, (4) leads to related optimization-based methods/approaches such as KRF and MWUM that lead to similar improvements. XXX. GIVE THE OBJECTIVE AND RELAXATION AND RESULTS AND LEAVE IT AT THAT. BTW, also, LMO showed that the practice follows the theory for graphs up to thousands to tens of thousands of nodes~\cite{LMO09}, but, aside from this, these methods haven't really been implemented. XXX. WE'LL PROBABLY SAY IN THE CONCLUSION THAT LOCAL ARV METHODS ARE OPEN. \subsection{Embedding aspects}\label{subsection:embeddings} \textcolor{red}{I have to embed ( :-) ) this section in previous sections since most of the material below is already mentioned above.} These two approaches have an interpretation in terms of embeddings (which is of theoretical interest, but which also can be seen empirically), and this provides a way to think of the two together as complementary (in the global case, but we will do the local case below). To start, return to min-cut, and note that that objective, while tractable, often leads to trivial solutions, in that it cuts on a very small number of nodes, e.g., one or two. Thus, it is common to normalize by the volume. (Alternatively, one could have a balance constraint or use methods that explicitly or implicitly provide a bias toward large sets of nodes.) In this case, one can define the expansion of a set of nodes to be the usual expansion-SA-to-Volume measure. XXX. DEFINE. If the graph is weighted or has variable degrees, then one could define the conductance to be the usual conductance-SA-to-Vol measure. (They are the same for unweighted/degree-regular graphs, but not otherwise and the differences are important in theory/practice.) Note that the input to these objectives is just a graph $G=(V,E,W)$. These objectives are intractable, but they capture the qualitative idea of what most people think of by clusters---relatively large, relatively a lot of stuff inside, relatively less interaction between the cluster and its complement. The point is that spectral and flow both give approximation algorithms for expansion/conductance. Spectral relaxes one way, i.e., the $\{-1,1\}$ constraint to being $\pm1$ on average, and this can be solved by computing eigenvectors or diffusions Flow relaxes another way, i.e, the $\{0,1\}$ constraint to be a metric, and this can be written as the solution of an LP, which can then be solved with flow methods like Dinic or related methods. From this embedding perspective, it is clear that spectral embeds the graph not just on a like, and it then does a sweep cut, but it also embeds the graph in a complete graph, i.e., the Laplacian of a complete graph, and it is here that the duality is the tightest. It is here that the quadratic approximation result of Cheeger arises. And note that for spectral methods random walks mix quickly on constant-degree expanders and slowly on line graphs, and this can be sued to construct the Guattery-Miller cockroach graphs. XXX. SOMETHING ABOUT CONNECTIONS WITH L2. Similarly, flow methods embed you in an $\ell_1$ space, and Bourgain's embedding lemma says that an $O(\log n)$ distortion can be expected, and this happens on constant degree expanders. We have gone through this in some detail since many of the ideas are essential for constructing local versions of these methods. Many of these results are ``obvious'' to certain research areas and not even thought about by other research areas, so we went through them in the more well-developed global partitioning area, so we can draw similarities with local methods. \subsection{Some additional observations} There is, of course, the usual theory-practice gap. Theoretical work typically considers the problem of two clusters, and applied work is often more interested in more than two clusters. The latter, while also using a lot of algorithm engineering, often uses ideas that boil down to the former. Here, we mention a few things. \paragraph{Practical issues: going beyond spectral and flow.} Global graph partitioning methods basically boil down to some combination of spectral and flow, also couples with ideas like multi-resolution. The canonical example of this is METIS~\cite{karypis98_metis,karypis98metis}. This, coupled with doing things ``right,'' e.g., doing the sweep cut right with spectral of doing something analogous like Goldberg does with flow algorithms, can have a big effect on the quality of the result. We won't go into a lot of detail about this, but we expect that a lot of these ideas can also lead to improved local and \section{Spectral methods} The material in this section is based on \cite{S11}. Then $$ vol(E(S,S^c)) = \frac{x^T L x}{(c_1-c_2)^2} $$ and $$ x^TDx = c_1^2 vol(S) + c_2^2 vol(S^c). $$ Depending on how we define $c_1$ and $c_2$ we obtain different problems. For example, if $$ c_1= \sqrt{\frac{vol(S^c)}{vol(S)}}, \ c_2= -\sqrt{\frac{vol(S)}{vol(S^c)}} $$ then the following problem \begin{align}\label{spcutLD} \mbox{minimize} & \quad \frac{x^TLx}{x^TDx} \\\nonumber \mbox{subject to} &\quad x\in\{c_1,c_2\}^{|\mathcal{V}|} \end{align} is equivalent to the Sparsest-Cut problem \eqref{spcut}. If $c_1=1$ and $c_2=0$ then \eqref{spcutLD} is equivalent to the Minimum-Isoperimetry problem \eqref{isop}. It is clear that for different values of $c_1$ and $c_2$ we impose different regularization on the optimal partition of the graph. However for any values of $c_1$ and $c_2$ problem \eqref{spcutLD} is still NP-hard. A possible way to overcome this difficulty is to relax the solution set to $\{u\in\mathbb{R}^{|\mathcal{V}|} \ | 1_{|\mathcal{V}|}^T D u = 0\}$, where $1_{|\mathcal{V}|}$ is a column vector of ones of length $|\mathcal{V}|$. This relaxation is equivalent to finding the second largest eigenvalue of the normalized graph Laplacian $\mathcal{L}$, which justifies the name spectral relaxation. The relaxed problem is \begin{align}\label{spcutLD} \mbox{minimize} & \quad \frac{x^TLx}{x^TDx} \\\nonumber \mbox{subject to} &\quad x\in\{u\in\mathbb{R}^{|\mathcal{V}|} \ | \ 1_{|\mathcal{V}|}^T D u = 0\}, \end{align} which is a non-convex problem, but all local solutions are of the same importance since all have the same objective value. The local solutions differ only by a positive constant factor and a sign change. The reason for imposing the orthogonality constraint $1_{|\mathcal{V}|}^T D x = 0$ is because otherwise \eqref{spcutLD} has a trivial solution, the eigevector $D^{1/2} 1_{|\mathcal{V}|}$ (any other constant vector can be used instead of $1_{|\mathcal{V}|}$) of the zero (the smallest) eigevalue of $\mathcal{L}$. Let us define $\delta_S(v_i,v_j) = |x_i - x_j|^2$.\ Using $\delta_S$ we can rewrite the numerator in the objective function as $$ x^T L x = \sum_{i=1}^{|\mathcal{V}|}\sum_{j=1}^{|\mathcal{V}|} C_{ij}\delta_S(v_i,v_j). $$ Another way of viewing the spectral relaxation is by noticing that if instead of defining $x : \mathcal{V} \to \{c_1,c_2\}$ to be an indicator function, we allow it to be any arbitrary function $x: \mathcal{V} \to \mathbb{R}$, then we relax the cut-metric $\delta_S$ to be the Euclidean square distance. \section{Local methods} \subsection{The Local-Isoperimetry Problem} Recall that we introduced the Sparsest-Cut and Minimum-Isoperimetry problems as regularizations of the Min-Cut problem that were designed to achieve some control over the size of the set identified. The Local-Isoperimetry problem is an integer problem that adds more regularization (or a constraint) to achieve a local solution~\cite{MOV12_JMLR}. The problem depends on a graph $G$ and a reference set $R$ that must be contained within the solution. We also make the size of the final set a parameter $k$. The new program is \begin{align} \hat{\phi}(\mathcal{G};R,k) \quad := \quad \mbox{minimize} & \; \tilde{\phi}(S)= \frac{vol(E(S,S^c))}{vol(S)} \\\nonumber \mbox{subject to} & \; S\subset \mathcal{V},\ vol(S) \le k \\ \label{eq:localiso} & \; R \subseteq S. \nonumber \end{align} It differs from the Minimum-Isoperimetry problem because we have replaced the volume bound with a constant $k$ and require that the solution set $S$ contain the reference set $R$. The idea with this problem is that we wish to localize the solution to the Minimum-Isoperimetry problem around a subset of nodes $R$. We'll often take $R$ to be a single node. Note that if $k = vol(G)$ then this problem may be solved as a parametric flow problem using techniques akin to those in~\cite{HochbaumNcutprime}. If $k = vol(G)/2$, then we are able to use $|V|$ runs of this algorithm to solve the Minimum-Isoperimetry problem by setting $R$ to be each vertex. So solving this is NP-hard in general. We illustrate one counter-intuitive feature of this problem in Figure~\ref{fig:local-iso}. First, let us note that one of the global solutions of the Minimum-Isoperimetry problem must be a connected subgraph. However, the solution set of the Local-Isoperimetry problem may be a disconnected graph. This occurs if there is another set in the graph with a sufficiently small isoperimetry value such that it is better to pay the price for disconnecting the reference graph $R$ and returning the union of the two. \begin{figure}[t] \subfigure[The Local-Isometry solution may be disconnected. The singleton node is the set $R$.]{\includegraphics[scale=0.33]{U3A-local-iso-1-500}}\hspace{0.01\linewidth}% \subfigure[Increasing $k$ results in a connected set for this problem.]{\includegraphics[scale=0.33]{U3A-local-iso-1-5000}} \caption{The singleton node in (a) is not a bug. The single disconnected node was the set $R$. \textfinish{Add more discussion.} \textfinish{add PGP figures}} \label{fig:local-iso} \end{figure} \section{Discussion and conclusion} \label{sec:disc_conclusion} Although the optimization approach we have adopted is designed to highlight similarities between different variants of locally-biased graph algorithms, it is also worth emphasizing that there are a number of quite different and complementary perspectives people in different research communities have adopted thus far on these methods. For example: \textbf{Theoretical and empirical.} The theoretical implications of these locally-biased algorithms are often used to improve the performance of long-standing problems in theoretical computer science by improving runtimes, improving approximation constants, and handling special cases. Empirically, these algorithms are used to \emph{study} real-world data and to accelerate and improve performance on discovery and prediction tasks. Due to the strong locality, the fast runtimes for theory often manifest as extremely fast algorithms in practice. Well-implemented strongly-local algorithms often have runtimes in milliseconds even on billion-edge graphs~\cite{KG14,FCSRM16_TR}. \textbf{Algorithmic and statistical.} Some of the work is motivated by having better algorithmic results, e.g., being fast and/or being a rigorous approximation algorithm, i.e., worst-case guarantees in terms of approximating the optimal solution of a combinatorial problem, while other work has provided an interpretation in terms of statistical properties, e.g., explicitly optimizing a regularized objective \cite{WSST14} or implicitly having output that are nice in some sense, i.e., well-connected output cluster \cite{ZLM13}. Often, locally-biased algorithms alone suffice as the result is an improvement to some downstream activity that will necessarily look at all the data anyway. \textbf{Optimization and operational.} The locally-biased methods tend to result from stating an optimization problem and solving it with some sort of black box or white box. Strongly-local algorithms often arise by studying a specific procedure on a graph and showing that it satisfies some condition, e.g., that it terminates so quickly that it cannot explore the entire graph, that it leads to a solution with certain quality-of-approximation guarantees, etc. See, for instance, the spectral algorithms \cite{ACL06,AL08,AP09,C09,ST13,CS14,KG14} and the flow-based algorithms \cite{KKOV07_TR,OSVV08,OZ14}. In light of these complementary approaches as well as the ubiquity with which graphs are used to model data, we expect that locally-biased graph algorithms and our optimization perspective on locally-biased graph algorithms will find increasing usefulness in many application areas. \section{Empirical Evaluation} \label{sec:empirical} In this section, we illustrate differences among global, weakly local, and strongly local solutions to the problems discussed in Section~\ref{sec:localgraphpart}. Additionally, we discuss differences between spectral and flow methods (which is equivalently $\ell_2$ vs. $\ell_1$ metrics for node distances). To do so, we make use of the following real-world undirected and unweighted networks. \begin{compactitem} \item \textbf{US-Senate}. Each node in this network is a Senator that served in a single term (two years) of Congress. Our data cover the period from year $1789$ to $2008$. Senators appear as multiple nodes if they served in multiple terms. Edges are based on the similarity of voting records between Senators and thresholded at the maximum similarity such that the graph remains connected. Edge-weights are discarded. For a detailed discussion of this data-set we refer the reader to \cite{multiplex_Mucha}. This graph has $8974$ nodes and $153804$ edges. This graph has two large clusters with small conductance ratio, i.e., downward-slopping network community profile; see Figure $6$ in \cite{Jeub15} for details. The first cluster consists of all the nodes before the year $1913$ and the second cluster consists of nodes after that year. \item \textbf{CA-GrQc}. The data for this graph is a general relativity and quantum cosmology collaboration network. Details can be found in the Stanford Network Analysis Project.\footnote{\url{http://snap.stanford.edu/data}} This graph has $4158$ nodes and $13422$ edges. This graph has many clusters of small size with small conductance ratio, while large clusters have large conductance ratio, i.e., upward-slopping network community profile; see Figure $6$ in \cite{Jeub15} for details. \item \textbf{FB-Johns55}. This graph is a Facebook anonymized data-set on a particular day in September $2005$ for a student social network at John Hopkins university. The graph is unweighted and it represents ``friendship" ties. The data form a subset of the Facebook100 data-set from \cite{TKMP11,TMP12}. This graph has $5157$ nodes and $186572$ edges. This is an expander-like graph, all small and large clusters have about the same conductance ratio, i.e., flat network community profile; see Figure $6$ in \cite{Jeub15} for details. \item \textbf{US-Roads}. The data for this graph is from the National Highway Planning Network \cite{LLDM08_communities_CONF}. Each node in this network is an intersection between two highways and the edges represent segments of the highways themselves. \end{compactitem} Note that the small-scale vs. large-scale clustering properties of the first three networks have been characterized previously~\cite{Jeub15}. In addition, it is known that US-Roads has a downward-sloping network community profile. \paragraph{Global, weakly local, and strongly local solutions} We first demonstrate differences among global, weakly local, and strongly local algorithms. Let us start with a comparison among spectral algorithms. By comparing algorithms that use that same metric, i.e., $\ell_2$, to measure distances among nodes we minimize factors that can affect the solution, and we focus on weak vs. strong locality. In all figures we show the solution obtained by an algorithm without applying any rounding procedure. We illustrate the importance of the nodes by colouring and size; details are explained in the captions of the figures and in the text. The layout for all graphs has been obtained using the force-directed algorithm \cite{H05}, which is available from the graph-tool project.\footnote{\url{https://graph-tool.skewed.de}} For US-Senate, the comparison is shown in Figure \ref{Fig_senate_spectral}. Figures \ref{Fig_senate_spectral_1} and \ref{Fig_senate_spectral_2} show the solutions of global algorithms, Spectral relaxation and MOV global ($z=1_{|\mathcal{V}|}$ and then we orthogonalize $z$ with respect to $D1_{|\mathcal{V}|}$), respectively. As expected, the US-Senate graph has two large clusters, i.e., before the year $1913$ and after that year, that partition along the one-dimensional time axis. This global structure is nicely captured by Spectral relaxation and MOV global in Figures \ref{Fig_senate_spectral_1} and \ref{Fig_senate_spectral_2}, respectively. Given an input seed set, Figures \ref{Fig_senate_spectral_3} and \ref{Fig_senate_spectral_4} illustrate the weakly and strongly local solutions by MOV and $\ell_1$-regularized Page-Rank, respectively. For MOV in Figures \ref{Fig_senate_spectral_3} we set $z_i = 1$ for all $i$ in the input seed set and $z_i=0$ for all $i$ outside the input seed set. Then we orthogonalize $z$ with respect to $D\cdot1_{|\mathcal{V}|}$. For $\ell_1$-regularized Page-Rank, we only give a single node as an input seed set, i.e., $h_i=1$ where $i$ is the input node and $h_i=0$ for all other nodes. Moreover, we set the locality parameter $\epsilon$ large enough such that the solution is very sparse, i.e., strongly local. In Figures \ref{Fig_senate_spectral_3} and \ref{Fig_senate_spectral_4}, we demonstrate the input seed sets by nodes with a blue halo around them. In Figure \ref{Fig_senate_spectral_3}, the cluster which is found by MOV consists of the nodes which have large mass concentration around the input seed set, i.e., the nodes around the input seed set that have large size and are coloured with a bright red shade. MOV recovers this cluster by examining the whole graph; each node has a weight assigned to it in Figure \ref{Fig_senate_spectral_3}. On the other hand, a similar cluster is found in Figure \ref{Fig_senate_spectral_4} by using $\ell_1$-regularized Page-Rank without examining the whole graph. This is possible because nodes of the graph have zero weight assigned and need not be considered. \emph{This speed and data advantage, along with the sparsity-based implicit regularization~\cite{GM14_ICML}, are some of the reasons that strongly-local algorithms, such as $\ell_1$-PageRank, are used so often in practice~\cite{LLDM09_communities_IM,Jeub15}.} \begin{figure*} \centering \subfigure[Global spectral relaxation]{\label{Fig_senate_spectral_1}\includegraphics[scale=0.22]{senate_eig2} \put(-220,0){ \begin{tikzpicture} \draw (0,0) -- (7.0,0); \foreach \x in {0,2.5,3.1,7.0} \draw (\x cm,3pt) -- (\x cm,-3pt); \draw (0.1,0) node[below=3pt] (a) {\tiny $1789$} node[above=3pt] {}; \draw (2.5,0) node[below=3pt] {\tiny $1860$} node[above=3pt] (c) {}; \draw (3.1,0) node[below=3pt](d) {} node[above=3pt] {\tiny $1913$}; \draw (6.9,0) node[above=3pt] (f) {\tiny $2008$} node[below=3pt] {}; \end{tikzpicture} } } \subfigure[MOV with global seed]{\label{Fig_senate_spectral_2} \includegraphics[scale=0.22]{senate_mov} }\\ \subfigure[MOV with local seed]{\label{Fig_senate_spectral_3} \includegraphics[scale=0.22]{senate_mov_local} } \subfigure[ACL $\ell_1$-regularized Page-Rank]{\label{Fig_senate_spectral_4}\includegraphics[scale=0.22]{senate_l1pr}} \caption{US-Senate.\ This figure shows the solutions of (a) Spectral relaxation, (b) MOV with global bias, (c) MOV with local bias and (d) strongly-local $\ell_1$-regularized Page-Rank. We use a heat map to represent the weights of the nodes. For Spectral relaxation and MOV bright yellow means large positive and bright red means large negative. For $\ell_1$-regularized Page-Rank bright yellow means large positive and bright red means small positive. The blue halo around a node in Figures \ref{Fig_senate_spectral_3} and \ref{Fig_senate_spectral_4} means that this node is included in the seed set. The size of the nodes shows the weights of the solution in absolute value. If a weight is nearly zero then the corresponding node is barely~visible. } \label{Fig_senate_spectral} \end{figure*} In Figure \ref{Fig_grqc_spectral}, we present global, weakly local, and strongly local solutions for the less well-partitionable and thus less easily-visualizable CA-GrQc graph. As already mentioned in the description of this data-set, this graph has many small clusters with small conductance ratio and large clusters have large ratio. This is also justified in our experiment by the fact that global methods, such as the Spectral relaxation and MOV global in Figures \ref{Fig_grqc_spectral_1} and \ref{Fig_grqc_spectral_2}, respectively, recover small clusters. The two global procedures find small clusters which are presented in Figures \ref{Fig_grqc_spectral_1} and \ref{Fig_grqc_spectral_2} with red, orange and yellow colours. However, since there are many small clusters of small conductance ratio, one might want to find different clusters than the ones obtained by Spectral relaxation and MOV global. This is possible using localized procedures such as MOV and $\ell_1$-regularized Page-Rank. Given two different seed sets we demonstrate in Figures \ref{Fig_grqc_spectral_3} and \ref{Fig_grqc_spectral_4} that MOV successfully finds other clusters than the ones obtained by the global methods. The same is shown in Figures \ref{Fig_grqc_spectral_5} and \ref{Fig_grqc_spectral_6} for $\ell_1$-regularized Page-Rank. Notice that MOV assigns weights (perhaps small) to all the nodes of the graph; on the other hand, $\ell_1$-regularized Page-Rank, as a strongly local procedure, assigns weights only to a small number of nodes, without examining all of the graph. \begin{figure} \centering \subfigure[Spectral relaxation]{\label{Fig_grqc_spectral_1}\includegraphics[scale=0.111]{CA-GrQc-cc_eig2}} \subfigure[MOV global]{\label{Fig_grqc_spectral_2}\includegraphics[scale=0.111]{CA-GrQc-cc_mov}} \\ \subfigure[MOV local, seed $1$]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_grqc_spectral_3}\includegraphics[scale=0.095]{CA-GrQc-cc_mov_local}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_mov_local_zoom}}}}; \draw[black] (9.0,4.05)rectangle (9.6,3.76); \draw[black,->](8.33,3.3)--(9,3.75); \end{tikzpicture} } \subfigure[MOV local, seed $2$]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_grqc_spectral_4}\includegraphics[scale=0.095]{CA-GrQc-cc_mov_local_diff_seed}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_mov_local_diff_seed_zoom}}}}; \draw[black] (8.8,4.53)rectangle (9.2,4.28); \draw[black,->](8.33,3.3)--(8.8,4.28); \end{tikzpicture} } \\ \subfigure[$\ell_1$-regularized Page-Rank, seed $1$]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_grqc_spectral_5}\includegraphics[scale=0.095]{CA-GrQc-cc_l1pr}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_l1pr_zoom}}}}; \draw[black] (9.0,4.05)rectangle (9.6,3.76); \draw[black,->](8.33,3.3)--(9,3.75); \end{tikzpicture} } \subfigure[$\ell_1$-regularized Page-Rank, seed $2$]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_grqc_spectral_6}\includegraphics[scale=0.095]{CA-GrQc-cc_l1pr_diff_seed}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_l1pr_diff_seed_zoom}}}}; \draw[black] (8.8,4.53)rectangle (9.2,4.28); \draw[black,->](8.33,3.3)--(8.8,4.28); \end{tikzpicture} } \caption{CA-GrQc.\ This figure shows the solutions of Spectral relaxation, MOV with global input, MOV with local input and $\ell_1$-regularized Page-Rank. The meaning of the colours of the nodes and its sizes is the same as in Figure \ref{Fig_senate_spectral}.} \label{Fig_grqc_spectral} \end{figure} We now use the FB-Johns55 graph which has an expander-like behaviour at all size scales, i.e., all small and large clusters have large conductance ratio. See Figure $6$ in \cite{Jeub15} for details. We present the results of this experiment in Figure \ref{Fig_fb_spectral}. Notice that in Figures \ref{Fig_fb_spectral_1} and \ref{Fig_fb_spectral_2} the global methods identify three main clusters, one small (red nodes), one medium size (orange nodes) and one large (yellow nodes). All these clusters have similar conductance ratio. In Figures \ref{Fig_fb_spectral_3} and \ref{Fig_fb_spectral_4} we show that MOV can recover the medium or the small size clusters, respectively, by giving a localized seed set. In Figures \ref{Fig_fb_spectral_5} and \ref{Fig_fb_spectral_6} we illustrate that using $\ell_1$-regularized Page-Rank one can find very similar clusters while exploiting the strongly local running time of the method. \begin{figure} \centering \subfigure[Spectral relaxation]{\label{Fig_fb_spectral_1}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_eig2}} \subfigure[MOV global]{\label{Fig_fb_spectral_2}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_mov}} \\ \subfigure[MOV local]{\label{Fig_fb_spectral_3} \includegraphics[scale=0.11]{Johns-Hopkins55-cc_mov_local}} \subfigure[MOV local, seed $2$]{\label{Fig_fb_spectral_4}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_mov_local_diff_seed}} \\ \subfigure[$\ell_1$-regularized Page-Rank]{\label{Fig_fb_spectral_5}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_l1pr}} \subfigure[$\ell_1$-regularized Page-Rank, seed $2$]{\label{Fig_fb_spectral_6}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_l1pr_diff_seed}} \caption{FB-Johns55.\ This figure shows the solutions of Spectral relaxation, MOV with global input, MOV with local input and $\ell_1$-regularized Page-Rank. The meaning of the colours of the nodes and its sizes is the same as in Figure \ref{Fig_senate_spectral}.} \label{Fig_fb_spectral} \end{figure} Let us now present the performance of flow-based algorithms on the same graphs. We begin with US-Senate in Figure \ref{Fig_senate_flow}. In this figure, the red nodes are part of the solution of Flow Improve or Local Flow Improve, depending on the experiment; the yellow nodes are part of the seed set only; and the orange nodes are in both the solution of the flow algorithm and the input seed set. In Figure \ref{Fig_senate_flow_1}, we used as an input seed set to Flow Improve the cluster obtained by applying sweep cut with respect to the conductance ratio on the Spectral relaxation solution. Figures \ref{Fig_senate_flow_2} and \ref{Fig_senate_flow_3} present a clear distinction between Flow Improve and Local Flow Improve, weakly and strongly local algorithms, respectively. For both figures, the input seed set is located at about the middle of the graph. Flow Improve as a weakly local algorithm examines the whole graph and returns a cluster which includes the period before $1913$. Also, it includes big part of the input seed set in the cluster due to the overlapping regularization term in the denominator of its objective function. See the definition of the objective function $\phi_R$ for Flow Improve in Section \ref{sec:localgraphpart}. On the other hand, in Figure \ref{Fig_senate_flow_3} Local Flow Improve as a strongly local algorithm does not examine the whole graph and its solution is concentrated only around the input seed set. \begin{figure*} \centering \subfigure[Local Flow Improve, seed: Spectral relaxation + sweep cut]{\label{Fig_senate_flow_1}\includegraphics[scale=0.15]{senate_flowI_s_local_seed1} \hspace{-5.5cm} \begin{tikzpicture} \draw (0,0) -- (4.8,0); \foreach \x in {0,1,1.5,4.8} \draw (\x cm,3pt) -- (\x cm,-3pt); \draw (0.1,0) node[below=3pt] (a) {\tiny $1789$} node[above=3pt] {}; \draw (1,0) node[below=3pt] {\tiny $1860$} node[above=3pt] (c) {}; \draw (1.5,0) node[below=3pt](d) {} node[above=3pt] {\tiny $1913$}; \draw (4.7,0) node[above=3pt] (f) {\tiny $2008$} node[below=3pt] {}; \end{tikzpicture} } \subfigure[Flow Improve, local seed set]{\label{Fig_senate_flow_2}\includegraphics[scale=0.15]{senate_flowI_w_local_seed2}} \subfigure[Local Flow Improve, local seed set]{\label{Fig_senate_flow_3}\includegraphics[scale=0.15]{senate_flowI_s_local_seed2}} \caption{US-Senate.\ This figure shows the solutions of Flow Improve and Local Flow Improve for various input seed sets. Red nodes are only Flow Improve or Local Flow Improve, depending on the experiment; yellow nodes are only seed set; and orange nodes are both part of the flow algorithm and the seed set.} \label{Fig_senate_flow} \end{figure*} The distinction that we discussed in the previous paragraph between Flow Improve and Local Flow Improve is easy to visualize in the relatively well-structured US-Senate, but it is not so clear in all graphs. For example, in Figure \ref{Fig_cagrqc_flow} we present the performance of these two algorithms for the CA-GrQc graph. Since this graph has only small clusters of small conductance ratio, Flow Improve and Local Flow Improve find the same clusters. This is clearly shown by comparing Figures \ref{Fig_cagrqc_flow_1} and \ref{Fig_cagrqc_flow_2} and Figures \ref{Fig_cagrqc_flow_3} and \ref{Fig_cagrqc_flow_4}. A similar performance is observed for the FB-Johns55 graph in Figure \ref{Fig_fb_flow}, except that the solution of Flow Improve and Local Flow Improve are not exactly the same but only very similar. \begin{figure} \centering \subfigure[Flow Improve, seed $1$]{\label{Fig_cagrqc_flow_1} \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\includegraphics[scale=0.095]{CA-GrQc-cc_flowI_w_local_seed1}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_flowI_w_local_seed1_zoom}}}}; \draw[black] (9.0,4.05)rectangle (9.6,3.76); \draw[black,->](8.33,3.3)--(9,3.75); \end{tikzpicture} } \subfigure[Local Flow Improve, seed $1$]{\label{Fig_cagrqc_flow_2} \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\includegraphics[scale=0.095]{CA-GrQc-cc_flowI_s_local_seed1}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_flowI_s_local_seed1_zoom}}}}; \draw[black] (9.0,4.05)rectangle (9.6,3.76); \draw[black,->](8.33,3.3)--(9,3.75); \end{tikzpicture} }\\ \subfigure[Flow Improve, local seed $2$]{\label{Fig_cagrqc_flow_3} \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\includegraphics[scale=0.095]{CA-GrQc-cc_flowI_w_local_seed2}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_flowI_w_local_seed2_zoom}}}}; \draw[black] (8.8,4.53)rectangle (9.2,4.28); \draw[black,->](8.33,3.3)--(8.8,4.28); \end{tikzpicture} } \subfigure[Local Flow Improve, seed $2$]{\label{Fig_cagrqc_flow_4} \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\includegraphics[scale=0.095]{CA-GrQc-cc_flowI_s_local_seed2}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.04]{CA-GrQc-cc_flowI_s_local_seed2_zoom}}}}; \draw[black] (8.8,4.53)rectangle (9.2,4.28); \draw[black,->](8.33,3.3)--(8.8,4.28); \end{tikzpicture} } \caption{CA-GrQc.\ This figure shows the solutions of Flow Improve and Local Flow Improve for various input seed sets. Red nodes are only Flow Improve or Local Flow Improve, depending on the experiment; yellow nodes are only seed set; and orange nodes are both part of the flow algorithm and the seed set.} \label{Fig_cagrqc_flow} \end{figure} \begin{figure} \centering \subfigure[Flow Improve, seed $1$]{\label{Fig_fb_flow_1}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_flowI_w_local_seed1}} \subfigure[Local Flow Improve, seed $1$]{\label{Fig_fb_flow_2}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_flowI_s_local_seed1}}\\ \subfigure[Flow Improve, seed $2$]{\label{Fig_fb_flow_3}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_flowI_w_local_seed2}} \subfigure[Local Flow Improve, seed $2$]{\label{Fig_fb_flow_4}\includegraphics[scale=0.11]{Johns-Hopkins55-cc_flowI_s_local_seed2}} \caption{FB-Johns55. This figure shows the solutions of Flow Improve and Local Flow Improve for various input seed sets. Red nodes are only Flow Improve or Local Flow Improve, depending on the experiment; yellow nodes are only seed set; and orange nodes are both part of the flow algorithm and the seed set.} \label{Fig_fb_flow} \end{figure} \paragraph{Flow vs. spectral, or $\ell_1$ vs. $\ell_2$} Spectral algorithms measure distances of the nodes based on the $\ell_2$ norm. Generally this means that the nodes of the graph are embedded on the real line. On the other hand, flow algorithms measure distances of the nodes based on the $\ell_1$ norm. The solution to flow-based algorithms that we discussed is binary, either a node is added in the solution with weight $1$ or it is not and it has weight $0$. In this case, the objective function $\|Bx\|_{1,C}$ of the flow algorithms is a locally-biased variation on $\cut(S)$, where $S$ is constructed based on the binary $x$.\ Therefore, the flow algorithms aim to find a balance between finding good cuts and identifying the input seed set. This implies that the flow algorithms try to minimize the absolute number of edges that cross the partition, but at the same time they try to take into account the volume regularization effect of the denominator in the objective function. In this section, we will try to isolate the effect of $\ell_1$ and $\ell_2$ metrics in the output solution. We do this by employing MQI and spectral MQI, which are flow (i.e., $\ell_1$) and spectral (i.e., $\ell_2$) problems, respectively. The first set of results is shown in Figure \ref{Fig_senate_l1vl2}. Notice in Figure \ref{Fig_senate_l1vl2_1} and \ref{Fig_senate_l1vl2_2} that MQI and spectral MQI + sweep cut recover the large clusters, i.e., before and after the year $1913$. There are only minor differences between the two solutions. Moreover, observe that Spectral MQI returns a solution which is not binary. This is illustrated in Figure \ref{Fig_senate_l1vl2_3}, where the weights of the nodes are real numbers. Then sweep cut is applied on the solution of spectral MQI to obtain a binary solution with small conductance ratio, i.e., Figure \ref{Fig_senate_l1vl2_2}. The previous example did not reveal any difference between MQI and spectral MQI other than the fact that spectral MQI has to be combined with the sweep cut rounding procedure to obtain a binary solution. In Figure \ref{Fig_usroads_l1vl2}, we present a result showing where the solutions have substantial differences. The graph that we used for this is the US-Roads, and the input seed set consists of nodes near Minneapolis together with some suburban areas around the city. Notice in Figures \ref{Fig_usroads_l1vl2_1} that MQI, i.e., $\ell_1$ metric, shrinks the boundaries of the input seed set. However, MQI does not accurately recover Minneapolis. The reason is the volume regularization which is imposed by the denominator of the objective function of MQI. This regularization forces the solution to have large volume. On the other hand, spectral MQI + sweep cut in Figure \ref{Fig_usroads_l1vl2_2} recovers Minneapolis. The reason is that for spectral MQI the regularization effect of the denominator is unimportant since the objective function is invariant to scalar multiplications of the solution vector. It's the solution of spectral MQI, i.e., the eigenvector of smallest eigenvalue, which is presented in Figure \ref{Fig_usroads_l1vl2_3}, that is concentrated closely around Minneapolis. Due to this concentration of the eigenvector around Minneapolis, the sweep is successful. Briefly, spectral MQI, which is a continuous relaxation of MQI, implicitly offers an additional level of volume regularization, which turns out to be useful in this example. Finally, we present another set of results using the FB-Johns55 graph in Figure \ref{Fig_fb_l1vl2}. As we saw before, notice that for this less well-structured graph the solutions of MQI and spectral MQI + sweep cut are nearly the same. This happens because the regularization effect of the denominator of MQI and the regularization imposed by spectral MQI have nearly the same effect on this graph. This is also justified by the fact that MQI in Figure \ref{Fig_fb_l1vl2_1} and spectral MQI without sweep cut in Figure \ref{Fig_fb_l1vl2_3} recover nearly the same cluster. \begin{figure*} \centering \subfigure[MQI, seed: Spectral relaxation + sweep cut]{\label{Fig_senate_l1vl2_1}\includegraphics[scale=0.15]{senate_mqi} \hspace{-5.5cm} \begin{tikzpicture} \draw (0,0) -- (4.8,0); \foreach \x in {0,1,1.5,4.8} \draw (\x cm,3pt) -- (\x cm,-3pt); \draw (0.1,0) node[below=3pt] (a) {\tiny $1789$} node[above=3pt] {}; \draw (1,0) node[below=3pt] {\tiny $1860$} node[above=3pt] (c) {}; \draw (1.5,0) node[below=3pt](d) {} node[above=3pt] {\tiny $1913$}; \draw (4.7,0) node[above=3pt] (f) {\tiny $2008$} node[below=3pt] {}; \end{tikzpicture} } \subfigure[Spectral MQI + sweep cut, seed: Spectral relaxation + sweep cut]{\label{Fig_senate_l1vl2_2}\includegraphics[scale=0.15]{senate_smqi_sweep}} \subfigure[Spectral MQI, seed: Spectral relaxation + sweep cut]{\label{Fig_senate_l1vl2_3}\includegraphics[scale=0.15]{senate_smqi}} \caption{US-Senate.\ This figure shows the solutions of MQI, spectral MQI, spectral MQI + sweep cut given the solution of Spectral relaxation + sweep cut as an input seed set. For Figures \ref{Fig_senate_l1vl2_1} and \ref{Fig_senate_l1vl2_2} the red nodes are only MQI or spectral MQI + sweep cut depending on the experiment; yellow nodes are only seed set; and orange nodes are both part of the flow or spectral algorithm and the seed set. Figure \ref{Fig_senate_l1vl2_3} shows the solution of spectral MQI without sweep cut. For Figure \ref{Fig_senate_l1vl2_3} we use a heat map to represent the weights of the nodes. Bright yellow means large positive and bright red means small positive. The size of the nodes shows the weights of the solution in absolute value. } \label{Fig_senate_l1vl2} \end{figure*} \begin{figure*} \centering \subfigure[MQI, seed: Minneapolis and suburban areas]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_usroads_l1vl2_1}\includegraphics[scale=0.133]{usroads_mqi}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.056]{usroads_mqi_zoom}}}}; \draw[black] (8,4.75)rectangle (8.35,4.5); \draw[black,->](7.5,4.0)--(8,4.5); \end{tikzpicture} } \subfigure[Spectral MQI + sweep cut, seed: Minneapolis and suburban areas]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_usroads_l1vl2_2}\includegraphics[scale=0.133]{usroads_smqi_sweep}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.056]{usroads_smqi_sweep_zoom}}}}; \draw[black] (8,4.75)rectangle (8.35,4.5); \draw[black,->](7.5,4.0)--(8,4.5); \end{tikzpicture} } \subfigure[Spectral MQI, seed: Minneapolis and suburban areas]{ \begin{tikzpicture}[every node/.style={anchor=center}] \node(a) at (8,4){\label{Fig_usroads_l1vl2_3}\includegraphics[scale=0.133]{usroads_smqi}}; \node(b) at (7.5,3.3){\frame{\colorbox{white}{\includegraphics[scale=0.056]{usroads_smqi_zoom}}}}; \draw[black] (8,4.75)rectangle (8.35,4.5); \draw[black,->](7.5,4.0)--(8,4.5); \end{tikzpicture} } \caption{US-Roads.\ This figure shows the solutions of MQI, spectral MQI, spectral MQI + sweep cut given Minneapolis and its suburban areas as an input seed set. The meaning of the colours of the nodes and its sizes is the same as in Figure \ref{Fig_senate_l1vl2}.} \label{Fig_usroads_l1vl2} \end{figure*} \begin{figure*} \centering \subfigure[MQI]{\label{Fig_fb_l1vl2_1}\includegraphics[scale=0.14]{Johns-Hopkins55-cc_mqi}} \subfigure[Spectral MQI + sweep cut]{\label{Fig_fb_l1vl2_2}\includegraphics[scale=0.14]{Johns-Hopkins55-cc_smqi_sweep}} \subfigure[Spectral MQI]{\label{Fig_fb_l1vl2_3}\includegraphics[scale=0.14]{Johns-Hopkins55-cc_smqi}} \caption{FB-Johns55. This figure shows the solutions of MQI, spectral MQI, spectral MQI + sweep cut for a given input seed set. The meaning of the colours of the nodes and its sizes is the same as in Figure \ref{Fig_senate_l1vl2}.} \label{Fig_fb_l1vl2} \end{figure*} \section{Conductance partitioning} \section{ Sparsest-Cut, Minimum-Conductance, and Spectral~Relaxations}\label{sec:global_cond_spcut} In this section, we present two ubiquitous combinatorial optimization problems: Sparsest-Cut and Minimum-Conductance. These problems are NP-hard \cite{SM90,LR99}, but they can be relaxed to tractable convex optimization problems~\cite{ARV09}. We discuss one of the commonly used relaxation techniques which will motivate part of our discussion for local graph clustering algorithms. Both Sparsest-Cut and Minimum-Conductance give different ways of balancing the size of a partition with its quality. Sparsest-Cut finds the partition that minimizes the ratio of the fraction of edges that are removed divided by the scaled product of volumes of the two disjoint sets of nodes defined by removing those edges. In particular, if we have a partition $(S, S^c)$, where $S^c$ is defined in Section \ref{sec:prelim}, then $\cut(S)$ is the number of edges that are removed, and the scaled product of volumes $\vol(S)\vol(S^c)/\vol(\mathcal{V})$ is the volume of the disjoint sets of nodes. Putting all together in an optimization problem, we obtain \begin{align}\label{spcut} \tilde{\phi}(\mathcal{G}):= \mbox{minimize} & \ \tilde{\phi}(S):= \frac{\cut(S)}{\frac{1}{\vol(\mathcal{V})}\vol(S)\vol(S^c)} \\\nonumber \mbox{subject to} &\ S\subset \mathcal{V}. \end{align} We shall use the term \emph{expansion of a set} to refer to the ratio $\tilde{\phi}(S)$, and the term \emph{expansion of the graph} to refer to $\tilde{\phi}(\mathcal{G})$, in which case this problem is known as the Sparsest-Cut problem. Another way of balancing the partition is \begin{align}\label{isop} \phi(\mathcal{G}):= \mbox{minimize} & \ \phi(S):= \frac{\cut(S)}{\min(\vol(S),\vol(S^c))} \\\nonumber \mbox{subject to} &\ S\subset \mathcal{V}. \end{align} In this case, we divide by the minimum of $\vol(S)$ and $\vol(S^c)$, rather than their product. We shall use the term \emph{conductance of a set} to refer to the ratio of cut to volume $\phi(S)$, and the term \emph{conductance of the graph} to refer to $\phi(\mathcal{G})$, in which case this problem is known as the Minimum-Conductance problem. The difference between the two problems \eqref{spcut} and \eqref{isop} is that the former regularizes based on the number of connections lost among pairs of nodes, while the latter regularizes based on the size of the small side of the partition. Optimal solutions to these problems differ by a factor of $2$: \begin{equation}\label{eq:equiv_cond_sparsecut} \frac{1}{2}\tilde{\phi}(S) \le \phi(S) \le \tilde{\phi}(S), \end{equation} leading to the two objectives $\tilde{\phi}(S)$ and $\phi(S)$ being almost substitutable from a theoretical computer science perspective~\cite{ARV09}. However, this does not mean that the actual obtained solutions by solving \eqref{spcut} and \eqref{isop} are similar; in general, they are not. There are three major relaxation techniques for the NP-hard problems \eqref{spcut} and \eqref{isop}: spectral relaxation; all pairs multi-commodity flow or linear programming (LP) relaxation; and semidefinite programming (SDP) relaxation. For detailed descriptions about the LP and SDP relaxations we refer the reader to \cite{LR99} and \cite{ARV09}, respectively. We focus here on spectral relaxation since similar relaxations are widely used for the development of local clustering methods, which we discuss in subsequent sections. \subsection{Spectral Relaxation}\label{subsec:spcut} Spectral graph partitioning is one of the best known relaxations of the Sparsest-Cut \eqref{spcut} and Minimum-Conductance \eqref{isop}. The relaxation is the same for both problems, although the diversity of derivations of spectral partitioning does not always make this connection clear. For a partition $(S, S^c)$, let's associate a vector $x\in\{c_1,c_2\}^{|\mathcal{V}|}$ such that $x_i=c_1$ if $v_i\in S$ and $x_i=c_2$ if $v_i\in S^c$. (For simplicity, think of $c_1 = 1$ and $c_2 = 0$, so $x$ is the set indicator vector.) The spectral clustering relaxation uses a continuous relaxation of the set indicator vector in problems \eqref{spcut} and \eqref{isop} to produce an eigenvector. The relaxed problem is \begin{align}\label{spcutLD} \lambda_2 := \mbox{minimize} & \ \frac{\|Bx\|_{2,C}^2}{2\|x\|_{2,D}^2} \\\nonumber \mbox{subject to} &\ 1_{|\mathcal{V}|}^T D x = 0\\\nonumber &\ x\in\mathbb{R}^{|\mathcal{V}|} - \{0_{|\mathcal{V}|}\}. \end{align} (The denominator $\|x\|_{2,D}^2 = \sum_i |x_i|^2 d_i$.) To see why \eqref{spcutLD} is a continuous relaxation of \eqref{spcut} we make two observations. First, notice that for all $x$ in $\mathbb{R}^{|\mathcal{V}|} - \{0_{|\mathcal{V}|}\}$ such that $1_{|\mathcal{V}|}^T D x = 0$ the denominator in \eqref{spcutLD} satisfies $2\vol(\mathcal{V})\|x\|_{2,D}^2 = \sum_{i=1}^{|\mathcal{V}|}\sum_{j=1}^{|\mathcal{V}|} d_id_j|x_i-x_j|^2$. Therefore, $\lambda_2$ in \eqref{spcutLD} is equivalent to \begin{align}\label{eq:1434} \lambda_2 = \mbox{minimize} & \ \frac{\|Bx\|_{2,C}^2}{\frac{1}{\vol(\mathcal{V})}\sum_{i=1}^{|\mathcal{V}|}\sum_{j=1}^{|\mathcal{V}|} d_id_j|x_i-x_j|^2} \\\nonumber \mbox{subject to} &\ 1_{|\mathcal{V}|}^T D x = 0\\\nonumber &\ x\in\mathbb{R}^{|\mathcal{V}|} - \{0_{|\mathcal{V}|}\}. \end{align} The optimal value of the right hand side in \eqref{eq:1434} is equivalent to the optimal value of right hand side in the following expression \begin{align}\label{eq:1435} \lambda_2 = \mbox{minimize} & \ \frac{\|Bx\|_{2,C}^2}{\frac{1}{\vol(\mathcal{V})}\sum_{i=1}^{|\mathcal{V}|}\sum_{j=1}^{|\mathcal{V}|} d_id_j|x_i-x_j|^2} \\\nonumber \mbox{subject to} & \ x\in\mathbb{R}^{|\mathcal{V}|} - \{0_{|\mathcal{V}|},1_{|\mathcal{V}|}\}. \end{align} To prove this notice that the objective function of the right hand side in \eqref{eq:1435} is invariant to constant shifts of $x$, i.e., $x$ and $x + c 1_{|\mathcal{V}|}$ have the same objective function, where $c$ is a constant. Therefore, if $\tilde{x}$ is an optimal solution of the right hand side in \eqref{eq:1435} then $\hat{x} = \tilde{x} - \frac{1^T_{|\mathcal{V}|}D\tilde{x}}{\vol(\mathcal{V})} 1_{|\mathcal{V}|}$ has the same optimal objective value and also $1_{|\mathcal{V}|}^TD \hat{x} = 0$. Second, by restricting the solution in \eqref{eq:1435} in $\{0,1\}^{|\mathcal{V}|}$ instead of $\mathbb{R}^{|\mathcal{V}|}$ we get that $\cut(S) = \|Bx\|_{2,C}^2$ and $\sum_{i=1}^{|\mathcal{V}|}\sum_{j=1}^{|\mathcal{V}|} d_id_j|x_i-x_j|^2 = \vol(S)\vol(S^c)$. Using these two observations, it is easy to see that \eqref{spcutLD} is a continuous relaxation of \eqref{spcut}. Using \eqref{eq:equiv_cond_sparsecut}, it is easy to see that \eqref{spcutLD} is a relaxation for \eqref{isop} as well. The quality of approximation of relaxation \eqref{spcutLD} to Sparsest-Cut \eqref{spcut} is given by Cheeger's inequality \cite{LS88,SJ89} $${\lambda_2}/{\vol(\mathcal{V})}\le \tilde{\phi}(\mathcal{G}) \le {(8 \lambda_2)^{1/2}}/{\vol(\mathcal{V})} $$ while the approximation guarantee for the Minimum-Conductance problem \eqref{isop} is $$ {\lambda_2}/{2}\le \phi(\mathcal{G}) \le (2 \lambda_2)^{1/2}, $$ which can be found in \cite{bM89}. (A generalization of these bounds holds for arbitrary vectors~\cite{Mihail}.) Both of these approximation ratios can be realized by rounding procedures described below. Another form of relaxation is the combinatorial model relaxation, which is formulation as problem \eqref{spcutLD} by ignoring the orthogonality constraint and restricting $x\in\{0,1\}^{|\mathcal{V}|}$ instead of $x\in\mathbb{R}^{|\mathcal{V}|}$. An extensive study of the spectral and the combinatorial mode relaxation can be found in \cite{Hoc13}, while empirical comparisons between these relaxations are discussed in \cite{HLB13}. \subsection{Rounding} \label{subsec:rounding} In practice, the solution obtained by the spectral relaxation is unlikely to lie in $\{0,1\}^{|\mathcal{V}|}$, i.e., it is unlikely to be the indicator vector of a set. Therefore, it is necessary to have an efficient post-processing procedure where the solution is rounded to a set. At the same time it is important to guarantee that the rounded solution has good worst-case guarantees in terms of the conductance or sparsest cut objective. One of the most efficient and theoretically justified rounding procedures for spectral relaxation is the \emph{sweep cut}. The sweep cut procedure is summarized in the following steps. \begin{compactenum}\footnotesize \item Input: the solution $x\in\mathbb{R}^{|\mathcal{V}|}$ of \eqref{spcutLD}. \item Sort the indices of $x$ in decreasing order with respect to the values of the components of $x$. Let $i_1,i_2,\dots,i_{|\mathcal{V}|}$ be the sorted indices. \item Using the sorted indices generate a collection of sets $\mathcal{S}_{j}:=\{i_1,i_2,\dots,i_j\}$ for each $j\in \{1,2,\dots,|\mathcal{V}|\}$. \item Compute the conductance or sparsest-cut objective for each set $S_j$ and return the minimum. \end{compactenum} Notice that sweep cut can be used to obtain approximate solutions for both Sparsest-Cut and Minimum-Conductance. In fact, the proof for the upper inequalities of the approximation guarantees of spectral relaxation to Sparsest-Cut and Minimum-Conductance are obtained by using the sweep cut procedure \cite{LS88,SJ89}. \section{Regularization, $\ell_2$ versus $\ell_1$} In the previous section we followed a theoretical computer science perspective. We discussed tractable combinatorial problems, i.e., min-cut and $s$-$t$ min-cut, and NP-hard combinatorial problems, i.e., minimum conductance and isoperimetry. We discussed how to relax the latter NP-hard problems in order to obtain continuous versions that are ``good" approximations in terms of worst-case objective value. In this section we make a step back and study the regularization properties that are imposed by min-cut and certain spectral problems. Understanding the regularization properties that are imposed by each problem on the solution will allow us to understand the structure of the output solution in practice. In subsequent sections we will use the regularization problems by min-cut and spectral to understand what properties are imposed on the solution by current state-of-the-art local algorithms. \paragraph{Properties of $s$-$t$ Min-Cut} We now discuss regularization properties for the $s$-$t$ min-cut problem. Similar properties hold for min-cut since this is the best $s$-$t$ min-cut over all pairs of nodes in $\mathcal{V}$. It can be shown \cite{} that the following optimization problem \begin{align}\label{eq:mincutL1} \nonumber \mbox{minimize} & \quad \|Bx\|_{1,C}\\ \mbox{subject to} & \quad x_s =1, \ x_t = 0 \\\nonumber &\quad x\ge 0. \end{align} is equivalent to $s$-$t$ min-cut, which we introduced in Section \ref{subsec:mincutmaxflow}. The solution to $s$-$t$ min-cut is binary and lies in $[0,1]^{|\mathcal{V}|}$, see \cite{}. Thus, the solution defines a partition $S, S^c$. Since we assumed that the given graph has edges with unit-weights then the $s$-$t$ min-cut objective penalizes partitions $S,S^c$ that have many outgoing edges from $S$ to $S^c$. Intuitively, this implies that min-cut looks for partitions which have clear boundaries. For many applications, depending on the given graph such a property might be very important. Later on we will see that there are local algorithms which preserve this property close to a given input seed set. \paragraph{Properties of spectral} Let us now replace the $\ell_1$ norm in \eqref{eq:mincutL1} with the $\ell_2$ norm to obtain the following problem \begin{align}\label{eq:spectralL2} \nonumber \mbox{minimize} & \quad \|Bx\|_{2,C}\\ \mbox{subject to} & \quad x_s =1, \ x_t = 0 \\\nonumber &\quad x\ge 0. \end{align} From the preliminaries we know that $\|Bx\|_{2,C} = x^T L x$. This problem is related to spectral relaxation \eqref{spcutLD}. The only difference between \eqref{eq:spectralL2} and \eqref{spcutLD} is that we do not scale the $\ell_2$ norm in the constraint with the diagonal matrix $D$. This means that we do not put weights on nodes based on their degree, i.e., all nodes are of the same importance. Similarly to our discussion on spectral relaxation it can be shown that problem \eqref{eq:spectralL2} is a relaxation of the Minimum-Conductance and Minimum-Isoperimetry problems where all nodes are of the same importance. However, in this section our focus is the properties imposed by the $\ell_2$ norm and not the relaxation properties of problem \eqref{eq:spectralL2}. The solution to \eqref{eq:spectralL2} satisfies $$ L x = \lambda_2 x, $$ where $\lambda_2$ is the second smallest eigenvalue of $L$ for a unit-length eigenvector $x$. Since $L$ is the Laplacian operator of the given graph, then the eigenvector which satisfies the previous equation can be seen as the solution to a diffusion process. This means that the solution to problem \eqref{eq:spectralL2} is generally ``smooth", meaning that the values of the components of $x$ are in $\mathbb{R}$ and they are not binary. This is an important difference between \eqref{eq:mincutL1} and \eqref{eq:spectralL2}, which is due to the geometry that $\ell_1$ and $\ell_2$ norms define. Intuitively, the ``smoothness" of the solution given by spectral relaxation implies that we loose the property of \eqref{eq:mincutL1}, which is to preserve boundaries. If one takes a clustering perspective, then this drawback can be fixed by combining spectral with some sweep procedure which will possibly ignore components of $x$ with very small values, i.e., nodes outside the boundary. This two step-procedure is provably at most as good as \eqref{eq:mincutL1} if our objective is to find the best partition with the fewest outgoing edges. However, one should not conclude that using $\ell_1$ norm than $\ell_2$ defines a better geometry. The decision between $\ell_1$ and $\ell_2$ is application depended. If one wants to preserve boundaries then they should use the $\ell_1$ norm. If one wants to perform a diffusion on a given graph, then one should use the $\ell_2$ norm. \section{Locally-biased graph partitioning methods}\label{sec:localgraphpart} All of the algorithms described in Section~\ref{sec:global_cond_spcut} are ``global,'' in that they touch all of the nodes of the input graph at least once, and thus they have a running time that is at least linear in the size of the input graph. Informally, locally-biased graph clustering algorithms find clusters near a specified seed set of vertices, in many cases without even touching the entire input graph. In this section, we will describe several local graph clustering algorithms, each of which has somewhat different properties. To understand the seemingly-quite-different algorithms we will discuss, we will distinguish local graph clustering algorithms based on three major features. \begin{compactenum} \item \textit{Weakly or strongly local algorithms}. Weakly local algorithms are those that are biased toward a local part of the graph but may ``touch'' the entire input graph during the computation---i.e., they formally have running times that scale with the size of the graph. Strongly-local graph algorithms are those that only access a small portion of the entire input graph in order to perform their computation---i.e., they formally have running times that are linear in the size of the output or input set of nodes but independent of the size of the input graph. We show that the difference between weakly and strongly local algorithms often translates to whether we penalize the solution by adding an $\ell_1$ norm penalty implicitly to the objective function and/or by restricting the feasible region of the problem by adding implicitly a locality constraint. Both ways result in strongly local algorithms. \item \textit{Localizing bias.} Current local graph clustering algorithms are supervised, i.e., one has to give a reference set of seed nodes. We discuss two major ways that this information is incorporated in the problem. First, the bias to the input is incorporated in the objective function of the problem through a localizing construction we will call the \emph{reference cut graph}. Second, the bias to the input is incorporated to the problem as a constraint. \item \textit{$\ell_1$ and $\ell_2$ metrics.} The final major feature that distinguishes locally-biased algorithms is how the cut measure is treated. Via the cut metric~\cite{DezaLaurent97}, this can be viewed as embedding the vertices of $\mathcal{G}$ and evaluating their distance in the embedding. Two distances are explicitly or implicitly used in locally-biased algorithms: the $\ell_1$ and the $\ell_2$ metric spaces. The former results in local flow algorithms, and the latter results in local spectral algorithms. The distinction between these two is very important in practice, as we show by empirical evaluation in subsequent sections. \end{compactenum} \noindent The local graph clustering algorithms that we consider in the following sections and their basic properties with respect to the above three features are given in Table \ref{table1}. \begin{table} \centering \begin{tabular}{lcccc} \toprule \multicolumn{1}{c}{Methods} & Locality & Bias & Metric \\ \midrule Flow-Improve \cite{AL08_SODA} & Weak & Objective & $\ell_1$ \\ MOV \cite{MOV12_JMLR} & Weak & Constraint & $\ell_2$ \\ Local Flow-Improve \cite{OZ14} & Strong & Objective & $\ell_1$ \\ MQI \cite{kevin04mqi} & Strong & Constraint & $\ell_1$ \\ spectral MQI \cite{Chung07_localcutsLAA} & Strong & Constraint & $\ell_2$ \\ $\ell_1$-reg. Page-Rank \cite{ACL06,GM14_ICML} & Strong & Objective & $\ell_2$ \\ \bottomrule \end{tabular} \caption{State-of-the-art local graph clustering methods and their properties with respect to the three features that are discussed in Section \ref{sec:localgraphpart}.} \label{table1} \end{table} \subsection{A general localizing construction} We describe these locally-biased graph methods in terms of an augmented graph we call the \emph{reference cut graph}. We should emphasize that this is a conceptual construction to highlight the similarities and differences between locally-biased algorithms in Table~\ref{table1}---in particular, these algorithms do \emph{not} explicitly construct the reference cut graph. Let $h,g\in\mathbb{R}^{|\mathcal{V}|}$, $h,g\ge 0$, and $g-h \ge 0$, and let $\alpha$, $\beta$, and $\gamma$ be parameters specified below.. Then the reference cut graph is constructed from a simple, undirected graph $\mathcal{G}$ as follows: \begin{compactenum}\footnotesize \item Add a source and sink node $s$ and $t$. \item Add edges from $s$ to each node in $\mathcal{V}$. \item Add edges from each node in $\mathcal{V}$ to $t$. \item Weight each original edge by $\gamma$. \item Weight the edges from $s$ to $\mathcal{V}$ by $\alpha h$, where $\alpha \ge 0$. \item Weight the edges from $t$ to $\mathcal{V}$ by $\beta (g - h)$, $\beta \ge 0$. \end{compactenum} Let $H := diag(h)$, $G := diag(g)$ and $Z = G-H$. Then we can also view the augmented graph through its incidence matrix and weight matrix: \[ \tilde{B} = \begin{bmatrix} 1_{|\mathcal{V}|} & -I_{|\mathcal{V}|} & 0 \\ 0 & B & 0 \\ 0 & I_{|\mathcal{V}|} & -1_{|\mathcal{V}|} \end{bmatrix}, \ \tilde{C}= \begin{bmatrix} \alpha H & 0 & 0 \\ 0 & \gamma C & 0 \\ 0 & 0 & \beta Z \end{bmatrix}, \] respectively. The above construction might look overly complicated. However, we will see in the following subsections that it simplifies for local spectral and flow graph clustering algorithms with specific settings for $h,g$ and $\gamma$. \subsection{Weakly-Local and Strongly-Local flow methods} Although finding the set of minimum conductance is NP-hard in general, there are a number of cases and variations that admit polynomial time algorithms and can be solved via Max-Flow/Min-Cut or a parametric Max-Flow method. These algorithms begin with a reference set $R$ of nodes, and they return a smaller (or not much larger) set of nodes that is a better partition in terms of the conductance ratio $\phi$. Typically, the returned value is \emph{optimal} for a variation on the Minimum-Conductance and/or the Sparsest-Cut objective. The methods themselves are highly flexible and apply to other variations of Minimum-Conductance and Sparsest-Cut. For the sake of simplicity, we will describe them for conductance. All of the following procedures adopt the following meta-algorithm starting with working set $W$ initialized to an input reference set of nodes $R$, values $\alpha_1, \beta_1, \gamma_1$ and vectors $h = d_R$, $g=d$, where $d_R$ is a vector of length $|\mathcal{V}|$ with components equal to $d_i$'s for nodes $v_i\in R$ and zeros for nodes $v_i\in R^c$. Figure~\ref{fig:cut-graph} illustrates the construction of an augmented graph based on the previous setting of $\alpha,\beta,\gamma$ and $h,g$. \noindent \textbf{\footnotesize The Local-Flow Meta-algorithm} \begin{compactenum} \footnotesize \item Initialize $W_1 = R$, $k = 1$, $h=d_R$, $g=d$ and $\alpha_1, \beta_1, \gamma_1$ based on $R$. \item Create the reference cut graph $\tilde{B}_k, \tilde{C}_k$ based on $R$ and $\alpha_k, \beta_k, \gamma_k$. \item Solve the $s,t$-Min-Cut problem associated with $\tilde{B}_k,\tilde{C}_k$ \item Set $W_{k+1}$ to be the $s$-side of the cut \item\label{step:cond-or-variant} Check if $W_{k+1}$ has smaller conductance (or some variant of it, as we will make specific in the text below) than before and stop if not and return $W_k$ \item Update $\alpha_{k+1}, \beta_{k+1}, \gamma_{k+1}$ based on $W_{k+1}$ and $R$ \item Set $k \to k+1$ \item Repeat starting from state 2 \end{compactenum} \begin{figure} \subfigure[A simple graph]{\includegraphics[width=0.3\linewidth]{cut-graph-init}} \hfill \subfigure[Adding the source $s$ and sink $t$]{\includegraphics[width=0.3\linewidth]{cut-graph-1}} \hfill \subfigure[The reference cut graph, with weights indicated]{\includegraphics[width=0.3\linewidth]{cut-graph}} \caption{The construction of the reference cut graph begins by adding a source node $s$ connected to the reference set $R$ and a sink node $t$ connected to the rest of the graph. Then we add weights to the network based on the degrees and three parameters $\alpha, \beta, $ and $\gamma$. Each edge to the source node is weighted by $\alpha \cdot \text{degree}$, each edge to the sink node is weighted by $\beta \cdot \text{degree}$, and each internal edge is weighted by $\gamma$. Note that one of the choices of $\alpha, \beta, $ or $\gamma$ will be 1, but various papers adopt different choices, and so we leave it general.} \label{fig:cut-graph} \end{figure} Next, we describe several procedures that are instantiations of this basic Local-Flow Meta-algorithm. \paragraph{MQI} The first algorithm we consider is the MQI procedure due to Lang and Rao~\cite{kevin04mqi}. This method is designed to take the reference set $R$ with $\vol(R) \le \vol(G)/2$ and identify a subset of it $S \subseteq R$. The method instantiates the Local-Flow Meta-algorithm using $\alpha_k = \cut(W_k), \gamma_k = \vol(W_k), \beta_k = \infty$ and $h=d_R$, $g=d$.\ The idea with this method is that the reference cut graph will have an $s,t$-Min-Cut value strictly less than $\alpha_k \gamma_k$ if and only if there is a strict subset $S \subset W_k$ that has conductances less than $\alpha_k / \gamma_k$. (See~\cite{kevin04mqi} for the proof.) If there is such a set, then the result $W_{k+1}$ will be a set with conductance less than $\alpha_k / \gamma_k$. Since $\alpha_k$ and $\gamma_k$ are picked based on the current working set $W_k$, at each step the algorithm monotonically improves the conductance. Also, each step minimizes the objective \begin{align*} \mbox{minimize} & \; \| \tilde{B}x \|_{1,\tilde{C}_k} \label{eq:mqi-l1} \\ \mbox{subject to} & \; x_i = 0 \ \forall v_i\in R^c, x_s = 1, x_t = 0. \end{align*} In fact, when the algorithm terminates, that means that there is no subset of $R$ with conductance less than $\alpha_k / \gamma_k$. Hence, we have solved the following variation on the conductance problem \begin{align*} \mbox{minimize} & \; \phi(S)= \frac{\cut(S)}{\vol(S)} \\ \mbox{subject to} & \; S \subseteq R. \end{align*} The key difference from the standard conductance problem is that we have restricted ourselves to a \emph{subset} of the reference set $R$. Procedurally, this is guaranteed because the edges connecting $R^c$ to $t$ have weight infinity, so they will never be cut. Thus, operationally, MQI is always a strongly local algorithm since it only operates within the input seed set $R$. Nodes connected to $t$ with weight infinity can be agglomerated or merged into a mega-sink node $T$. The resulting graph has the same size as $R$ along with the source and sink. (This is how the MQI construction is described in the original paper.) The MQI problem can be solved using Max-Flow method on the resulting graph a logarithmic number of times \cite{kevin04mqi}. Therefore, the running time for solving MQI depends on the Max-Flow algorithm that is used. Details about running times of Max-Flow algorithms can be found in~\cite{GT14}. \paragraph{Flow-Improve} The Flow-Improve method due to Andersen and Lang~\cite{AL08_SODA} was inspired by MQI and designed to address the weakness that the algorithm will always find an output set within the reference set $R$, i.e., that is a subset of $R$. (As an illustration, see Figure~\ref{fig:mqi-vs-flow-improve}.) Again, Flow-Improve takes as input a reference set $R$ with volume less than half the graph. The idea behind Flow-Improve is that we want to find a set with conductance at least as good as $R$ and that also is highly correlated with $R$. To do this, consider the following variant of conductance \[ \phi_R(S) = \frac{\cut(S)}{\vol(S \cap R) - \theta \vol(S \cap R^c)} \] where $\theta = \vol(R)/\vol(R^c)$, and where the value is $\infty$ if the denominator is negative. For any set $S$, $\phi_R(S) \ge \phi(S)$. Thus, this modified conductance score is an upper-bound on the true conductance. Again, we are able to show that the Local-Flow Meta-algorithm can solve for the exact value of $\phi_R(S)$ in polynomial time. To do so, instantiate that algorithm with $\alpha_k = \phi_R(W_k)$, $\beta_k = \theta \phi_R(W_k)$, $\gamma_k = 1$ and $h=d_R$, $g=d$.\ The value of a cut on set $S$ in the resulting graph is \[ \cut(S) + \alpha_k \vol(R) - \alpha_k [\vol(S \cap R) - \theta \vol(S \cap R^c)]. \] (See~\cite{AL08} for the justification.) Andersen and Lang show that the algorithm monotonically reduces $\phi_R(W_{k})$ at each iteration as well. Each iteration now solves the $s,t$-Min-Cut problem \begin{align} \mbox{minimize} & \; \| \tilde{B}x \|_{1,\tilde{C}_k} \label{eq:FlowImprove} \\ \mbox{subject to} & \; x_s = 1, x_t = 0. \nonumber \end{align} In order to match precisely their Flow-Improve procedure, we would need to modify our Meta-algorithm to check the value of $\phi_R(W_k)$, instead of conductance (at Step~\ref{step:cond-or-variant} of the Local-Flow Meta-algorithm above), for monotonic decrease. The authors also show that this procedure terminates in a finite number of iterations. At termination, the Flow-Improve algorithm has exactly solved $\mbox{minimize} \; \phi_R(S), S \subseteq V$. This can be considered a locally-biased variation of the conductance objective, where we penalize departure from the reference set $R$. Consequently, the solutions will tend to identify small conductance sets nearby $R$. Flow-Improve is a very useful algorithm, but it has two small weaknesses. The first is that it is a weakly local algorithm. At each step, we have to solve a Min-Cut problem that is the size of the original graph. The second is that the Min-Cut problems do not have integer weights. (Note that $\theta$ will almost never be an integer.) Most fast Max-Flow/Min-Cut procedures and implementations assume integer weights. For instance, many implementations of the push-relabel method (hipr~\cite{goldberg95_push}) only allows integer weights. Boykov and Kolmogorov's solver is a notable exception~\cite{Boykov-2004-maxflow}. Similarly to MQI, the running time of solving the Max-Flow/Min-Cut problem depends on the particular solver that is used. A summary of Max-Flow/Min-Cut methods can be found in \cite{GT14}. \begin{figure} \centering \subfigure[MQI]{\includegraphics[width=0.30\linewidth,trim=80mm 33mm 00mm 57mm,clip]{U3A-mqi-spectral}} \subfigure[Flow-Improve]{\includegraphics[width=0.30\linewidth,trim=80mm 33mm 00mm 57mm,clip]{U3A-flow-improve-spectral}} \subfigure[The difference]{\includegraphics[width=0.30\linewidth,trim=80mm 33mm 00mm 57mm,clip]{U3A-flow-improve-mqi-diff}} \caption{The results of running MQI and Flow-Improve on the reference set produced by a spectral partitioning method on the geometric graph (i.e., run global spectral \S\ref{subsec:spcut}; then round with a sweep-cut \S\ref{subsec:rounding}; and then refine using MQI and Flow-Improve). The Flow-Improve method identifies the optimal set from Figure~\ref{fig:small-cluster} in this case, whereas MQI cannot because it searches only within the given set $R$.} \label{fig:mqi-vs-flow-improve} \end{figure} \paragraph{Local-Flow-Improve} The Local-Flow-Improve algorithm due to Orecchia and Zhu~\cite{OZ14} sought to address the weak-locality of the Flow-Improve method and create a strongly local flow based method. This involved two key innovations: a modification to the construction and objective that enables strong locality; and an algorithm to realize that strong locality. This Local-Flow-Improve method essentially interpolates between MQI and Flow-Improve. In one limit, it is strictly local to the reference graph and exactly reproduces the MQI output. In the other limit, it is exactly Flow-Improve. To do this, Orecchia and Zhu alter the objective function used for Flow-Improve to place an additional penalty on deviating from the set $R$. They describe this as increasing the weight of connections $\beta_k$ in the reference cut graph by scaling these by a value $\kappa \ge 1$. If $\kappa = 1$, then their construction is exactly that of Flow-Improve. If $\kappa = \infty$, then this construction is equivalent to that of MQI. The effect of $\kappa$ is illustrated in Figure~\ref{fig:local-flow-improve}. In terms of the optimization framework, their modification corresponds to using \[ \phi_R'(S;\kappa) = \frac{\cut(S)}{\vol(S \cap R) - \theta \kappa \vol(S \cap R^c)} \ \] where $\kappa \ge 1$ and $\theta = \vol(R)/\vol(R^c)$ as in Flow-Improve, and again the value is $\infty$ if the denominator is negative. This result corresponds to instantiating the Local-Flow Meta-algorithm using $\alpha_k = \phi_R'(W;\kappa)$, $\beta_k = \phi_R'(W;\kappa) \theta \kappa$ and $h=d_R$, $g=d$. The second innovation is that they describe an algorithm to solve the Min-Cut problem on the reference cut graph that does not need to explore the entire graph. This second piece used a novel modification of Dinic's procedure~\cite{D70} to compute a Max-Flow/Min-Cut that exploited the locality. We refer interested readers back to Orecchia and Zhu for details of this second somewhat complicated construction. In our recent work~\cite{Veldt-preprint-simple-local-flow}, however, we describe a simplified framework for the Local-Flow-Improve method that shows that the strong locality in their modification results from implicitly regularizing the Flow-Improve objective with a $\ell_1$ norm regularizer. (This will mirror strongly local spectral results in the forthcoming spectral section.) In fact, our recent work~\cite{Veldt-preprint-simple-local-flow} shows that each iteration exactly solves \begin{align} \mbox{minimize} & \; \| \tilde{B} x\|_{1,\tilde{C}_k'} + \varepsilon \| D x \|_1 \label{eq:FlowImprove-l1} \\ \mbox{subject to} & \; x_s = 1, x_t = 0, \nonumber \end{align} where $\tilde{C}_k'$ is a small perturbation on the above definition and $\varepsilon$ is a locality parameter. The volume of the output cluster $S$ of the method in \cite{Veldt-preprint-simple-local-flow} is bounded $\vol(S) \le \vol(R)(1+2/\epsilon) + E(R,R^c)$, where $\epsilon:= \vol(R)/\vol(R^c) + \delta$ and $\delta\ge 0$ is a constant. That work also describes a simple procedure to realize the strong locality that leverages any Max-Flow/Min-Cut solver on a sequence of sub-problems whose size is bounded independent of the graph. \begin{figure} \subfigure[Reference set $R$.]{\includegraphics[trim=0 60mm 90mm 10mm, clip,width=0.24\linewidth]{U3A-flow-improve-1-R}} \subfigure[The Flow-Improve result.]{\includegraphics[trim=0 60mm 90mm 10mm, clip,width=0.24\linewidth]{U3A-flow-improve-1-0}} \subfigure[Local-Flow-Improve $\kappa \!= \! e^3$]{\includegraphics[trim=0 60mm 90mm 10mm, clip,width=0.24\linewidth]{U3A-flow-improve-1-3}} \subfigure[Local-Flow-Improve $\kappa \!= \! e^5$]{\includegraphics[trim=0 60mm 90mm 10mm, clip,width=0.24\linewidth]{U3A-flow-improve-1-5}} \caption{The results of running Flow-Improve compared with Local-Flow-Improve with a reference set $R$. The Flow-Improve result returns a fairly large set whereas the Local-Flow-Improve results produce successively smaller sets as the penalty $\kappa$ increases. When $\kappa = e^5$ then the result simply fills in the hole in the reference~set.} \label{fig:local-flow-improve} \end{figure} \subsection{Weakly-and-Strongly local spectral methods} There are spectral analogues for each of the three flow-based constructions on the augmented graph. The development of these ideas occurred in parallel, largely independently, and it was not obvious until recently that the ideas were very related. Here, we make these connections explicit. Of the three flow constructions, the simplest is the MQI objective. We begin with it. \paragraph{SpectralMQI} The method we call SpectralMQI was proposed as the Dirichlet partitioning problem in~\cite{Chung07_localcutsLAA}. Given a graph $\mathcal{G}$ and a subset of vertices, consider finding the set $S$ of minimal local conductance $\phi'(S) = \cut(S)/\vol(S)$ such that $S \subseteq R$, where, again, $R$ is a reference set specified in the input. Note that the only difference from conductance is that we don't have the minimum in the denominator. A spectral algorithm to find an approximate minimizer of this is to solve the generalized eigenvector problem \begin{align*} \lambda_R = \mbox{minimize} & \quad \frac{\|Bx\|_{2,C}^2}{\|x\|_{2,D}^2} \\ \mbox{subject to} &\quad x_{i} = 0 \ \forall v_i \in R^c. \end{align*} The solution vector $x$ and value $\lambda_R$ are related to the smallest eigenvalue of the sub-matrix of the normalized Laplacian corresponding to the nodes in $R$. (Note that we take the sub-matrix of the normalized Laplacian, rather than the normalized Laplacian on the sub-graph induced by $R$.) A sweep-cut over the eigenvector $x$ produces a set $S$ that satisfies a Cheeger inequality with respect to the best possible solution~\cite{Chung07_localcutsLAA}. This definition of local conductance is also called $\texttt{NCut'}$ by Hochbaum~\cite{Hochbaum-2010-ncutp}, who gave a polynomial time algorithm to compute it that is closely related to the MQI procedure. In this case, if $R$ has volume less than half the graph, then this is exactly the spectral analogue of the MQI procedure and the result is a Cheeger-like bound on the overall conductance. \paragraph{MOV} The Mahoney-Orecchia-Vishnoi (MOV) objective~\cite{MOV12_JMLR} is a spectral analogue of the FlowImprove method, with a few subtle differences. The goal is to find a small Rayleigh quotient, as in \eqref{spcutLD}, that is highly correlated with an input vector $z\in\mathbb{R}^{|V|}$, where $z$ represents the seed or local-bias. Given this, the MOV objective is \begin{align*} \mbox{minimize} & \ \frac{\|Bx\|_{2,C}^2}{\|x\|_{2,D}^2} \\\nonumber \mbox{subject to} &\ 1_{|\mathcal{V}|}^T D x = 0 \\\nonumber &\ (z^T D x)^2 \ge \kappa &\ x\in\mathbb{R}^{\mathcal{V}}. \end{align*} The solution of this problem represents an embedding of the nodes in $\mathcal{V}$ which is locally biased, i.e., large values for components/nodes that are considered important and small or zero values for the rest. According to \cite{MOV12_JMLR}, there is a constant $\rho$, i.e., the optimal dual variable for the locally-biased constraint, such that the solution to the MOV problem satisfies $ (L+\rho D) x = \rho D z$.\ The null space of $L$ is the vector $1_{|\mathcal{V}|}$, and assuming that $1_{|\mathcal{V}|}Dz = 0$, then the solution to the previous system is unique. One final detail is that the MOV construction fixes $\| x \|_2 = 1$. Consequently, the MOV solution is \begin{equation} \label{eq:mov} x = c (L + \rho D)^\dagger D z \quad c = (\| (L + \rho D)^\dagger D z \|_2)^{-1}. \end{equation} In the MOV paper~\cite{MOV12_JMLR}, they show that $\rho$ can be chosen such that $x^T D z = \kappa$, the desired correlation strength with the input vector $z$, through a simple bisection procedure. Solving the linear system \eqref{eq:mov} results in a weakly-local method that satisfies another Cheeger-like inequality. Recent extensions show that it is possible to get multiple locally-biased vectors that are akin to eigenvectors from this setup~\cite{HM14_JRNL,LBM16_TR}. The methodology is able to leverage the large number of Laplacian system solvers \cite{V12} that can find an approximate solution to \eqref{eq:mov} in nearly linear time. The pseudo-inverse allows us to ``pass through'' $\rho = 0$ and approach $\rho = -\lambda_2$. (This system is singular at $\rho = 0$ and $\rho = \lambda_2$.) What is interesting is that taking the limit $\rho \to -\lambda_2$ directly maps to the spectral relaxation~\eqref{spcutLD}. Thus, the $\rho$ parameter interpolates between the global spectral relaxation~\eqref{spcutLD} and a spectral-version of the Min-Cut problem in each step of FlowImprove. Based on the reference cut graph, the MOV objective is $\mbox{minimize} \ \|\tilde{B}\tilde{x}\|_{2,\tilde{C}}^2$, where $\tilde{x} : = [1; x; 0]$. The reference graph cut setting is $\gamma=1$, $g=d$, $h=Dz$ and $\alpha=\beta=\rho\ge0$ controls the desired strength of the correlation to the input vector $z$.\ Notice that the MOV problem is a spectral, i.e., $\ell_2$, version of the $s,t$-Min-Cut problem. This observation was made first in \cite{GM14_ICML,GM15_KDD}. If $\rho$ is extremely large, the solution to the above problem would have perfect correlation with the input vector $h$. As $\rho \to 0$, we decrease the effective correlation with the input vector $h$. (These arguments are formal in~\cite{MOV12_JMLR}.) \section{Preliminaries and Notation} \label{sec:prelim} \textbf{Graph assumptions:} We use the letter $\mathcal{G}$ to denote a given connected graph. We assume that $\mathcal{G}$ is undirected with no self-loops. Many of the constructions we will use operate on \emph{weighted graphs} and so we assume that each edge may have a positive capacity. Graphs that are unweighted should have all of their capacities set to $1$. \textbf{Nodes, edges and cuts:} Let $\mathcal{V} = \{v_1,v_2\dots,v_n\}$ be a given set of $|\mathcal{V}|$ nodes of graph $\mathcal{G}$. We denote with $e_{ij}$ an edge in the graph between nodes $v_i$ and $v_j$. Let $\mathcal{E}$ be a given set of $|\mathcal{E}|$ edges of graph $\mathcal{G}$. A subset $S\subset\mathcal{V}$ of nodes can be used to define a partitioning of $\mathcal{V}$ into $S$ and $S^c:=\mathcal{V} \backslash S$. We define a cut as subset $E \subset \mathcal{E}$ which partitions the graph $\mathcal{G}$ into two sets. Given a partition $S\subset\mathcal{V}$ and $S^c$, then $E(S,S^c)=\{e_{ij}\in \mathcal{E} \ | \ v_i\in S \mbox{ and } v_j\in S^c\}$ is the set of edges with one side in $S$ and the other side in $S^c$. If the partition is clear from the context we write the cut set as $E$ instead of $E(S,S^c)$.\ Let $c_{ij}$ be a weight of the edge $e_{ij}$, then we define the cut $S$ as \begin{equation} \cut(S) := \cut(E(S,S^c)) := \sum_{e_{ij}\in E(S,S^c)} c_{ij}. \end{equation} The volume of a set $S$ is \begin{equation} \vol(S) := \sum_{v_i \in S} \sum_{e_{ij} \in \mathcal{E}} c_{ij}. \end{equation} For simplicity of notation, we will drop the input $\mathcal{G}$ in $\mathcal{V}$ and $\mathcal{E}$ if it is clear from the context that we are referring to a single graph $\mathcal{G}$. \textbf{Matrix notation:} We denote with $A\in\mathbb{R}^{|\mathcal{V}|\times |\mathcal{V}|}$ the adjacency matrix for a given graph, where $A_{ij} = c_{ij}$ $\forall e_{ij}\in \mathcal{E}$ and zero elsewhere. Let $d_i$ be the degree of node $v_i\in\mathcal{V}$, $D\in\mathbb{R}^{|\mathcal{V}|\times |\mathcal{V}|}$ be the degree diagonal matrix $D_{ii} = d_i$, $L=D-A$ be the graph Laplacian, and $\mathcal{L} = D^{-1/2}LD^{-1/2}$ be the symmetric normalized graph Laplacian. Note that the volume of a subset $S$ is $\vol(S) = \sum_{v_i\in S} d_{i}$. We denote with $B\in\mathbb{R}^{|\mathcal{E}|\times |\mathcal{V}|}$ the incidence matrix of the given graph $\mathcal{G}$. Every row of the incidence matrix corresponds to an edge $e_{ij}\in \mathcal{E}$ in $\mathcal{G}$. Assuming arbitrary ordering of the edges of the graph, in this paper we define the rows of the incidence matrix as $B_{e_{ij}} = e_i - e_j$ $\forall e_{ij}\in \mathcal{E}$, where $e_i\in\mathbb{R}^{|\mathcal{V}|}$ is equal to one at the $i^{th}$ position and zero elsewhere. Finally, $C\in\mathbb{R}^{|\mathcal{E}|\times|\mathcal{E}|}$ is a diagonal matrix of the weights of each edge, i.e., $C_{ij} = c_{ij}$ $\forall i,j$. In this notation, the Laplacian matrix $L = B^T C B$. \textbf{Norms:} For all $x \in \mathbb{R}^{|\mathcal{E}|}$, we define the weighted $\ell_1$ and $\ell_2$ norms $\|x\|_{1,C}: = \sum_{e_{ij}\in E} c_{ij} |x_{ij}|$ and $\|x\|_{2,C}^2: = \sum_{e_{ij}\in E} c_{ij} |x_{ij}|^2$, respectively. Given a partition $S,S^c$ and a vector $x\in\{0,1\}^{|\mathcal{V}|}$ such that $x_i=1$ if $v_i\in S$ and $x_i = 0$ if $v_i\in S^c$, then $\cut(S) = \|Bx\|_{1,C}$. Moreover, notice that $\cut(S) = \|Bx\|_{2,C}^2 = x^T B^T C B x = x^T L x$. \textbf{Miscellaneous:} We use $[x;y;z]$ to denote a single column vector where the individual vectors $x,y,z$ are stacked in this order. Finally, the vectors $0_{|\mathcal{V}|}$ and $1_{|\mathcal{V}|}$ are the all zeros and ones vectors of length $|\mathcal{V}|$, respectively. \subsection{Strongly-local flow and spectral methods} \paragraph{$\ell_1$-Regularized Page-Rank} The $\ell_1$-regularized Page-Rank problem was initially studied in \cite{GM14_ICML} and then further refined in \cite{FCSRM16_TR}. In the latter work, the problem is defined as \begin{align}\label{eq:l1pr} \mbox{minimize} & \quad \frac{1}{2}\|\tilde{B}\tilde{x}\|_{\tilde{C},2}^2 + \epsilon \|Dx\|_1, \end{align} where $\tilde{x} : = [1; x; 0]$.\ The reference cut graph setting for \eqref{eq:l1pr} is $g=d$ and $h\ge 0$ is a vector that satisfies $\|h\|_1=1$ and $\|h\|_\infty \ge \epsilon$.\ The latter condition is to guarantee that the solution to \eqref{eq:l1pr} is not the zero vector. Moreover, $\alpha = \beta $ and $\gamma = (1-\alpha)/2$. Similarly to $z$ for MOV, the vector $h$ controls the input seed set and the weights of nodes in that set. The larger the weights the more the solution will be correlated with the corresponding nodes in the input seed set. The solution vector to problem \eqref{eq:l1pr} is component-wise non-negative and the parameter $\alpha$ controls how much energy is concentrated close to the input seed set. Formally, based on theoretical guarantees in \cite{ACL06} the vector $h$ should be an indicator vector for a single seed node, around which there is a target cluster of nodes $C$. The algorithm is not guaranteed to find the exact target cluster $C$, but if $C$ has conductance less than $\alpha/10$ then it is guaranteed to return a cluster with conductance of $\mathcal{O}(\sqrt{\alpha \log(\vol(C))})$. We refer the reader to \cite{ACL06} for a detailed description of the theoretical graph clustering guarantees. The idea of $\ell_1$-regularized Page-Rank graph clustering initially appeared in \cite{ACL06} in the form of implicit regularization. In particular, the authors in \cite{ACL06} suggest solving a personalized Page-Rank linear system approximately. In \cite{FCSRM16_TR,GM14_ICML}, the authors noticed that the termination criteria in \cite{ACL06} are related to the first-order optimality conditions of the above $\ell_1$-regularized Page-Rank problem, and they draw the connection to explicit $\ell_1$ regularization. It is shown in \cite{FCSRM16_TR} that solving the $\ell_1$-regularized Page-Rank problem has the same Cheeger-like worst-case approximation guarantees to the Minimum-Conductance problem as the original procedure in \cite{ACL06}. However, there is an important technical difference: one advantage of solving the $\ell_1$-regularized problem is that the locality of the solution is a property of the optimization problem as opposed to a property of an algorithm. In particular, by solving the $\ell_1$-regularized problem it is guaranteed to obtain the same solution regardless of the algorithm used. In comparison, applying the procedure in \cite{ACL06}, where the output depends on the setting of the procedure, i.e., the strategy for choosing nodes to be updated at every iteration, leads to somewhat different solutions, depending on the specific settings chosen. Let $x^*$ be the optimal solution of \eqref{eq:l1pr} and $S^*$ be the set of nodes where $x^*$ is non-zero. In \cite{FCSRM16_TR}, it is shown that many standard optimization algorithms such as iterative soft-thresholding or block iterative soft-thresholding solve \eqref{eq:l1pr} with running time $\mathcal{O}(\vol(S^*)/\alpha)$, which can be independent of the volume of the whole graph $\vol(\mathcal{V})$. This opens up the possibility of the use of these algorithms more generally. For details about the algorithms, we refer the reader to~\cite{FCSRM16_TR}.
62b8ebb950ff8629269944ee195d517b04ffbd9c
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{\@startsection {section}{1}{\z@}{3.25ex plus 1ex minus .2ex}{1.5ex plus .2ex}{\large\bf}} \def\subsection{\@startsection{subsection}{2}{\z@}{3.25ex plus 1ex minus .2ex}{1.5ex plus .2ex}{\normalsize\bf}} \@addtoreset{equation}{section} \makeatother \title{ Prohorov-type local limit theorems on\\ abstract Wiener spaces } \author{Alberto Lanconelli\footnote{Dipartimento di Matematica, Universit\'a degli Studi di Bari Aldo Moro, Via E. Orabona 4, 70125 Bari - Italia. E-mail: \emph{[email protected]}}} \date{\empty} \begin{document} \maketitle \numberwithin{equation}{section} \bigskip \begin{abstract} We prove that the density of $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}$, where $\{X_n\}_{n\geq 1}$ is a sequence of independent and identically distributed random variables taking values on a abstract Wiener space, converges in $\mathcal{L}^1$ to the density of a certain Gaussian measure which is absolutely continuous with respect to the reference Wiener measure. The crucial feature in our investigation is that we do not require the covariance structure of $\{X_n\}_{n\geq 1}$ to coincide with the one of the Wiener measure. This produces a non trivial (different from the constant function one) limiting object which reflects the different covariance structures involved. The present paper generalizes the results proved in \cite{LS 2016} and deepens the connection between local limit theorems on (infinite dimensional) Gaussian spaces and some key tools from the Analysis on the Wiener space, like the Wiener-It\^o chaos decomposition, Ornstein-Uhlenbeck semigroup and Wick product. We also verify and discuss our main assumptions on some examples arising from the applications: dimension independent Berry-Esseen-type bounds and weak solutions of stochastic differential equations. \end{abstract} Key words and phrases: local limit theorems, abstract Wiener spaces, Wick product, Ornstein-Uhlenbeck semigroup \\ AMS 2000 classification: 60F25, 60G15, 60H07 \allowdisplaybreaks \section{Introduction and statement of the main result} Local limit theorems are central limit theorems for densities. Given a sequence $\{X_n\}_{n\geq 1}$ of independent and identically distributed random variables, one aims to prove that the density of the standardized sum $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{nVar(X_1)}}$ converges in some sense to the standard normal density function. This problem has attracted the attention of several authors: we recall the classical papers by Prohorov \cite{Prohorov}, which is concerned with convergence in $\mathcal{L}^1$, Gnedenko \cite{Gnedenko}, who studies uniform convergence and Ranga Rao and Varadarajan \cite{Ranga Rao Varadarajan}, where point-wise convergence is investigated. All these three classical results rely on Fourier transform techniques applied to convolutions of densities. More recently, Barron \cite{Barron} proved that the relative entropy, also known as Kullback-Leibler divergence, of the law of $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{nVar(X_1)}}$ with respect to the standard Gaussian measure tends to zero, monotonically along a certain subsequence. This type of convergence is stronger than the one considered by Prohorov \cite{Prohorov} and hence improves his result. \\ When the sequence $\{X_n\}_{n\geq 1}$ takes values on an infinite dimensional space, then the validity of general local limit theorems is not guaranteed. In fact, Bloznelis \cite{Blozenis} has shown a counterexample to the validity of Prohorov's theorem on general Hilbert spaces. We also mention the paper by Davydov \cite{Davydov}, where a variant of a local limit theorem for Banach space valued random variables is proposed.\\ The aim of the present paper is to prove a version of Prohorov's theorem for an abstract Wiener space valued random sequence. More precisely, we show that the density of $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}$, where $\{X_n\}_{n\geq 1}$ is a sequence of independent and identically distributed random variables taking values on an abstract Wiener space, converges in $\mathcal{L}^1$ to the density of a certain Gaussian measure which is absolutely continuous with respect to the reference Wiener measure. The choice of working in a space endowed with a Gaussian measure, which is the limiting object of the central/local limit theorem, brings important advantages: the roles of scaling operator and convolution product are naturally played by the Ornstein-Uhlenbeck semigroup and Wick product, respectively (see Theorem \ref{wick product convolution} below). This fact together with certain norm inequalities proved in previous papers (see Theorem \ref{Young inequality} below) become the key ingredients for the proof of our local limit theorem. Furthermore, we do not require the covariance operator of the $X_n$'s to be the identity, which is the covariance of the Wiener measure; we set three standing assumptions which allow for a richer family of Gaussian measures as limiting objects.\\ \noindent To be more specific, we recall that $(H,W,\mu)$ is an \emph{abstract Wiener space} if $(H,\langle\cdot,\cdot\rangle_H)$ is a separable Hilbert space, which is continuously and densely embedded in the Banach space $(W,|\cdot|_W)$, and $\mu$ is a Gaussian probability measure on the Borel sets of $W$ such that \begin{eqnarray}\label{Gaussian characteristic} \int_{W}e^{i\langle w,\varphi\rangle}d\mu(w)=e^{-\frac{1}{2}|\varphi|_H^2},\quad\mbox{ for all }\varphi\in W^*. \end{eqnarray} Here $W^*\subset H$ denotes the dual space of $W$, which in turn is dense in $H$, and $\langle\cdot,\cdot\rangle$ stands for the dual pairing between $W$ and $W^*$. We will refer to $H$ as the \emph{Cameron-Martin} space of $W$. Observe that $\langle w,\varphi\rangle=\langle w,\varphi\rangle_H$ if $w\in H$ and $\varphi\in W^*$. Among the most important examples of abstract Wiener spaces, we recall: the Euclidean space $\mathbb{R}^d$ endowed with the standard $d$-dimensional Gaussian measure (in this case we have $W=H=\mathbb{R}^d$); the classical Wiener space $C_0([0,1])$ of continuous functions starting at zero endowed with the classical Wiener measure (in this case $W=C_0([0,1])$ and $H=H_0^1([0,1])$). In the sequel we denote by $\Vert\cdot\Vert_p$ the norm in the space $\mathcal{L}^p(W,\mu)$ for $p\geq 1$.\\ We now describe our set of assumptions. Let $\{X_n\}_{n\geq 1}$ be a sequence of independent and identically distributed random variables taking values on $W$. \begin{assumption}\label{square integrable density} The law of the $X_n$'s is absolutely continuous with respect to $\mu$ with a density $f$ belonging to $\mathcal{L}^2(W,\mu)$ \end{assumption} \noindent Assumption \ref{square integrable density} has several important implications. First of all, it yields the finiteness of all the moments of the scalar random variable $\langle X_n,\varphi\rangle$ for $\varphi\in W^*$. In fact, for any $m\in\mathbb{N}$ and $\varphi\in W^*$ a simple application of the Cauchy-Schwartz inequality gives \begin{eqnarray*} E[|\langle X_n,\varphi\rangle|^m]&=&\int_W|\langle w,\varphi\rangle|^md\mu_{X_n}(w)\\ &=&\int_W|\langle w,\varphi\rangle|^m f(w)d\mu(w)\\ &\leq&\Big(\int_W|\langle w,\varphi\rangle|^{2m} d\mu(w)\Big)^{\frac{1}{2}}\cdot\Big(\int_W f^2(w)d\mu(w)\Big)^{\frac{1}{2}}\\ &=&\sqrt{(2m-1)!!}|\varphi|_H^{m}\Big(\int_W f^2(w)d\mu(w)\Big)^{\frac{1}{2}} \end{eqnarray*} entailing the finiteness of the moments. Moreover, taking $m=2$ in the previous inequality, we see that if $\{\varphi_j\}_{j\geq 1}$ is a sequence in $W^*$ converging in the norm of $H$ to $h\in H$, then \begin{eqnarray*} E[|\langle X_n,\varphi_j\rangle-\langle X_n,\varphi_i\rangle|^2]&=&E[|\langle X_n,\varphi_j-\varphi_i\rangle|^2]\\ &\leq&\sqrt{3}|\varphi_j-\varphi_i|_H^2\Big(\int_W f^2(w)d\mu(w)\Big)^{\frac{1}{2}}. \end{eqnarray*} Therefore, $\{\langle X_n,\varphi_j\rangle\}_{j\geq 1}$ turns out to be a Cauchy sequence in $\mathcal{L}^2(W,\mu)$ and we can define $\langle X_n,h\rangle$ almost surely as the limit of this sequence.\\ Another consequence of Assumption \ref{square integrable density} is that, according to the Wiener-It\^o chaos decomposition theorem, the density $f$ can be represented as $\sum_{k\geq 0}\delta^k(f_k)$, where \begin{eqnarray*} \delta^0(f_0)=\int_Wf(w)d\mu(w)=1 \end{eqnarray*} and for $k\geq 1$, $f_k\in H^{\hat{\otimes}k}$, the space of symmetric elements of $H^{\otimes k}$, while $\delta^k(f_k)$ stands for the multiple Wiener-It\^o integral of $f_n$. Hence, we can write \begin{eqnarray*} Var(\langle X_n, h\rangle)&=&E[\langle X_n, h\rangle^2]-E[\langle X_n, h\rangle]^2\nonumber\\ &=&\int_W\langle w, h\rangle^2f(w)d\mu(w)-\Big(\int_W\langle w, h\rangle f(w)d\mu(w)\Big)^2\nonumber\\ &=&\int_W(\delta^2(h^{\otimes 2})+|h|_H^2)f(w)d\mu(w)-\Big(\int_W\langle w, h\rangle f(w)d\mu(w)\Big)^2\nonumber\\ &=&\int_W\delta^2(h^{\otimes 2})f(w)d\mu(w)+|h|_H^2-\Big(\int_W\langle w, h\rangle f(w)d\mu(w)\Big)^2\nonumber\\ &=&2\langle f_2,h^{\otimes 2}\rangle_{H^{\otimes 2}}-\langle f_1,h\rangle_H^2+|h|_H^2\nonumber\\ &=&\langle 2f_2-f_1^{\otimes 2},h^{\otimes 2}\rangle_{H^{\otimes 2}}+|h|_H^2. \end{eqnarray*} This gives the identity \begin{eqnarray}\label{variance chaos expansion} Var(\langle X_n, h\rangle)&=&\langle 2f_2-f_1^{\otimes 2},h^{\otimes 2}\rangle_{H^{\otimes 2}}+|h|_H^2 \end{eqnarray} that will play a crucial role in the sequel. Our next assumption concerns the behaviour of $Var(\langle X_n,h\rangle)$ for $h\in H$. \begin{assumption}\label{positive trace class} For every $h\in H$, \begin{eqnarray}\label{positive} Var(\langle X_n,h\rangle)\geq |h|_H^2 \end{eqnarray} and for some orthonormal basis $\{e_j\}_{j\geq 1}$ of $H$, \begin{eqnarray}\label{trace class} \sum_{j\geq 1}[Var(\langle X_n,e_j\rangle)-1]<+\infty. \end{eqnarray} \end{assumption} \noindent Observe that, according to equation (\ref{variance chaos expansion}), inequality (\ref{positive}) is equivalent to the positivity of $2f_2-f_1^{\otimes 2}$; in fact for all $h\in H$ we have \begin{eqnarray*} \langle 2f_2-f_1^{\otimes 2},h^{\otimes 2}\rangle_{H^{\otimes 2}}&=&Var(\langle X_n, h\rangle)-|h|_H^2. \end{eqnarray*} Moreover, inequality (\ref{trace class}) ensures that $2f_2-f_1^{\otimes 2}$ is of trace class since \begin{eqnarray*} \sum_{j\geq 1}\langle 2f_2-f_1^{\otimes 2},e_j^{\otimes 2}\rangle_{H^{\otimes 2}}&=&\sum_{j\geq 1}[Var(\langle X_n, e_j\rangle)-1]. \end{eqnarray*} \noindent Our last assumption is the following. \begin{assumption}\label{integrable quadratic exponential} For some orthonormal basis $\{e_j\}_{j\geq 1}$ of $H$, \begin{eqnarray}\label{series covariance} \sum_{i,j\geq 1}[Cov(\langle X_n,e_i\rangle,\langle X_n,e_j\rangle)-\delta_{ij}]^2<1. \end{eqnarray} \end{assumption} \noindent This is a more technical condition: it serves to guarantee that the limiting Gaussian measure in our local limit theorem possesses a square integrable density with respect to the reference Wiener measure (see Proposition \ref{properties quadratic exp} below). Note that the left hand side in (\ref{series covariance}) corresponds to $|2f_2-f_1^{\otimes 2}|^2_{H^{\otimes 2}}$; in fact, \begin{eqnarray*} |2f_2-f_1^{\otimes 2}|^2_{H^{\otimes 2}}&=&\sum_{i,j\geq 1}\langle 2f_2-f_1^{\otimes 2},e_i\otimes e_j\rangle_{H^{\otimes 2}}^2\\ &=&\sum_{i,j\geq 1}\langle 2f_2-f_1^{\otimes 2},e_i\hat{\otimes}e_j\rangle_{H^{\otimes 2}}^2\\ &=&\sum_{i,j\geq 1}\Big[\int_W\delta^2(e_i\hat{\otimes}e_j)f(w)d\mu(w)\\ &&-\int_W\langle w,e_i\rangle f(w)d\mu(w)\cdot\int_W\langle w,e_j\rangle f(w)d\mu(w)\Big]^2\\ &=&\sum_{i,j\geq 1}\Big[\int_W\langle w,e_i\rangle\cdot\langle w,e_j\rangle f(w)d\mu(w)-\langle e_i,e_j\rangle_H\\ &&-\int_W\langle w,e_i\rangle f(w)d\mu(w)\cdot\int_W\langle w,e_j\rangle f(w)d\mu(w)\Big]^2\\ &=&\sum_{i,j\geq 1}[Cov(\langle X_n,e_i\rangle,\langle X_n,e_j\rangle)-\delta_{ij}]^2. \end{eqnarray*} Hence, Assumption \ref{integrable quadratic exponential} is equivalent to \begin{eqnarray*} \Big|f_2-\frac{f_1^{\otimes 2}}{2}\Big|^2_{H^{\otimes 2}}<\frac{1}{4}. \end{eqnarray*} The reason for this assumption comes from the advantages of utilizing the Wiener-It\^o chaos expansion and the hyper-contractive properties of the Ornstein-Uhlenbeck semigroup (see Nelson \cite{Nelson}), which are absent in the $\mathcal{L}^1(W,\mu)$ space.\\ The next proposition is a first step towards the main result. Its proof is deferred to Section 3. \begin{proposition}\label{properties quadratic exp} Let Assumptions \ref{square integrable density}, \ref{positive trace class} and \ref{integrable quadratic exponential} be in force and set $g_2:=f_2-\frac{f_1^{\otimes 2}}{2}$, where $f_1$ and $f_2$ are the first and second order kernels in the chaos decomposition of $f$, respectively. Then, \begin{eqnarray*} \xi=\sum_{k\geq 0}\frac{\delta^{2k}(g_2^{\otimes k})}{k!} \end{eqnarray*} is the square integrable density of a Gaussian measure on $W$ with characteristic functional given by \begin{eqnarray*} \exp\Big\{-\langle g_2,h^{\otimes 2}\rangle-\frac{\Vert h\Vert_H^2}{2}\Big\}. \end{eqnarray*} Moreover, if for any $n\in\mathbb{N}$ the random variables $\mathcal{X}_1,...,\mathcal{X}_n$ are independent with law $\xi d\mu$, then \begin{eqnarray}\label{WID} \frac{\mathcal{X}_1+\cdot\cdot\cdot+\mathcal{X}_n}{\sqrt{n}}\quad\mbox{ has law }\quad\xi d\mu. \end{eqnarray} \end{proposition} We will refer to the Gaussian measure $\xi d\mu$ on $W$ as the \emph{Gaussian measure induced by} $f$. We are now ready to state the main result of the present paper. The proof is postponed to Section 4. \begin{theorem}\label{main theorem} Let $\{X_n\}_{n\geq 1}$ be a sequence of independent and identically distributed random variables taking values on $W$ and fulfilling Assumptions \ref{square integrable density}, \ref{positive trace class} and \ref{integrable quadratic exponential}. Let also $\mathcal{X}$ be a random variable on $W$ whose law is the Gaussian measure induced by $f$. Then, for any $\alpha\in ]0,1[$ the density of \begin{eqnarray}\label{limit} \sqrt{\alpha}\cdot\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}+\sqrt{1-\alpha}\cdot Z \end{eqnarray} converges in $\mathcal{L}^1(W,\mu)$ as $n$ tends to infinity to the density of \begin{eqnarray}\label{limit2} \sqrt{\alpha}\cdot\mathcal{X}+\sqrt{1-\alpha}\cdot Z \end{eqnarray} with speed of convergence of order $\frac{1}{\sqrt{n}}$. Here $Z$ is a $W$-valued random variable with law $\mu$ which is independent of the sequence $\{X_n\}_{n\geq 1}$ and of $\mathcal{X}$. \end{theorem} We note that if the covariance structure of the variables $X_n$'s coincides with the one of the Wiener measure $\mu$, i.e. $Var(\langle X,h\rangle)=|h|_H^2$, then Assumptions \ref{positive trace class} and \ref{integrable quadratic exponential} are trivially satisfied. This corresponds to the classical assumption in multidimensional central/local limit theorems. Moreover, Theorem \ref{main theorem} generalizes the result proved in \cite{LS 2016}, where it is assumed that $f_1=0$ and $f_2=0$ (from a probabilistic point of view, this is equivalent to the requirements $E[X_n]=0$ and $Var(\langle X,h\rangle)=|h|_H^2$).\\ The reason why the local limit theorem we prove concerns the sequence in (\ref{limit}) instead of $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}$, which would be the natural object of investigation, is of technical nature. In fact, observe first of all that adding the term containing $Z$ has the effect of smoothing the density of the sequence $X_n$ through the Ornstein-Uhlenbeck semigroup. Moreover, according to Theorem \ref{wick product convolution} below the Wick product behaves as a Gaussian convolution, i.e. is the necessary tool to handle densities of sum of independent random variables on Gaussian spaces. The Wick product is however an unbounded bilinear form on any $\mathcal{L}^p(W,\mu)$ space; to get something in $\mathcal{L}^p(W,\mu)$ out of a Wick product, one has to smooth the densities involved through the Ornstein-Uhlenbeck semigroup. This is contained in Theorem \ref{Young inequality} below. The point here is that the regularization required by that theorem exhausts all the infinitesimal scaling effect of the factor $\frac{1}{\sqrt{n}}$ in $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}$. Hence, we are forced to add an extra smoothing component on the density of the $X_n$'s in order to prove the theorem. We remark however that such a smoothness assumption is also required by Lynnik \cite{Lynnik} in proving an information-theoretic central limit theorem.\\ The paper is organized as follows: in Section 2 the crucial assumptions of the main theorem, Theorem \ref{main theorem}, are verified and discussed on three important examples arising from the applications: the finite dimensional case in relation to dimension independent Berry-Esseen-type bounds, the class of measures of convolution type obtained by convolving the Wiener measure with a probability measure supported on the Cameron-Martin space, the analysis of weak solutions of a family of systems of stochastic differential equations; Section 3 contains the main preparatory theorems needed for the proof of the main result: the roles of the Wick product and Ornstein-Uhlenbeck semigroup in the manipulation of probability densities on infinite dimensional Gaussian spaces, Theorem \ref{wick product convolution} and Theorem \ref{Young inequality}, and the proof of Proposition \ref{properties quadratic exp}, which describes the properties of the limiting object of our local limit theorem; in Section 4 we prove Theorem \ref{main theorem} while in the Appendix we collect for the reader's convenience several useful formulas utilized throughout the paper. \section{Examples and applications} In this section we check and discuss the assumptions of Theorem \ref{main theorem} in some concrete examples arising from the applications. \subsection{The finite dimensional case: dimension independent Berry-Esseen-type bounds} We choose $W=H=\mathbb{R}^d$ and \begin{eqnarray*} \mu_d(A)=\int_{A}(2\pi)^{-\frac{d}{2}}e^{-\frac{|x|^2}{2}}dx,\quad A\in\mathcal{B}(\mathbb{R}^d) \end{eqnarray*} where $|\cdot|$ denotes the $d$-dimensional Euclidean norm and $\mathcal{B}(\mathbb{R}^d)$ is the collection of the Borel sets of $\mathbb{R}^d$. Then, $(W,H,\mu)$ is a finite dimensional abstract Wiener space. \\ Let $\{X_n\}_{n\geq 1}$ be a sequence of independent and identically distributed $d$-dimensional random vectors; assume that the law of $X_n$ on $\mathbb{R}^d$ is absolutely continuous with respect to $\mu_d$ with a density $f\in\mathcal{L}^2(\mathbb{R}^d,\mu_d)$ (this corresponds to Assumption \ref{square integrable density}). Moreover, observe that Assumptions \ref{positive trace class} and \ref{integrable quadratic exponential} can be easily rephrased in terms of the covariance matrix of the random vector $X_n$ (condition (\ref{trace class}) is always satisfied in the finite dimensional framework). If we denote by $\mu_n$ the law of \begin{eqnarray*} \sqrt{\alpha}\cdot\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}+\sqrt{1-\alpha}\cdot Z \end{eqnarray*} and by $\mu_{\mathcal{X}}$ the law of $\sqrt{\alpha}\cdot\mathcal{X}+\sqrt{1-\alpha}\cdot Z$, then Theorem \ref{main theorem} can rewritten as \begin{eqnarray}\label{berry-esseen} d_{TV}(\mu_n,\mu_{\mathcal{X}})\leq \frac{C}{\sqrt{n}} \end{eqnarray} where $d_{TV}$ stands for the distance in total variation and $C$ is a constant depending only on $f$ and $\alpha$. Inequality (\ref{berry-esseen}) represents a dimension independent Berry-Essen type bound. These type of estimates have been investigated by many authors under very mild conditions (existence of the third moment) and with a constant $C$ in (\ref{berry-esseen}) depending on the dimension $d$. See Bentkus \cite{Bentkus} for the best known value of $C$ and the references quoted there. We stress that in the aforementioned paper it is assumed that the covariance of the vector $X_n$ is the identity matrix; if we make the same assumption, then Assumptions \ref{positive trace class} and \ref{integrable quadratic exponential} are trivially satisfied. Therefore, compared to the paper \cite{Bentkus}, the additional regularity we impose to the $X_n$'s is Assumption \ref{square integrable density} together with the introduction of the smoothing parameter $\alpha\in ]0,1[$. To conclude, our approach needs more stringent assumptions on the law of the random vector $X_n$ but provides a bound which does not depend on the dimension of the image space of the random sequence $\{X_n\}_{n\geq 1}$ (see \cite{LS 2016} for a more details on this type of comparison). \subsection{The cases $f_2-\frac{f_1^{\otimes 2}}{2}=0$ and $f_2-\frac{f_1^{\otimes 2}}{2}=g^{\otimes 2}$} To ease the notation set $g_2:=f_2-\frac{f_1^{\otimes 2}}{2}$ and assume first that $g_2=0$. This means that according to (\ref{variance chaos expansion}) we have \begin{eqnarray*} Var(\langle X_n,h\rangle)=|h|^2_H\quad h\in H. \end{eqnarray*} Hence, Assumptions \ref{positive trace class} and \ref{integrable quadratic exponential} are trivially fulfilled. Moreover, the density $\xi$ in Proposition \ref{properties quadratic exp} reduces to the constant function one and we recover a slight generalization of the result proved in \cite{LS 2016} (where it is assumed that $f_1=0$ and $f_2=0$).\\ \noindent Now assume that $g_2=g^{\otimes 2}$ for some $g\in H$. Then, $g_2$ is a positive trace class element of $H^{\hat{\otimes} 2}$, which is equivalent to Assumption \ref{positive trace class}. Furthermore, if $|g|^2_H<\frac{1}{2}$, then also Assumption \ref{integrable quadratic exponential} is satisfied. In this case the density $\xi$ from Proposition \ref{properties quadratic exp} looks like \begin{eqnarray*} \sum_{k\geq 0}\frac{\delta^{2k}(g^{\otimes 2k})}{k!}. \end{eqnarray*} The last expression can be written in the formalism of the Wick Calculus (e.g. Holden et al. \cite{HOUZ}) as \begin{eqnarray*} \exp^{\diamond}\{\delta(g)^{\diamond 2}\}\quad\mbox{ or alternatively }\quad :\exp\{\delta(g)^2\}:\quad. \end{eqnarray*} It is proved in Aase et al. \cite{AOU} that \begin{eqnarray}\label{oksendal} \exp^{\diamond}\{\delta(g)^{\diamond 2}\}=\frac{1}{\sqrt{1+2 |g|_H^2}}\exp\Big\{\frac{\delta(g)^2}{1+2 |g|_H^2}\Big\}, \end{eqnarray} provided that $2|g|_H^2<1$. If we set $\tilde{g}=\frac{g}{|g|_H}$, then equation (\ref{oksendal}) becomes \begin{eqnarray*} \exp^{\diamond}\{\delta(g)^{\diamond 2}\}&=&\frac{1}{\sqrt{1+2 |g|_H^2}}\exp\Big\{\frac{\delta(g)^2}{1+2 |g|_H^2}\Big\}\\ &=&\frac{1}{\sqrt{1+2 |g|_H^2}}\exp\Big\{\frac{2 |g|_H^2}{1+2 |g|_H^2}\frac{\delta(\tilde{g})^2}{2}\Big\}\\ &=&\frac{1}{\sqrt{1+2 |g|_H^2}}\exp\Big\{\Big(1-\frac{1}{1+2 |g|_H^2}\Big)\frac{\delta(\tilde{g})^2}{2}\Big\}\\ &=&\frac{\frac{1}{\sqrt{2\pi(1+2 |g|_H^2)}}e^{-\frac{x^2}{2(1+2 |g|_H^2)}}}{\frac{1}{\sqrt{2\pi}}e^{-\frac{x^2}{2}}}\Big\vert_{x=\delta(\tilde{g})}. \end{eqnarray*} Hence, the density $\xi$ can be written more explicitly as \begin{eqnarray*} \sum_{k\geq 0}\frac{\delta^{2k}(g^{\otimes 2k})}{k!}=\frac{\frac{1}{\sqrt{2\pi(1+2 |g|_H^2)}}e^{-\frac{x^2}{2(1+2 |g|_H^2)}}}{\frac{1}{\sqrt{2\pi}}e^{-\frac{x^2}{2}}}\Big\vert_{x=\delta(\tilde{\tilde{g}})}. \end{eqnarray*} \subsection{Convolution measures} Let $(W,H,\mu)$ be an abstract Wiener space and let $X$ be a random variable taking values on $W$. Assume that $X=Z+Y$ where $Z$ and $Y$ are independent, the law of $Z$ is $\mu$ and the law of $Y$, say $\nu$, is supported on $H$. Then, the law of $X$ is given by $\mu\star\nu$ where $\star$ denotes the convolution of probability measures. This class of probability measures has an important role in the applications being a Gaussian (white noise) perturbation of a probability measure on the Hilbert space $H$. Poincar\'e-type inequalities with respect to this class of measures have been investigated in \cite{L 2016} and \cite{DLS 2016}.\\ Observe that the measure $\mu\star\nu$ is absolutely continuous with respect to $\mu$ with a density given by \begin{eqnarray}\label{density convolution} \frac{d (\mu\star\nu)}{d \mu}=\int_H\mathcal{E}(h)d\nu(h) \end{eqnarray} (here $\mathcal{E}(h)$ denotes the stochastic exponential: see (\ref{def stochastic exponential}) in the Appendix below). We now want to check that the measure $\mu\star\nu$ fulfills the assumptions of Theorem \ref{main theorem} . First of all, we need to verify the membership of (\ref{density convolution}) to $\mathcal{L}^2(W,\mu)$. According to the Minkowsky integral inequality we can write \begin{eqnarray*} \Big\Vert\int_H\mathcal{E}(h)d\nu(h)\Big\Vert_2&\leq&\int_H\Vert\mathcal{E}(h)\Vert_2d\nu(h)\\ &=&\int_H\exp\Big\{\frac{|h|_H^2}{2}\Big\}d\nu(h). \end{eqnarray*} Therefore, the membership of (\ref{density convolution}) to $\mathcal{L}^2(W,\mu)$ is guaranteed if $\nu$ satisfies the following exponential integrability condition \begin{eqnarray}\label{exponential integrability} \int_H\exp\Big\{\frac{|h|_H^2}{2}\Big\}d\nu(h)<+\infty. \end{eqnarray} We now compute the variance of $\langle X,h\rangle$ for $h\in H$. We have: \begin{eqnarray*} Var(\langle X,h\rangle)&=&Var(\langle Z+Y,h\rangle)\\ &=&Var(\langle Z,h\rangle+\langle Y,h\rangle)\\ &=&Var(\langle Z,h\rangle)+Var(\langle Y,h\rangle)\\ &=&|h|_H^2+Var(\langle Y,h\rangle). \end{eqnarray*} This yields immediately (\ref{positive}). In addition, \begin{eqnarray*} \sum_{j\geq 1}[Var(\langle X,e_j\rangle)-1]=\sum_{j\geq 1}Var(\langle Y,e_j\rangle). \end{eqnarray*} Since the measure $\nu$, the law of $Y$, is supported on the Hilbert space $H$ and satisfies the condition (\ref{exponential integrability}), it follows from the previous equality that also (\ref{trace class}) from Assumption \ref{positive} is satisfied. We now verify the last assumption; by construction (the independence of $Z$ and $Y$) we can write \begin{eqnarray*} &&\sum_{i,j\geq 1}[Cov(\langle X,e_i\rangle,\langle X,e_j\rangle)-\delta_{ij}]^2\\ &=&\sum_{i,j\geq 1}[Cov(\langle Z,e_i\rangle+\langle Y,e_i\rangle,\langle Z,e_j\rangle+\langle Y,e_j\rangle)-\delta_{ij}]^2\\ &=&\sum_{i,j\geq 1}[Cov(\langle Z,e_i\rangle,\langle Z,e_j\rangle)+Cov(\langle Y,e_i\rangle,\langle Y,e_j\rangle)-\delta_{ij}]^2\\ &=&\sum_{i,j\geq 1}Cov(\langle Y,e_i\rangle,\langle Y,e_j\rangle)^2\\ &\leq&\Big(\sum_{i\geq 1}Var(\langle Y,e_i\rangle)\Big)^2. \end{eqnarray*} Hence, Assumption \ref{integrable quadratic exponential} is fulfilled if, for instance, \begin{eqnarray}\label{last convolution measure} \sum_{i\geq 1}Var(\langle Y,e_i\rangle)<1. \end{eqnarray} \subsection{Weak solutions of a class of stochastic differential equations} Consider the system of stochastic differential equations \begin{eqnarray} \left\{ \begin{array}{ll}\label{SDE} dX_t=b_1(Y_t)dt+dB_t^1, & X_0=x \\ dY_t=b_2(X_t)dt+dB_t^2, & Y_0=y \end{array}\right. \end{eqnarray} where $\{(B_t^1,B_t^2)\}_{t\in [0,1]}$ is a two-dimensional standard Brownian motion defined on the probability space $(\Omega,\mathcal{F},\mathcal{P})$, $x,y\in\mathbb{R}$ and $b_1,b_2:\mathbb{R}\to\mathbb{R}$ are measurable functions. If for $i\in\{1,2\}$ the Novikov condition \begin{eqnarray}\label{Novikov} E\Big[\exp\Big\{\frac{1}{2}\int_0^1|b_i(B_t^i)|^2dt\Big\}\Big]<+\infty \end{eqnarray} is satisfied, then by means of the Girsanov theorem one can assert that the process $\{\tilde{B}_t\}_{t\in [0,1]}$, where $\tilde{B}_t=(\tilde{B}_t^1,\tilde{B}_t^2)$ and for $i\neq j\in\{1,2\}$ \begin{eqnarray*} \tilde{B}_t^i=B_t^i-\int_0^tb_i(B_s^j)ds \end{eqnarray*} is a standard two-dimensional Brownian motion under the probability measure $d\mathcal{Q}:=\mathcal{E}d\mathcal{P}$ with \begin{eqnarray*} \mathcal{E}:=\exp\Big\{\int_0^1b_1(B^2_t)dB_t^1+\int_0^1b_2(B^1_t)dB_t^2-\frac{1}{2}\int_0^1|b_1(B^2_t)|^2+|b_2(B^1_t)|^2 dt\Big\}. \end{eqnarray*} As a consequence, the process $\{B_t\}_{t\in [0,1]}$ becomes a weak solution of the system (\ref{SDE}) with respect to the probability space $(\Omega,\mathcal{F},\mathcal{Q})$ and the noise $\{\tilde{B}_t\}_{t\in [0,1]}$. \\ Suppose that we want to investigate the law of the translated Brownian motion $\{\tilde{B}_t\}_{t\in [0,1]}$ under the original measure $\mathcal{P}$. We consider its first component, i.e. \begin{eqnarray}\label{def of X} \tilde{B}_t^1=B_t^1-\int_0^tb_1(B_s^2)ds,\quad t\in [0,1], \end{eqnarray} and we observe that the independence of $B^1$ and $B^2$ implies that the law of the process $\{\tilde{B}^1_t\}_{t\in [0,1]}$ under the measure $\mathcal{P}$ is obtained by convolving the Wiener measure (which is the law of the process $\{B_t^1\}_{t\in [0,1]}$) with a probability measure supported on the Cameron-Martin space (which is the law of the process $\{-\int_0^tb_1(B_s^2)ds\}_{t\in [0,1]}$). This means that we can proceed the investigation via the general framework of convolution measures described in the previous subsection. \\ Let $W$ be the classical Wiener space $C_0([0,1];\mathbb{R})$ of continuous functions $w$ defined on the interval $[0,1]$ with values on $\mathbb{R}$ and such that $w(0)=0$; $H$ the Cameron-Martin space $H_0^1([0,1];\mathbb{R})$ of absolutely continuous functions with square integrable derivative; $\mu$ the classical Wiener measure on $W$.\\ Denote by $X$ the process $t\mapsto\tilde{B}_t^1$ and by $Z$ and $Y$ the processes $t\mapsto B_t^1$ and $t\mapsto -\int_0^tb_1(B_s^2)ds$, respectively. According to (\ref{def of X}), we have $X=Z+Y$ with $Z$ independent of $Y$; the law of $Z$ is $\mu$ while the law of $Y$, say $\nu$, is supported on $H$. We can write explicitly the density of the law of $X$ with respect to $\mu$ via formula (\ref{density convolution}): \begin{eqnarray}\label{5} \int_H\mathcal{E}(h)d\nu(h)&=&\int_H\exp\Big\{\int_0^1\dot{h}_tdw_t-\frac{1}{2}\int_0^1\dot{h}_t^2dt\Big\}d\nu(h)\nonumber\\ &=&\int_W\exp\Big\{-\int_0^1b_1(\tilde{w}_t)dw_t-\frac{1}{2}\int_0^1b_1(\tilde{w}_t)^2dt\Big\}d\mu(\tilde{w}). \end{eqnarray} In the second equality we utilized the fact that the measure $\nu$ is the image of (an independent copy of) $\mu$ through the map $t\mapsto -\int_0^tb_1(\tilde{w}_s)ds$. According to the previous subsection, the density in (\ref{5}) belongs to $\mathcal{L}^2(W,\mu)$ if condition (\ref{exponential integrability}) is fulfilled; in the present framework this is equivalent to \begin{eqnarray*} \int_H\exp\Big\{\frac{1}{2}\int_0^1b_1(\tilde{w}_t)^2dt\Big\}d\mu(\tilde{w})<+\infty, \end{eqnarray*} which is exactly the Novikov condition (\ref{Novikov}). Therefore, Assumption \ref{square integrable density} is equivalent to the Novikov condition (\ref{Novikov}). Furthermore, Assumption \ref{positive trace class} is satisfied according to the discussion of the previous subsection. Let us now focus on Assumption \ref{integrable quadratic exponential}. We know that inequality (\ref{last convolution measure}) is sufficient for that assumption to be true. Let $\{e_i\}_{i\geq 1}$ be an orthonormal bases of $H$; then \begin{eqnarray*} \sum_{i\geq 1}Var(\langle Y,e_i\rangle)&=&\sum_{i\geq 1}E[\langle Y,e_i\rangle^2]-(E[\langle Y,e_i\rangle])^2\\ &\leq&\sum_{i\geq 1}E[\langle Y,e_i\rangle^2]\\ &=&\sum_{i\geq 1}\int_W\Big(\int_0^1b_1(w_t)\dot{e}_i(t)dt\Big)^2d\mu(w)\\ &=&\int_W\sum_{i\geq 1}\Big(\int_0^1b_1(w_t)\dot{e}_i(t)dt\Big)^2d\mu(w)\\ &=&\int_W\int_0^1b^2_1(w_t)dtd\mu(w). \end{eqnarray*} To conclude, if \begin{eqnarray*} \int_0^1E[|b_1(B_t^2)|^2]dt<1 \end{eqnarray*} then Assumption \ref{integrable quadratic exponential} is satisfied. \section{Preliminary results} We are now going to collect several important results of independent interest that will play a crucial role in the proof of Theorem \ref{main theorem}. The next proposition establishes the existence in the Cameron-Martin space $H$ of the mean $E[X]$ of a random element $X$ on $W$. \begin{proposition}\label{existence of mean} Let $X$ be a random variable taking values on $W$. Assume that the law of $X$ is absolutely continuous with respect to $\mu$ with a density $f$ belonging to $\mathcal{L}^2(W,\mu)$. Then, the expectation $E[X]$ of $X$ belongs to $H$ and coincides with the first kernel in the chaos decomposition of $f$. Moreover, the density of $X-E[X]$ is given by $f\diamond\mathcal{E}(-E[X])$. \end{proposition} \begin{proof} The expectation of $X$ is defined to be the unique element $E[X]\in W$ such that \begin{eqnarray}\label{def expectation} E[\langle X,\varphi\rangle]=\langle E[X],\varphi\rangle,\quad\mbox{ for all }\quad\varphi\in W^*. \end{eqnarray} Its existence is guaranteed by the finiteness all the moments of $\langle X,\varphi\rangle$, as explained in the first section. Moreover, \begin{eqnarray*} E[\langle X,\varphi\rangle]&=&\int_W\langle w,\varphi\rangle f(w)d\mu(w)\\ &=&\langle f_1,\varphi\rangle_H \end{eqnarray*} where we utilized equation (\ref{orthogonality chaos expansion}) below and the identity $\langle w,\varphi\rangle=\delta^1(\varphi)$. Comparing with equation (\ref{def expectation}), this shows that $E[X]=f_1\in H$. In addition, using identity (\ref{S-transform}) we can write \begin{eqnarray*} \int_We^{i\langle\cdot,\varphi\rangle}(f\diamond\mathcal{E}(-E[X]))d\mu&=&e^{-\frac{|\varphi|_H^2}{2}} \int_W\mathcal{E}(i\varphi)(f\diamond\mathcal{E}(-E[X]))d\mu\\ &=&e^{-\frac{|\varphi|_H^2}{2}}\int_W\mathcal{E}(i\varphi)fd\mu \int_W\mathcal{E}(i\varphi)\mathcal{E}(-E[X])d\mu\\ &=&e^{-\frac{|\varphi|_H^2}{2}-i\langle E[X],\varphi\rangle_H}\int_W\mathcal{E}(i\varphi)fd\mu\\ &=&e^{-i\langle E[X],\varphi\rangle_H}\int_We^{i\langle\cdot,\varphi\rangle}fd\mu\\ &=&e^{-i\langle E[X],\varphi\rangle_H}E\Big[e^{i\langle X,\varphi\rangle}\Big]\\ &=&E\Big[e^{i\langle X-E[X],\varphi\rangle}\Big]. \end{eqnarray*} Here we employed the fact that $\langle E[X],\varphi\rangle_H=\langle E[X],\varphi\rangle$ since $E[X]\in H$ and $\varphi\in W^*$. The proof is complete. \end{proof} \noindent Since we are dealing with random elements taking values on possibly infinite dimensional abstract Wiener spaces, it is not clear whether we can find an operator acting on densities which replicates the role of the classic convolution product on Euclidean spaces. The next theorem tells that the Wick product fulfills precisely this requirement. Similar results for the Poisson and chi-square distributions can be found in \cite{LS 2013} and \cite{LSportelli 2012}. In the sequel $\Gamma(\lambda)$ for $\lambda\in [0,1]$ denotes the operator defined in (\ref{def Gamma}); it corresponds to the Ornstein-Uhlenbeck semigroup $T_t$ via the relation $\Gamma(e^{-t})=T_t$ for any $t\geq 0$. \begin{theorem}\label{wick product convolution} Let $X_1,...,X_n$ be independent random variables taking values on $W$ and denote by $\mu_{X_1},...,\mu_{X_n}$ the corresponding laws on $W$, respectively. Assume that the measures $\mu_{X_1},...,\mu_{X_n}$ are absolutely continuous with respect to $\mu$. Then, for any $\alpha_1,...\alpha_n\in [0,1]$ such that $\alpha_1 + \cdot\cdot\cdot+\alpha_n =1,$ we have \begin{eqnarray}\label{gaussian wick} \Gamma(\sqrt{\alpha_1})\frac{d\mu_{X_1}}{d\mu}\diamond\cdot\cdot\cdot\diamond\Gamma(\sqrt{\alpha_n})\frac{d\mu_{X_n}}{d\mu} & = & \frac{d\mu_{\sqrt{\alpha_1} X_1+\cdot\cdot\cdot+\sqrt{\alpha_n} X_n}}{d\mu}, \end{eqnarray} where $\frac{dQ}{d\mu}$ denotes the Radon-Nikodym derivative of the measure $Q$ with respect to the reference measure $\mu$ and $\mu_{\sqrt{\alpha_1} X_1+\cdot\cdot\cdot+\sqrt{\alpha_n} X_n}$ denotes the law of the random variable $\sqrt{\alpha_1} X_1+\cdot\cdot\cdot+\sqrt{\alpha_n} X_n$. \end{theorem} \begin{proof} See Proposition 3.1 in \cite{LS 2016}. \end{proof} \noindent According to the previous result, the Wick product can be considered to be a Gaussian analogue of the classic convolution product. From this point of view, the following theorem corresponds to a sharp Young-type inequality in the Gaussian framework. \begin{theorem}\label{Young inequality} Let $\alpha_1,...,\alpha_n\in [0,1]$ be such that $\alpha_1+\cdot\cdot\cdot+\alpha_n = 1$ and let $p_1,...,p_n,r \in [1$, $+\infty]$ satisfy the following condition \begin{eqnarray*} \frac{\alpha_1}{p_1-1}+\cdot\cdot\cdot+\frac{\alpha_n}{p_n-1}=\frac{1}{r-1}. \end{eqnarray*} If $f_i\in\mathcal{L}^{p_i}(W,\mu)$ for each $i=1,...,n$, then $\Gamma(\sqrt{\alpha_1})f_1 \diamond \cdot\cdot\cdot\diamond\Gamma(\sqrt{\alpha_n})f_n\in\mathcal{L}^r(W,\mu)$. More precisely, \begin{eqnarray}\label{gaussian young} \parallel\Gamma(\sqrt{\alpha_1})f\diamond\cdot\cdot\cdot\diamond\Gamma(\sqrt{\alpha_n})f_n\parallel_r & \leq & \parallel f_1\parallel_{p_1}\cdot\cdot\cdot\parallel f_n\parallel_{p_n}. \end{eqnarray} \end{theorem} \begin{proof} See Theorem 4.7 in \cite{DLS 2011}. \end{proof} \subsection{Proof of Proposition \ref{properties quadratic exp}} We first prove that $\xi\in\mathcal{L}^2(W,\mu)$: \begin{eqnarray*} \Vert\xi\Vert^2_2&=&\Big\Vert\sum_{k\geq 0}\frac{\delta^{2k}(g_2^{\otimes k})}{k!}\Big\Vert_2^2\\ &=&\sum_{k\geq 0}\frac{(2k)!}{(k!)^2}|g_2^{\otimes k}|^2_{H^{\otimes 2k}}\\ &=&\sum_{k\geq 0}{2k \choose k}|g_2|^{2k}_{H^{\otimes 2}}\\ &=&\frac{1}{\sqrt{1-4|g_2|^{2}_{H^{\otimes 2}}}} \end{eqnarray*} provided that \begin{eqnarray*} |g_2|^{2}_{H^{\otimes 2}}<\frac{1}{4}. \end{eqnarray*} Here we used the generating function of the central binomial coefficients, i.e \begin{eqnarray*} \sum_{k\geq 0}{2k \choose k}x^k=\frac{1}{\sqrt{1-4x}},\quad\mbox{ for }x<\frac{1}{4}. \end{eqnarray*} On the other hand, as explained in the introduction, Assumption \ref{integrable quadratic exponential} serves precisely to guarantee that $|g_2|^{2}_{H^{\otimes 2}}<\frac{1}{4}$, entailing the square integrability of $\xi$.\\ We now prove that $\xi$ is non negative by means of a characterization of positivity proposed by Nualart and Zakai in \cite{Nualart Zakai}. According to that paper, we need to prove that the function \begin{eqnarray}\label{def F} h\in H\mapsto F(h)=\int_We^{i\langle w,h\rangle}\xi(w)d\mu(w) \end{eqnarray} is positive definite, i.e. for all $z_1,...,z_n\in\mathbb{C}$ and $h_1,...,h_n\in H$ one has \begin{eqnarray*} \sum_{j,l=1}^nz_jF(h_i-h_l)\bar{z}_l\geq 0. \end{eqnarray*} It is not difficult to see that \begin{eqnarray*} F(h)&=&\sum_{k\geq 0}\frac{(-1)^{k}}{k!}\langle g_2^{\otimes k},h^{\otimes 2k}\rangle_{H^{\otimes 2k}}\cdot e^{-\frac{|h|_H^2}{2}}\\ &=&\sum_{k\geq 0}\frac{(-1)^{k}}{k!}\langle g_2,h^{\otimes 2}\rangle_{H^{\otimes 2}}^k\cdot e^{-\frac{|h|_H^2}{2}}\\ &=&\exp\Big\{-\langle g_2,h^{\otimes 2}\rangle_{H^{\otimes 2}}-\frac{|h|_H^2}{2}\Big\}. \end{eqnarray*} Since by Assumption \ref{positive trace class} the kernel $g_2=f_2-\frac{f_1^{\otimes 2}}{2}$ is positive, symmetric and of trace class, the function $\exp\{-\langle g_2,h^{\otimes 2}\rangle_{H^{\otimes 2}}\}$ is the characteristic functional of a Gaussian measure on $H$ with covariance operator equal to $2g_2$ and hence is positive definite. Moreover, $\exp\Big\{-\frac{|h|_H^2}{2}\Big\}$ is the characteristic functional of the Wiener measure $\mu$, entailing its positive definiteness. This show that the function in (\ref{def F}) is positive definite being the product of two functions of this type. This in turn also proves that $\xi d\mu$ is a Gaussian measure on $W$ obtained convolving the Wiener measure with a Gaussian measure on $H$ (whose covariance operator is equal to $2g_2$).\\ We are now left with the proof of (\ref{WID}). Note that according to Theorem \ref{wick product convolution}, since the law of the $\mathcal{X}_i$'s has density $\xi$ with respect to $\mu$, then the law of $\frac{\mathcal{X}_1+\cdot\cdot\cdot+\mathcal{X}_n}{\sqrt{n}}$ has density $\Gamma(1/\sqrt{n})\xi\diamond\cdot\cdot\cdot\diamond\Gamma(1/\sqrt{n})\xi$ ($n$-times). Therefore, (\ref{WID}) is equivalent to \begin{eqnarray*} \underbrace{\Gamma(1/\sqrt{n})\xi\diamond\cdot\cdot\cdot\diamond\Gamma(1/\sqrt{n})\xi}_{\mbox{$n$-times}}=\xi. \end{eqnarray*} By definition of Wick product (see (\ref{def wick product}) below), \begin{eqnarray*} \Gamma(1/\sqrt{n})\xi\diamond\cdot\cdot\cdot\diamond\Gamma(1/\sqrt{n})\xi &=&\sum_{k\geq 0}\sum_{i_1+\cdot\cdot\cdot+i_n=k}\frac{n^{-i_1}\cdot\cdot\cdot n^{-i_n}}{i_1!\cdot\cdot\cdot i_n!}\delta^{2k}(g_2^{\otimes i_1}\hat{\otimes}\cdot\cdot\cdot\hat{\otimes} g_2^{\otimes i_n})\\ &=&\sum_{k\geq 0}\frac{n^{-k}}{k!}\delta^{2k}\Big(\sum_{i_1+\cdot\cdot\cdot+i_n=k}\frac{k!}{i_1!\cdot\cdot\cdot i_n!}g_2^{\otimes i_1}\hat{\otimes}\cdot\cdot\cdot\hat{\otimes} g_2^{\otimes i_n}\Big)\\ &=&\sum_{k\geq 0}\frac{n^{-k}}{k!}\delta^{2k}((\underbrace{g_2+\cdot\cdot\cdot+g_2}_{\mbox{$n$-times}})^{\otimes k})\\ &=&\sum_{k\geq 0}\frac{\delta^{2k}(g_2^{\otimes k})}{k!}\\ &=&\xi. \end{eqnarray*} The proof is complete \section{Proof of Theorem \ref{main theorem}} We are now ready to prove our local limit theorem. The proof will be essentially based on the algebraic and analytical properties of the Wick product and Ornstein-Uhlenbeck semigroup, through the results described in the previous section.\\ Let $f\in\mathcal{L}^2(W,\mu)$ denote the common density of the $X_n$'s with respect to the measure $\mu$. According to Proposition \ref{existence of mean}, the density of $X_n-E[X_n]$ is given by $f\diamond\mathcal{E}(-f_1)$, where $f_1\in H$ denotes the first kernel in the Wiener-It\^o chaos decomposition of $f$. To ease the notation, we set \begin{eqnarray*}\label{def f tilde} \tilde{f}:=f\diamond\mathcal{E}(-f_1). \end{eqnarray*} We remark that if $f=1+\delta^1(f_1)+\delta^2(f_2)+\cdot\cdot\cdot$, then by definition of Wick product (\ref{def wick product}) and stochastic exponential (\ref{stoch exp. decomposition}) we get \begin{eqnarray}\label{X-E[X]} f\diamond\mathcal{E}(-f_1)&=&1+\delta^1(f_1-f_1)+\delta^2\Big(f_2+f_1\otimes (-f_1)+\frac{(-f_1)^{\otimes 2}}{2}\Big)+\cdot\cdot\cdot\nonumber\\ &=&1+\delta^2\Big(f_2-\frac{f_1^{\otimes 2}}{2}\Big)+\cdot\cdot\cdot. \end{eqnarray} From Theorem \ref{wick product convolution} we know that the density of $\frac{X_1+\cdot\cdot\cdot+X_n-nE[X_1]}{\sqrt{n}}$ is given by \begin{eqnarray*} \Gamma(1/\sqrt{n})\tilde{f}\diamond\cdot\cdot\cdot\diamond\Gamma(1/\sqrt{n})\tilde{f}=(\Gamma(1/\sqrt{n})\tilde{f})^{\diamond n} \end{eqnarray*} where $g^{\diamond n}$ stands for $g\diamond\cdot\cdot\cdot\diamond g$ ($n$-times). Moreover, the density of the random variable in (\ref{limit}) can be written as \begin{eqnarray*} \Gamma(\sqrt{\alpha})[(\Gamma(1/\sqrt{n})\tilde{f})^{\diamond n}]\diamond\Gamma(\sqrt{1-\alpha})1=[\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}]^{\diamond n}. \end{eqnarray*} Here we utilized the functorial property (\ref{functor}) and the identity $\Gamma(\lambda)1=1$ (note that the density of $Z$ with respect to $\mu$ is one). Observe in addition that we can write without ambiguity the right hand side of the previous equation as $\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}$ (again as a consequence of the interplay between Ornstein-Uhlenbeck semigroup and Wick product). Analogously, since the density of $\mathcal{X}$ is $\xi$, we get that the density of $\sqrt{\alpha}\mathcal{X}+\sqrt{1-\alpha}Z$ is $\Gamma(\sqrt{\alpha})\xi$.\\ Our aim is to prove that \begin{eqnarray*} \lim_{n\to +\infty}\Vert \Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}-\Gamma(\sqrt{\alpha})\xi\Vert_1=0. \end{eqnarray*} First of all, exploiting the associativity and distributivity of the Wick product, together with Proposition \ref{properties quadratic exp}, we can write \begin{eqnarray*} \Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}-\Gamma(\sqrt{\alpha})\xi&=&\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}-\Gamma(\sqrt{\alpha}/\sqrt{n})\xi^{\diamond n}\\ &=&(\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}-\Gamma(\sqrt{\alpha}/\sqrt{n})\xi)\diamond\\ &&\diamond\sum_{j=0}^{n-1} \Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond j}\diamond\Gamma(\sqrt{\alpha}/\sqrt{n})\xi^{\diamond n-1-j}. \end{eqnarray*} Now, we take the $\mathcal{L}^1(W,\mu)$-norm and we apply Theorem \ref{Young inequality} (actually we need only the $\mathcal{L}^1$-form of the inequality which was proven before in the paper \cite{LS 2010}) and the triangle inequality to obtain \begin{eqnarray}\label{1} \Vert\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}-\Gamma(\sqrt{\alpha})\xi\Vert_1&\leq&\Vert\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha)n})\tilde{f}-\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha) n})\xi\Vert_1\nonumber\\ &&\cdot\Big\Vert\sum_{j=0}^{n-1} \Gamma(1/\sqrt{n})\tilde{f}^{\diamond j}\diamond\Gamma(1/\sqrt{n})\xi^{\diamond n-1-j}\Big\Vert_1\nonumber\\ &\leq&\Vert\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha)n})\tilde{f}-\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha) n})\xi\Vert_1\nonumber\\ &&\cdot\sum_{j=0}^{n-1} \Vert\Gamma(1/\sqrt{n})\tilde{f}^{\diamond j}\diamond\Gamma(1/\sqrt{n})\xi^{\diamond n-1-j}\Vert_1. \end{eqnarray} Let us now focus the attention on the last sum. Invoking once again Theorem \ref{Young inequality} and exploiting the fact that $\tilde{f}$ and $\xi$ are density functions (their $\mathcal{L}^1(W,\mu)$-norms are equal to one) we get \begin{eqnarray*} &&\sum_{j=0}^{n-1}\Vert\Gamma(1/\sqrt{n})\tilde{f}^{\diamond j}\diamond\Gamma(1/\sqrt{n})\xi^{\diamond n-1-j}\Vert_1\\ &\leq&\sum_{j=0}^{n-1}\Vert\Gamma(\sqrt{n-1}/\sqrt{n})\tilde{f}\Vert_1^{j}\Vert\Gamma(\sqrt{n-1}/\sqrt{n})\xi\Vert_1^{n-j-1}\\ &\leq&\sum_{j=0}^{n-1}\Vert \tilde{f}\Vert_1^{j}\Vert\xi\Vert_1^{n-j-1}\\ &=&n. \end{eqnarray*} Plugging this last estimate in (\ref{1}) we obtain \begin{eqnarray}\label{2} &&\Vert\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}-\Gamma(\sqrt{\alpha})\xi\Vert_1\nonumber\\ &\leq&\Vert\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha)n})\tilde{f}-\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha) n})\xi\Vert_1\nonumber\\ &&\cdot\sum_{j=0}^{n-1} \Vert\Gamma(1/\sqrt{n})\tilde{f}^{\diamond j}\diamond\Gamma(1/\sqrt{n})\xi^{\diamond n-1-j}\Vert_1\nonumber\\ &\leq&n\Vert\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha)n})\tilde{f}-\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha) n})\xi\Vert_1 \end{eqnarray} To ease the notation we set $\beta=\frac{\alpha}{1-\alpha}$ and we observe that, by the Nelson's hyper-contractive estimate \cite{Nelson} and the assumption $f\in\mathcal{L}^2(W,\mu)$, there exists $n_0\in\mathbb{N}$ big enough such that the function $\Gamma(\sqrt{\beta}/\sqrt{n_0})\tilde{f}$ belongs to $\mathcal{L}^2(W,\mu)$; therefore, we can write for all $n\geq n_0$ that \begin{eqnarray*} \Gamma(\sqrt{\beta}/\sqrt{n})\tilde{f}&=&\Gamma(\sqrt{n_0}/\sqrt{n})\Gamma(\sqrt{\beta}/\sqrt{n_0})\tilde{f}\\ &=&\sum_{k\geq 0}\Big(\frac{n_0}{n}\Big)^{\frac{k}{2}}\delta^k(\hat{f}_k)\\ &=&1+\frac{n_0}{n}\delta^2(\hat{f}_2)+\sum_{k\geq 3}\Big(\frac{n_0}{n}\Big)^{\frac{k}{2}}\delta^k(\hat{f}_k) \end{eqnarray*} where the $\hat{f}_k$'s are the kernels in the Wiener-It\^o chaos decomposition of $\Gamma(\sqrt{\beta}/\sqrt{n_0})\tilde{f}$. The same holds true for $\Gamma(\sqrt{\beta}/\sqrt{n})\xi$, i.e. \begin{eqnarray*} \Gamma(\sqrt{\beta}/\sqrt{n})\xi&=&\Gamma(\sqrt{n_0}/\sqrt{n})\Gamma(\sqrt{\beta}/\sqrt{n_0})\xi\\ &=&\sum_{k\geq 0}\Big(\frac{n_0}{n}\Big)^{\frac{k}{2}}\delta^k(\hat{g}_k)\\ &=&1+\frac{n_0}{n}\delta^2(\hat{g}_2)+\sum_{k\geq 3}\Big(\frac{n_0}{n}\Big)^{\frac{k}{2}}\delta^k(\hat{g}_k). \end{eqnarray*} where the $\hat{g}_k$'s are the kernels of $\Gamma(\sqrt{\beta}/\sqrt{n_0})\xi$. Note that by construction $\hat{f}_2=\hat{g}_2$ (recall (\ref{X-E[X]}) and the definition of $g_2$ in Proposition \ref{properties quadratic exp}) which implies \begin{eqnarray*} \Gamma(\sqrt{\beta}/\sqrt{n})\tilde{f}-\Gamma(\sqrt{\beta}/\sqrt{n})\xi=\sum_{k\geq 3}\Big(\frac{n_0}{n}\Big)^{\frac{k}{2}}\delta^k(\hat{f}_k-\hat{g}_k) \end{eqnarray*} Hence, for all $n\geq n_0$ \begin{eqnarray*} &&\Vert\Gamma(\sqrt{\beta}/\sqrt{n})\tilde{f}-\Gamma(\sqrt{\beta}/\sqrt{n})\xi)\Vert_1\\ &\leq&\Vert\Gamma(\sqrt{\beta}/\sqrt{n})\tilde{f}-\Gamma(\sqrt{\beta}/\sqrt{n})\xi)\Vert_2\\ &=&\Vert\Gamma(\sqrt{n_0}/\sqrt{n})\Gamma(\sqrt{\beta}/\sqrt{n_0})\tilde{f}-\Gamma(\sqrt{n_0}/\sqrt{n})\Gamma(\sqrt{\beta}/\sqrt{n})\xi)\Vert_2\\ &=&\Big(\sum_{k\geq 3}k!\Big(\frac{n_0}{n}\Big)^k|\hat{f}_k-\hat{g}_k|^2\Big)^{\frac{1}{2}}\\ &\leq&\Big(\frac{n_0}{n}\Big)^{\frac{3}{2}}\Big(\sum_{k\geq 3}k!|\hat{f}_k-\hat{g}_k|^2\Big)^{\frac{1}{2}}. \end{eqnarray*} Combining the last estimate with (\ref{2}) we conclude that \begin{eqnarray*} &&\Vert\Gamma(\sqrt{\alpha}/\sqrt{n})\tilde{f}^{\diamond n}-\Gamma(\sqrt{\alpha})\xi\Vert_1\\ &\leq&n\Vert\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha)n})\tilde{f}-\Gamma(\sqrt{\alpha}/\sqrt{(1-\alpha) n})\xi\Vert_1\\ &\leq&n\Big(\frac{n_0}{n}\Big)^{\frac{3}{2}}\Big(\sum_{k\geq 3}k!|\hat{f}_k-\hat{g}_k|^2\Big)^{\frac{1}{2}}\\ &=&\frac{C}{\sqrt{n}}\Big(\sum_{k\geq 3}k!|\hat{f}_k-\hat{g}_k|^2\Big)^{\frac{1}{2}}. \end{eqnarray*} The proof of the convergence is complete. \section{Appendix} In this section we recall for the reader's convenience few definitions and notations and collect some useful formulas that we utilized throughout the paper. For more details on the subject we refer the interested reader to one of the books \cite{Bogachev}, \cite{Janson}, \cite{Kuo Banach} and \cite{Nualart}.\\ For $f,g\in\mathcal{L}^2(W,\mu)$ with $f=\sum_{k\geq 0}\delta^k(f_k)$ and $g=\sum_{k\geq 0}\delta^k(g_k)$ one has the identity \begin{eqnarray}\label{orthogonality chaos expansion} \int_Wf(w)g(w)d\mu(w)=\sum_{k\geq 0}k!\langle f_k,g_k\rangle_{H^{\otimes k}}. \end{eqnarray} The \emph{stochastic exponential} is defined as \begin{eqnarray}\label{def stochastic exponential} w\in W\mapsto\mathcal{E}(h)(w):=\exp\Big\{\langle w,h\rangle-\frac{|h|_H^2}{2}\Big\},\quad h\in H. \end{eqnarray} Its chaos decomposition is given by \begin{eqnarray}\label{stoch exp. decomposition} \mathcal{E}(h)=\sum_{k\geq 0}\delta^k\Big(\frac{h^{\otimes k}}{k!}\Big). \end{eqnarray} For any $\lambda\in [0,1[$, we define the operator $\Gamma(\lambda)$ acting on $\mathcal{L}^2(W,\mu)$ as \begin{eqnarray}\label{def Gamma} \Gamma(\lambda)\Big(\sum_{k\geq 0}\delta^k(f_k)\Big):=\sum_{k\geq 0} \lambda^k\delta^k(f_k). \end{eqnarray} We observe that with $\lambda=e^{-\tau}$, $\tau\geq 0$, the operator $\Gamma(\lambda)$ coincides with the Ornstein-Uhlenbeck semigroup \begin{eqnarray*} (P_{\tau}f)(w):=\int_Wf\big(e^{-\tau}w+\sqrt{1-e^{-2\tau}}\tilde{w}\big)d\mu(\tilde{w}),\quad w\in W, \tau\geq 0, \end{eqnarray*} which is a contraction on $\mathcal{L}^p(W,\mu)$ for any $p\geq 1$. On the space $\mathcal{L}^2(W,\mu)$ one can define an unbounded multiplication between functions through the prescription \begin{eqnarray}\label{def wick product} \delta^k(f_k)\diamond\delta^j(f_j):=\delta^{k+j}(f_k\hat{\otimes} f_j) \end{eqnarray} where $\hat{\otimes}$ denotes the symmetric tensor product. This is named \emph{Wick product} of $\delta^k(f_k)$ and $\delta^j(f_j)$ and it is extended by linearity. It is easy to check that for $\lambda\in [0,1]$ and $f,g\in\mathcal{L}^2(W,\mu)$, \begin{eqnarray}\label{functor} \Gamma(\lambda)(f\diamond g)=\Gamma(\lambda)f\diamond\Gamma(\lambda)g \end{eqnarray} and for $h,l\in H$, \begin{eqnarray*} \mathcal{E}(h)\diamond\mathcal{E}(l)=\mathcal{E}(h+l) \end{eqnarray*} and \begin{eqnarray}\label{S-transform} \int_W (f\diamond g)\mathcal{E}(h)d\mu=\int_W f\mathcal{E}(h)d\mu\cdot\int_W g\mathcal{E}(h)d\mu. \end{eqnarray} For additional information on the Wick product (and its role in the theory of stochastic differential equations) we refer to the book by Holden et al. \cite{HOUZ}, the paper \cite{DLS 2013} and the references quoted there.
8e637972a60100a2422d1d94991177321bd0ac62
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} With the discovery of the Higgs Boson in 2012 by both the ATLAS~\cite{Higgs_atlas} and the CMS~\cite{Higgs_cms} collaborations, our understanding of the physics of the fundamental particles up to the $1$ TeV scale, otherwise known as the Standard Model (SM), has been experimentally confirmed. However, though we know with certainty that SM is the correct picture, we also know with an equal degree of certainty that it is not the complete one. The mathematical framework of the SM is inconsistent with that of general relativity, but the later has been experimentally verified. Other experimental discoveries such as the neutrino oscillations, along with theoretical questions such as the origin of mass, the matter antimatter asymmetry, the dark matter and dark energy, the strong CP problem etc provide ample evidence for the existence of new physics (NP) beyond the SM. Over the past three decades, flavour physics has emerged as an important testing ground for the existence of NP. For example, tensions between SM expectations and experimental results have been found in B physics for observables such as the isospin asymmetry $A_I(B\rightarrow K\mu^+\mu^-)$~\cite{isospin}, the longitudinal polarization fraction in $B_s\rightarrow K^*K^*$~\cite{long_pol_1, long_pol_2}, $R_{D^*}^{\tau,l}=\frac{\text{BR}(B\rightarrow D^*\tau\nu_{\tau})}{\text{BR}(B\rightarrow D^*l\nu_l)}$, $R_{D}^{\tau,l}=\frac{\text{BR}(B\rightarrow D\tau\nu_{\tau})}{\text{BR}(B\rightarrow Dl\nu_l)}$~\cite{rddst_1, rddst_2, rddst_3, rddst_4, rddst_5, rddst_6, rddst_7} and $R_K=\frac{\text{BR}(B\rightarrow K\mu^+\mu^-)}{\text{BR}(B\rightarrow Ke^+e^-)}$~\cite{b_kll_1, b_kll_2}. A number of theoretical works have been and are still being undertaken following these results. Top quark physics seems to be important for the search of NP in the up quark sector. However, no signal for NP has yet been detected in the top quark sector. Searches for NP have also been carried out in charm. A considerable amount of work on the $D\rightarrow PP$ and $D\rightarrow VP$ modes~\cite{pp_pv_1, pp_pv_2, pp_pv_3, pp_pv_4, pp_pv_5, pp_pv_6, pp_pv_7, pp_pv_8, pp_pv_9, pp_pv_10, pp_pv_11, pp_pv_12, pp_pv_13, pp_pv_14, pp_pv_15, pp_pv_16, pp_pv_17, pp_pv_18, pp_pv_19, pp_pv_20, pp_pv_21, pp_pv_22, vv_1} has been undertaken over the last thirty years. In 2012, LHCb~\cite{LHCb_charm} and CDF~\cite{CDF_charm} reported the first observation of a CP asymmetry between the $D^0\rightarrow\pi^+\pi^-$ and the $D^0\rightarrow K^+K^-$ modes. This was followed by a large amount of work~\cite{np_1, np_2, np_3, np_4, np_5, np_6, np_7, np_8, np_9, np_10, np_11, np_12, np_13, np_14, np_15, np_16, np_17}, where the authors mostly used NP models to explain the same. The $3.2\sigma$ hint has since then slowly disappeared. However, the $D\rightarrow VV$ hadronic modes have received lesser attention~\cite{vv_1, vv_2}. It is not an easy task to have a solid theoretical understanding of all the charm hadronic modes. This is due to the mass of the charm quark, which, unlike the bottom quark is not sufficiently heavy for the realization of the infinitely heavy quark limit. Hence, the well known approaches based on QCD that lead to satisfactory predictions for $B$ decays like the heavy quark effective theory~\cite{hqet_1, hqet_2}, the QCD-factorization~\cite{qfac_1, qfac_2}, the perturbative QCD approach~\cite{qper_1, qper_2, qper_3, qper_4} and the soft-collinear effective theory~\cite{scet}, fail to achieve the same for $D$ decays. The charm quark is also not light enough for the application of a chiral perturbation theory. Furthermore, in the case of hadronic vector final states, the calculation of the form factors poses greater difficulty than their $D\rightarrow P$ counterparts. In absence of any reliable and effective theoretical models, it might be a good idea to look for possible NP in $D\rightarrow VV$ decays in a model independent way. This is what we have tried to achieve in this paper. The method is similar to the one used in~\cite{ang_an_B} for $B$ decays. The idea is to define observables which can be experimentally measured, linear combinations of which have a value inconsistent with zero under the presence of new physics. Although extensively used for the $B\rightarrow VV$ decays, such a formalism, to the best of our knowledge, has not yet been pursued for the $D\rightarrow VV$ hadronic decays. Our paper is organised as follows. In section~\ref{ang_an} we describe the formalism for defining such model independent observables in detail. The next section (section~\ref{pol}) is about the connection between the polarization basis that we use for our analysis and the $a$, $b$, $c$ amplitudes used in the most general covariant expression for a heavy pseudoscalar to two-body vector decay amplitude. We then discuss our observables and the effect NP has on them in section~\ref{results}. We also extract numerical values for the amplitudes in the polarization basis for $D^0\rightarrow \overline{K^{*0}}\rho^0$ mode, which is the best measured $D\rightarrow VV$ mode in~\cite{pdg}. Finally, we summarize and conclude in section~\ref{summary}. \section{The time-dependent analysis\label{ang_an}} Consider a $D\rightarrow V_1V_2$ decay. Let the SM contribution to this decay be parametrized by a single decay amplitude along with a corresponding strong phase. The NP contribution will in general be characterised by a different decay amplitude with a different strong phase, along with a NP weak phase. The corresponding CP conjugate decay will then have the same components, with the sign of the NP weak phase reversed. A decay of the type $D\rightarrow\bar{V_1}\bar{V_2}$, where the bar denotes the CP conjugate state, and it's corresponding CP conjugate decay will in general be parametrized by some different SM and NP parameters. Hence, for each of the above cases, the decay amplitude for each of the three possible helicity states may be written as: \begin{eqnarray}\label{amplitude} A_{\lambda}&\equiv& \text{Amp}(D\rightarrow V_1V_2)=a_{\lambda}e^{i\delta_{\lambda}^a}+b_{\lambda}e^{i\phi}e^{i\delta_{\lambda}^b}\nonumber\\ \bar{A}_{\bar\lambda}&\equiv& \text{Amp}(\bar{D}\rightarrow \bar{V_1}\bar{V_2})=a_{\lambda}e^{i\delta_{\lambda}^a}+b_{\lambda}e^{-i\phi}e^{i\delta_{\lambda}^b}\nonumber\\ A_{\bar\lambda}&\equiv& \text{Amp}(D\rightarrow \bar{V_1}\bar{V_2})=c_{\lambda}e^{i\delta_{\lambda}^c}+d_{\lambda}e^{i\gamma}e^{i\delta_{\lambda}^d}\nonumber\\ \bar{A}_{\lambda}&\equiv& \text{Amp}(\bar{D}\rightarrow V_1V_2)=c_{\lambda}e^{i\delta_{\lambda}^c}+d_{\lambda}e^{-i\gamma}e^{i\delta_{\lambda}^d}. \end{eqnarray} In the above, $a_{\lambda}$, $c_{\lambda}$ represent the SM decay amplitudes; $\delta_{\lambda}^a$ and $\delta_{\lambda}^c$ denote the SM strong phases; $b_{\lambda}$, $d_{\lambda}$ are the NP decay amplitudes; $\delta_{\lambda}^b$ and $\delta_{\lambda}^d$ the NP strong phases and $\phi$, $\gamma$ are the NP weak phases. The helicity index $\lambda$ takes the values ${0, \parallel, \perp}$ denoting the longitudinal, transverse parallel and the transverse perpendicular helicities respectively. Using CPT invariance, the full amplitude for each of the above decays can be written as \begin{eqnarray}\label{amplitude} \mathcal{A}_{f}&\equiv&\text{Amp}\left(D(t)\rightarrow V_{1}V_{2}\right)={A_{f}}_{0}g_{0}+{A_{f}}_{\parallel}g_{\parallel}+i{A_{f}}_{\perp}g_{\perp}\nonumber\\ \mathcal{A}_{\bar f}&\equiv&\text{Amp}\left(D(t)\rightarrow\overline{V}_{1}\overline{V}_{2}\right)={A_{\bar f}}_{0}g_{0}+{A_{\bar f}}_{\parallel}g_{\parallel}+i{A_{\bar f}}_{\perp}g_{\perp}\nonumber\\ \bar{\mathcal{A}}_{f}&\equiv&\text{Amp}\left(\overline{D}(t)\rightarrow V_{1}V_{2}\right)={\overline{A}_{f}}_{0}g_{0}+{\overline{A}_{f}}_{\parallel}g_{\parallel}-i{\overline{A}_{f}}_{\perp}g_{\perp}\nonumber\\ \bar{\mathcal{A}}_{\bar f}&\equiv&\text{Amp}\left(\overline{D}(t)\rightarrow\overline{V}_{1}\overline{V}_{2}\right)={\overline{A}_{\bar f}}_{0}g_{0}+ {\overline{A}_{\bar f}}_{\parallel}g_{\parallel}-i{\overline{A}_{\bar f}}_{\perp}g_{\perp}. \end{eqnarray} where $g_{\lambda}$'s are basically functions of the angles describing the kinematics for the the corresponding decay. Based on the parametrization discussed above, one can now look into the time dependent decay rates of a neutral $D$ meson going to two vector final states and perform an angular analysis of all the $D(t)\to V_1V_2$, $D(t)\to\overline{V}_1\overline{V}_2$ modes and their CP conjugate processes. The general expressions for the time-dependent decay rates of a neutral meson $M^0$ going to two vector final states are given by \begin{eqnarray} \Gamma \left(M^0(t)\to V_1V_2\right)=&N_fe^{-\text{$\Gamma $t}}\left(\left|\mathcal{A}_f\right|^2+\left|\frac{q}{p} \bar{\mathcal{A}}_f\right|^2\right)\cosh\left(\text{y$\Gamma $t}\right) +\left(\left|\mathcal{A}_f\right|^2-\left|\frac{q}{p} \bar{\mathcal{A}}_f\right|^2\right)\cos\left(\text{x$\Gamma $t}\right)\nonumber&\\ &+2\text{Re}\left(\left(\frac{q}{p}\right)\mathcal{A}_{f}^{*}\bar{\mathcal{A}}_{f}\right)\sinh(\text{y$\Gamma$t}) -2\text{Im}\left(\left(\frac{q}{p}\right)\mathcal{A}_{f}^{*}\bar{\mathcal{A}}_{f}\right)\text{sin}\left(\text{x$\Gamma$t}\right)\nonumber&\\ \Gamma \left(M^0(t)\to \overline{V}_1 \overline{V}_2 \right)=&N_fe^{-\text{$\Gamma $t}}\left(\left|\mathcal{A}_{\bar f}\right|^2+\left|\frac{q}{p} \bar{\mathcal{A}}_{\bar f}\right|^2\right)\cosh\left(\text{y$\Gamma $t}\right) +\left(\left|\mathcal{A}_{\bar f}\right|^2-\left|\frac{q}{p} \bar{\mathcal{A}}_{\bar f}\right|^2\right)\cos\left(\text{x$\Gamma $t}\right)\nonumber&\\ &+2\text{Re}\left(\left(\frac{q}{p}\right)\mathcal{A}_{\bar f}^{*}\bar{\mathcal{A}}_{\bar f}\right)\sinh(\text{y$\Gamma$t}) -2\text{Im}\left(\left(\frac{q}{p}\right)\mathcal{A}_{\bar f}^{*}\bar{\mathcal{A}}_{\bar f}\right)\text{sin}\left(\text{x$\Gamma$t}\right)\nonumber&\\ \Gamma \left(\overline{M^0}(t)\to V_1V_2\right)=&N_fe^{-\text{$\Gamma $t}}\left(\left|\frac{p}{q} \mathcal{A}_f\right|^2+\left|\bar{\mathcal{A}}_f\right|^2\right)\cosh\left(\text{y$\Gamma $t}\right) -\left(\left|\frac{p}{q} \mathcal{A}_f\right|^2-\left|\bar{\mathcal{A}}_f\right|^2\right)\cos\left(\text{x$\Gamma $t}\right)\nonumber&\\ &+2\text{Re}\left(\left(\frac{p}{q}\right)\mathcal{A}_{f}\bar{\mathcal{A}}_{f}^{*}\right)\sinh(\text{y$\Gamma$t}) -2\text{Im}\left(\left(\frac{p}{q}\right)\mathcal{A}_{f}\bar{\mathcal{A}}_{f}^{*}\right)\text{sin}\left(\text{x$\Gamma$t}\right)\nonumber&\\ \Gamma \left(\overline{M^0}(t)\to \overline{V}_1 \overline{V}_2\right)=&N_fe^{-\text{$\Gamma $t}}\left(\left|\frac{p}{q} \mathcal{A}_{\bar f}\right|^2+\left|\bar{\mathcal{A}}_{\bar f}\right|^2\right)\cosh\left(\text{y$\Gamma $t}\right) -\left(\left|\frac{p}{q} \mathcal{A}_{\bar f}\right|^2-\left|\bar{\mathcal{A}}_{\bar f}\right|^2\right)\cos\left(\text{x$\Gamma $t}\right)\nonumber&\\ &+2\text{Re}\left(\left(\frac{p}{q}\right)\mathcal{A}_{\bar f}\bar{\mathcal{A}}_{\bar f}^{*}\right)\sinh(\text{y$\Gamma$t}) -2\text{Im}\left(\left(\frac{p}{q}\right)\mathcal{A}_{\bar f}\bar{\mathcal{A}}_{\bar f}^{*}\right)\text{sin}\left(\text{x$\Gamma$t}\right).& \end{eqnarray} Here $x=\frac{\Delta m}{\Gamma}$, $y=\frac{\Delta\Gamma}{2\Gamma}$, $\Delta m=m_{H}-m_{L}$, $\Delta\Gamma=\Gamma_{H}-\Gamma_{L}$. The indices H, L stand for the heavy and light mass eigenstates. $N_{f}$ is a time independent normalization factor. Decays occurring without $M^0-\overline{M^0}$ oscillations (pure decays) are given by the terms proportional to $\left|\mathcal{A}_{f}\right|^{2}$ or $\left|\bar{A}_{f}\right|^{2}$. Terms proportional to $\left|\frac{q}{p}\bar{\mathcal{A}}_{f}\right|^{2}$ or $\left|\frac{p}{q}\mathcal{A}_{f}\right|^{2}$ are due to decays following the $M^0-\overline{M^0}$ oscillation (decays after mixing). BABAR, while measuring the effect of CP-violation on the time-dependent decay rates in the $D\rightarrow K\pi$ system, parametrized the decay rates in powers of $\Gamma t$~\cite{BABAR}. They truncate the series at second order in $\Gamma t$. However the sensitivity of the coefficient of the second order term to the decay rate is less than that of the coefficient of the first order term~\cite{Ball}. We follow the same parametrization and keep the terms up to the quadratic order in $\Gamma t$. It is clear that our observables will be linear combinations of the coefficients of the sinh, sin, cosh and cos terms. For the case of neutral charm decays, the corresponding decay rates look like \begin{align}\label{decay rate} \Gamma \left(D^0(t)\to V_{1}V_{2}\right)=&N_fe^{-\text{$\Gamma $t}}\bigg(2\left|\mathcal{A}_{f}\right|^{2}+\left[2yRe\left(\frac{q}{p}\mathcal{A}_{f}^{*}\bar{\mathcal{A}}_{f}\right) -2xIm\left(\frac{q}{p}\mathcal{A}_{f}^{*}\bar{\mathcal{A}}_{f}\right)\right]\Gamma t\nonumber\\ &+\left[\frac{(y^{2}-x^{2})}{2}\left|\mathcal{A}_{f}\right|^{2}+\frac{(x^{2}+y^{2})}{2}\left|\frac{q}{p}\bar{\mathcal{A}}_{f}\right|^{2}\right](\Gamma t)^{2}\bigg)\nonumber\\ \Gamma \left(D^0(t)\to \overline{V}_{1} \overline{V}_{2}\right)=&N_fe^{-\text{$\Gamma $t}}\bigg(2\left|\mathcal{A}_{\bar f}\right|^{2}+\left[2yRe\left(\frac{q}{p}\mathcal{A}_{\bar f}^{*}\bar{\mathcal{A}}_{\bar f}\right) -2xIm\left(\frac{q}{p}\mathcal{A}_{\bar f}^{*}\bar{\mathcal{A}}_{\bar f}\right)\right]\Gamma t\nonumber\\ &+\left[\frac{(y^{2}-x^{2})}{2}\left|\mathcal{A}_{\bar f}\right|^{2}+\frac{(x^{2}+y^{2})}{2}\left|\frac{q}{p}\bar{\mathcal{A}}_{\bar f}\right|^{2}\right](\Gamma t)^{2}\bigg)\nonumber\\ \Gamma \left(\overline{D^0}(t)\to V_{1}V_{2}\right)=&N_fe^{-\text{$\Gamma $t}}\bigg(2\left|\bar{\mathcal{A}}_{f}\right|^{2}+\left[2yRe\left(\frac{p}{q}\bar{\mathcal{A}}_{f}^{*}\mathcal{A}_{f}\right) -2xIm\left(\frac{p}{q}\bar{\mathcal{A}}_{f}^{*}\mathcal{A}_{f}\right)\right]\Gamma t\nonumber\\ &+\left[\frac{(y^{2}-x^{2})}{2}\left|\bar{\mathcal{A}}_{f}\right|^{2}+\frac{(x^{2}+y^{2})}{2}\left|\frac{p}{q}\mathcal{A}_{f}\right|^{2}\right](\Gamma t)^{2}\bigg)\nonumber\\ \Gamma \left(\overline{D^0}(t)\to\overline{V}_{1}\overline{V}_{2}\right)=&N_fe^{-\text{$\Gamma $t}}\bigg(2\left|\bar{\mathcal{A}}_{\bar f}\right|^{2}+\left[2yRe\left(\frac{p}{q}\bar{\mathcal{A}}_{\bar f}^{*}\mathcal{A}_{\bar f}\right) -2xIm\left(\frac{p}{q}\bar{\mathcal{A}}_{\bar f}^{*}\mathcal{A}_{\bar f}\right)\right]\Gamma t\nonumber\\ &+\left[\frac{(y^{2}-x^{2})}{2}\left|\bar{\mathcal{A}}_{\bar f}\right|^{2}+\frac{(x^{2}+y^{2})}{2}\left|\frac{p}{q}\mathcal{A}_{\bar f}\right|^{2}\right](\Gamma t)^{2}\bigg) \end{align} Using eqns~(\ref{amplitude}) and~(\ref{decay rate}) we can write the time dependent decay rates as \begin{eqnarray} \Gamma\left(D^0(t)\to V_{1}V_{2}\right)&=&e^{-\Gamma t}\sum_{\lambda\leq\lambda'}\left(X^{V_{1}V_{2}}_{\lambda\lambda'}+Y^{V_{1}V_{2}}_{\lambda\lambda'}\,\Gamma t+Z^{V_{1}V_{2}}_{\lambda\lambda'}\,(\Gamma t)^2\right)g_{\lambda}g_{\lambda'}\nonumber\\ \Gamma\left(D^0(t)\to\overline{V}_{1}\overline{V}_{2}\right)&=&e^{-\Gamma t}\sum_{\lambda\leq\lambda'}\left(X^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda'}+Y^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda'}\,\Gamma t+Z^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda'}\,(\Gamma t)^2\right)g_{\lambda}g_{\lambda'}\nonumber\\ \Gamma\left(\overline{D^0}(t)\to V_{1}V_{2}\right)&=&e^{-\Gamma t}\sum_{\lambda\leq\lambda'}\left(\bar{X}^{V_{1}V_{2}}_{\lambda\lambda'}+\bar{Y}^{V_{1}V_{2}}_{\lambda\lambda'}\,\Gamma t+\bar{Z}^{V_{1}V_{2}}_{\lambda\lambda'}\,(\Gamma t)^2\right)g_{\lambda}g_{\lambda'}\nonumber\\ \Gamma\left(\overline{D^0}(t)\to\overline{V}_{1}\overline{V}_{2}\right)&=&e^{-\Gamma t}\sum_{\lambda\leq\lambda'}\left(\bar{X}^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda'}+\bar{Y}^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda'}\,\Gamma t+\bar{Z}^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda'}\,(\Gamma t)^2\right)g_{\lambda}g_{\lambda'}. \end{eqnarray} Thus by a time dependent angular analysis of the decay modes $D^0(t)\to V_{1}V_{2}$, $D^0(t)\to\overline{V}_{1}\overline{V}_{2}$, $\overline{D^0}(t)\to V_{1}V_{2}$, $\overline{D^0}(t)\to\overline{V}_{1}\overline{V}_{2}$, one can define 72 observables. However, not all of these observables are independent. For example, it can be easily verified that the $Z$'s (the coefficient of the $(\Gamma t)^2$) terms) are linear combinations of the $X$'s. In fact, observables extracted from the coefficients of the terms that are even powers i $\Gamma t$ can always be written as linear combinations of the $X$'s, while those extracted from the coefficients of the terms that are odd powers in $\Gamma t$ can be written as linear combinations of $Y$'s. Hence, in our analysis, we deal with a total of 48 observables (24 $X$'s and 24 $Y$'s). In terms of the amplitudes, these are given by: \begin{align}\label{pol_obs} X^{V_{1}V_{2}}_{\lambda\lambda}=&2(A_{f}^{*})_{\lambda}(A_{f})_{\lambda}\nonumber\\ X^{V_{1}V_{2}}_{i\perp}=&-4Im\left[(A_{f}^{*})_{i}(A_{f})_{\perp}\right]\nonumber\\ X^{V_{1}V_{2}}_{0\parallel}=&4Re\left[(A_{f})_{0}^{*}(A_{f})_{\parallel}\right]\nonumber\\ X^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda}=&2(A_{\bar f}^{*})_{\lambda}(A_{\bar f})_{\lambda}\nonumber\\ X^{\overline{V}_{1}\overline{V}_{2}}_{i\perp}=&-4Im\left[(A_{\bar f}^{*})_{i}(A_{\bar f})_{\perp}\right]\nonumber\\ X^{\overline{V}_{1}\overline{V}_{2}}_{0\parallel}=&4Re\left[(A_{\bar f})_{0}^{*}(A_{\bar f})_{\parallel}\right]\nonumber\\ \bar{X}^{V_{1}V_{2}}_{\lambda\lambda}=&2(\bar{A}_{f}^{*})_{\lambda}(\bar{A}_{f})_{\lambda}\nonumber\\ \bar{X}^{V_{1}V_{2}}_{i\perp}=&-4Im\left[(\bar{A}_{f}^{*})_{\perp}(\bar{A}_{f})_{i}\right]\nonumber\\ \bar{X}^{V_{1}V_{2}}_{0\parallel}=&4Re\left[(\bar{A}_{f})_{0}^{*}(\bar{A}_{f})_{\parallel}\right]\nonumber\\ \bar{X}^{\overline{V}_{1}\overline{V}_{2}}_{\lambda\lambda}=&2(\bar{A}_{\bar f}^{*})_{\lambda}(\bar{A}_{\bar f})_{\lambda}\nonumber\\ \bar{X}^{\overline{V}_{1}\overline{V}_{2}}_{i\perp}=&-4Im\left[(\bar{A}_{\bar f}^{*})_{\perp}(\bar{A}_{\bar f})_{i}\right]\nonumber\\ \bar{X}^{\overline{V}_{1}\overline{V}_{2}}_{0\parallel}=&4Re\left[(\bar{A}_{\bar f})_{0}^{*}(\bar{A}_{\bar f})_{\parallel}\right]\nonumber\\ Y^{V_{1}V_{2}}_{ii}=&2yRe\left[\frac{q}{p}(A_{f}^{*})_{i}(\bar{A}_{f})_{i}\right] -2xIm\left[\frac{q}{p}(A_{f}^{*})_{i}(\bar{A}_{f})_{i}\right]\nonumber\\ Y^{V_{1}V_{2}}_{\perp\perp}=&-2yRe\left[\frac{q}{p}(A_{f}^{*})_{\perp}(\bar{A}_{f})_{\perp}\right] +2xIm\left[\frac{q}{p}(A_{f}^{*})_{\perp}(\bar{A}_{f})_{\perp}\right]\nonumber\\ Y^{V_{1}V_{2}}_{i\perp}=&-2yRe\left[\frac{q}{p}i\left((A_{f}^{*})_{i}(\bar{A}_{f})_{\perp}+(A_{f}^{*})_{\perp}(\bar{A}_{f})_{i}\right)\right] +2xIm\left[\frac{q}{p}i\left((A_{f}^{*})_{i}(\bar{A}_{f})_{\perp}+(A_{f}^{*})_{\perp}(\bar{A}_{f})_{i}\right)\right]\nonumber\\ Y^{V_{1}V_{2}}_{0\parallel}=&2yRe\left[\frac{q}{p}\left((A_{f})_{0}^{*}(\bar{A}_{f})_{\parallel}+(A_{f})_{\parallel}^{*}(\bar{A}_{f})_{0}\right)\right] -2xIm\left[\frac{q}{p}\left((A_{f})_{0}^{*}(\bar{A}_{f})_{\parallel}+(A_{f})_{\parallel}^{*}(\bar{A}_{f})_{0}\right)\right]\nonumber\\ Y^{\overline{V}_{1}\overline{V}_{2}}_{ii}=&2yRe\left[\frac{q}{p}(A_{\bar f}^{*})_{i}(\bar{A}_{\bar f})_{i}\right] -2xIm\left[\frac{q}{p}(A_{\bar f}^{*})_{i}(\bar{A}_{\bar f})_{i}\right]\nonumber\\ Y^{\overline{V}_{1}\overline{V}_{2}}_{\perp\perp}=&-2yRe\left[\frac{q}{p}(A_{\bar f}^{*})_{\perp}(\bar{A}_{\bar f})_{\perp}\right] +2xIm\left[\frac{q}{p}(A_{\bar f}^{*})_{\perp}(\bar{A}_{\bar f})_{\perp}\right]\nonumber\\ Y^{\overline{V}_{1}\overline{V}_{2}}_{i\perp}=&-2yRe\left[\frac{q}{p}i\left((A_{\bar f}^{*})_{i}(\bar{A}_{\bar f})_{\perp}+(A_{\bar f}^{*})_{\perp}(\bar{A}_{\bar f})_{i}\right)\right] +2xIm\left[\frac{q}{p}i\left((A_{\bar f}^{*})_{i}(\bar{A}_{\bar f})_{\perp}+(A_{\bar f}^{*})_{\perp}(\bar{A}_{\bar f})_{i}\right)\right]\nonumber\\ \end{align} \begin{align} Y^{\overline{V}_{1}\overline{V}_{2}}_{0\parallel}=&2yRe\left[\frac{q}{p}\left((A_{\bar f})_{0}^{*}(\bar{A}_{\bar f})_{\parallel}+(A_{\bar f})_{\parallel}^{*}(\bar{A}_{\bar f})_{0}\right)\right] -2xIm\left[\frac{q}{p}\left((A_{\bar f})_{0}^{*}(\bar{A}_{\bar f})_{\parallel}+(A_{\bar f})_{\parallel}^{*}(\bar{A}_{\bar f})_{0}\right)\right]\nonumber\\ \bar{Y}^{V_{1}V_{2}}_{ii}=&2yRe\left[\frac{p}{q}(\bar{A}_{f}^{*})_{i}(A_{f})_{i}\right] -2xIm\left[\frac{p}{q}(\bar{A}_{f}^{*})_{i}(A_{f})_{i}\right]\nonumber\\ \bar{Y}^{V_{1}V_{2}}_{\perp\perp}=&-2yRe\left[\frac{p}{q}(\bar{A}_{f}^{*})_{i}(A_{f})_{i}\right] +2xIm\left[\frac{p}{q}(\bar{A}_{f}^{*})_{i}(A_{f})_{i}\right]\nonumber\\ \bar{Y}^{V_{1}V_{2}}_{i\perp}=&2yRe\left[\frac{p}{q}i\left((\bar{A}_{f}^{*})_{i}(A_{f})_{\perp}+(\bar{A}_{f}^{*})_{\perp}(A_{f})_{i}\right)\right] -2xIm\left[\frac{p}{q}i\left((\bar{A}_{f}^{*})_{i}(A_{f})_{\perp}+(\bar{A}_{f}^{*})_{\perp}(A_{f})_{i}\right)\right]\nonumber\\ \bar{Y}^{V_{1}V_{2}}_{0\parallel}=&2yRe\left[\frac{p}{q}\left((\bar{A}_{f})_{0}^{*}(A_{f})_{\parallel}+(\bar{A}_{f})_{\parallel}^{*}(A_{f})_{0}\right)\right] -2xIm\left[\frac{p}{q}\left((\bar{A}_{f})_{0}^{*}(A_{f})_{\parallel}+(\bar{A}_{f})_{\parallel}^{*}(A_{f})_{0}\right)\right]\nonumber\\ \bar{Y}^{\overline{V}_{1}\overline{V}_{2}}_{ii}=&2yRe\left[\frac{p}{q}(\bar{A}_{\bar f}^{*})_{i}(A_{\bar f})_{i}\right] -2xIm\left[\frac{p}{q}(\bar{A}_{\bar f}^{*})_{i}(A_{\bar f})_{i}\right]\nonumber\\ \bar{Y}^{\overline{V}_{1}\overline{V}_{2}}_{\perp\perp}=&-2yRe\left[\frac{p}{q}(\bar{A}_{\bar f}^{*})_{\perp}(A_{\bar f})_{\perp}\right] +2xIm\left[\frac{p}{q}(\bar{A}_{\bar f}^{*})_{\perp}(A_{\bar f})_{\perp}\right]\nonumber\\ \bar{Y}^{\overline{V}_{1}\overline{V}_{2}}_{i\perp}=&2yRe\left[\frac{p}{q}i\left((\bar{A}_{\bar f}^{*})_{i}(A_{\bar f})_{\perp}+(\bar{A}_{\bar f}^{*})_{\perp}(A_{\bar f})_{i}\right)\right] -2xIm\left[\frac{p}{q}i\left((\bar{A}_{\bar f}^{*})_{i}(A_{\bar f})_{\perp}+(\bar{A}_{\bar f}^{*})_{\perp}(A_{\bar f})_{i}\right)\right]\nonumber\\ \bar{Y}^{\overline{V}_{1}\overline{V}_{2}}_{0\parallel}=&2yRe\left[\frac{p}{q}\left((\bar{A}_{\bar f})_{0}^{*}(A_{\bar f})_{\parallel}+(\bar{A}_{\bar f})_{\parallel}^{*}(A_{\bar f})_{0}\right)\right] -2xIm\left[\frac{p}{q}\left((\bar{A}_{\bar f})_{0}^{*}(A_{\bar f})_{\parallel}+(\bar{A}_{\bar f})_{\parallel}^{*}(A_{\bar f})_{0}\right)\right] \end{align} where $\frac{q}{p}=re^{i(\alpha_{SM}+\alpha_{NP})} \footnote{Within the SM, CP violation in the charm sector is known to be very small ($\mathcal O(\lambda^5)$ in the Wolfenstein parameterization of the CKM matrix). However, no NP has yet been found in the charm sector, and hence, any NP effect, if present, has to be very tiny too. It is therefore not feasible to neglect the tiny CP violating phase in charm mixing when one tries to probe NP in the charm sector. Hence we parameterize the CP violating mixing phase in charm according to the SM ($\alpha_{SM}$) and the NP ($\alpha_{NP}$) counterparts.}$. These 48 observables can be written in terms of 28 independent parameters: $r$, $\alpha_{SM}, \alpha_{NP}$, $\phi$, $\chi$, three each of $a_{\lambda}$'s, $b_{\lambda}$'s, $c_{\lambda}$'s, $d_{\lambda}$'s, and eleven strong phase differences $\delta^{ab}_{\lambda}=\delta^{a}_{\lambda}-\delta^{b}_{\lambda}$, $\delta^{cd}_{\lambda}=\delta^{c}_{\lambda}-\delta^{d}_{\lambda}$, $\delta^{ac}_{\lambda}=\delta^{a}_{\lambda}-\delta^{c}_{\lambda}$, $\Delta^{a}_{\parallel}=\delta^{a}_{\perp}-\delta^{a}_{\parallel}$. The expressions of the observables in terms of the theoretical parameters can be found in appendix~\ref{obs}. \section{The $\textbf{a}$, $\textbf{b}$, $\textbf{c}$ amplitudes\label{pol}} In this section, we briefly discuss the relation of the polarization basis to the $a$, $b$ and $c$ amplitudes used in the most general covariant expression for a $D\rightarrow V_1V_2$ decay. We closely follow~\cite{Valencia} in the discussions of this section. Consider the decay $D^{0}\to V_{1}V_{2}$. Angular momentum conservation ($\vec{J}_{D}=\vec{L}_{V_{1}V_{2}}+\vec{S}_{V_{1}V_{2}}$), dictates that $\vec{L}_{V_{1}V_{2}}$ can be 0, 1, 2. Thus, one obtains three independent amplitudes corresponding to the three different $\vec{L}_{V_{1}V_{2}}$ values. The most general covariant amplitude for a $D^{0}\to V_{1}V_{2}$ decay can be written as~\cite{Valencia,Lipkin} \begin{align}\label{covariant amplitude} M_{\lambda_{1}\lambda_{2}}=&<V_{1}(\lambda_{1})V_{2}(\lambda_{2})|H_{wk}|D>\nonumber\\ =&\epsilon_{1\mu}^{*}\epsilon_{2\nu}^{*}\bigg[ag^{\mu\nu}+\frac{b}{m_{1}m_{2}}p_{2}^{\mu}p_{1}^{\nu}+i\frac{c}{m_{1}m_{2}}\epsilon^{\mu\nu\alpha\beta}p_{1\alpha}p_{2\beta}\bigg], \end{align} Here $\epsilon_{1}$, $\epsilon_{2}$ represent the polarization vectors and $m_{1}$, $m_{2}$ the masses of the vector mesons $V_{1}$ and $V_{2}$ respectively. The invariant amplitudes $a$, $b$, $c$ each carry the dimension of energy. The corresponding decay rate in terms of the $a$, $b$ and $c$ amplitudes is given by \begin{equation}\label{gamma_cov} \Gamma\left(D\to V_{1}V_{2}\right)=\frac{\left|\bf k\right|}{8\pi m_{D}^{2}}\left(2|a|^{2}+|xa+(x^{2}-1)b|^{2}+2(x^{2}-1)|c|^{2}\right], \end{equation} where $\left|\bf k\right|$ is the decay momentum and $x=\frac{p_{1}.p_{2}}{m_{1}m_{2}}=\frac{m_{D}^{2}-m_{1}^{2}-m_{2}^{2}}{2m_{1}m_{2}}$. The $D^{0}\to V_{1}V_{2}$ amplitude can also be written in the linear polarization basis as, \begin{align}\label{linear polarisation amplitude} A\left(D^{0}\to V_{1}V_{2}\right)=A_{0}\epsilon_{1}^{*L}\epsilon_{2}^{*L}-A_{\parallel}\vec{\epsilon}_{1}^{*T}.\vec{\epsilon}_{2}^{*T}/\sqrt{2} -iA_{\perp}\vec{\epsilon}_{1}^{*}\times\vec{\epsilon}_{2}^{*}.\hat{p}_{2}/\sqrt{2}, \end{align} where $\epsilon_{i}^{*L}=\vec{\epsilon}_{i}^{*}.\hat{p}_{2}$, $\vec{\epsilon}_{i}^{*T}=\vec{\epsilon}_{i}^{*}-\epsilon_{i}^{*L}\hat{p}_{2}$, $\hat{p}_{2}$ is a unit vector in the direction of the momentum of the meson $V_{2}$ in the $V_{1}$ rest frame. Comparing eqn.(\ref{linear polarisation amplitude}) to eqn.(\ref{amplitude}), it can easily be verified that $g_{0}=\epsilon_{1}^{*L}\epsilon_{2}^{*L}$, $g_{\parallel}=-\frac{1}{\sqrt{2}}\epsilon_{1}^{*T}.\epsilon_{2}^{*T}$ and $g_{\perp}=-\frac{1}{\sqrt{2}}\epsilon_{1}^{*}\times\epsilon_{2}^{*}.\hat{p}_{2}$. The corresponding decay width in the linear polarization basis is given by \begin{equation}\label{gamma_pol} \Gamma\left(D^{0}\to V_{1}V_{2}\right)=\frac{\left|\bf k\right|}{8\pi m_{D}^{2}}\left(|A_{0}|^{2}+|A_{\parallel}|^{2}+|A_{\perp}|^{2}\right). \end{equation} From eqns.(\ref{gamma_cov}) and (\ref{gamma_pol}), it is evident that the two bases are related as \begin{eqnarray} A_{\parallel}=\sqrt{2}a,\,\,\,A_{0}=-ax-b(x^{2}-1),\,\,\,A_{\perp}=\sqrt{2(x^{2}-1)}c. \end{eqnarray} \section{Results and discussions\label{results}} A priori, the dependence of the observables on the theoretical parameters may not be trivial as is evident from appendix~\ref{obs}. However, a careful and systematic study of these observables enables one to propose multiple ways for the identification and extraction of new physics as will be shown in this section. In what follows, $\lambda=\{0,\parallel,\perp\}$ and $i=\{0,\parallel\}$. Let us look at the SM case first. For this case the NP decay amplitudes $b_{\lambda}$, $d_{\lambda}$, the NP strong phases $\delta^{b}_{\lambda}$, $\delta^{d}_{\lambda}$ and the NP weak phases $\phi$ and $\chi$ and $\alpha_{NP}$ are all equal to 0. With these values, we get the following relations between the observables \begin{align} &X^{V_1V_2}_{\lambda\lambda}=\bar{X}^{\bar{V_1}\bar{V_2}}_{\lambda\lambda},\; X^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}=\bar{X}^{V_1V_2}_{\lambda\lambda},\; X^{V_1V_2}_{0\parallel}=\bar{X}^{\bar{V_1}\bar{V_2}}_{0\parallel},\label{SM_1}\\ &X^{\bar{V_1}\bar{V_2}}_{0\parallel}=\bar{X}^{V_1V_2}_{0\parallel},\; X^{V_1V_2}_{i\perp}=-\bar{X}^{\bar{V_1}\bar{V_2}}_{i\perp},\; X^{\bar{V_1}\bar{V_2}}_{i\perp}=-\bar{X}^{V_1V_2}_{i\perp}\label{SM_2}. \end{align} It is possible to construct linear combinations of the $Y$'s which depend directly on the phases $\alpha_{SM}$ and the decay phase difference $\delta_{\lambda}^{ac}$ as follows: \begin{align} &1+\frac{Y_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+Y_{\lambda \lambda }^{V_1 V_2}}{r^2 \bar{Y}_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+r^2 \bar{Y}_{\lambda \lambda }^{V_1 V_2}}=\frac{2y\cos\alpha_{SM}}{y\cos\alpha_{SM}+x\sin\alpha_{SM}}\label{SM_3}\\ &\frac{r^2 \bar{Y}_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+r^2 \bar{Y}_{\lambda \lambda }^{V_1 V_2}+Y_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+Y_{\lambda \lambda }^{V_1 V_2}}{r^2 \bar{Y}_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}-r^2 \bar{Y}_{\lambda \lambda }^{V_1 V_2}-Y_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+Y_{\lambda \lambda }^{V_1 V_2}}=\frac{x}{y}\tan\delta_{\lambda}^{ac}\label{SM_4} \end{align} Eqn.(\ref{SM_3}) is of particular interest. The right hand side of the equation is known to appreciable accuracy for the SM, and so is $r$ on the left hand side. The observables $Y$ can be measured experimentally. Hence this relation can be used as a smoking gun signal for detecting the presence of NP. One can go further and identify whether the NP manifests itself in pure decays or as a CP violating effect in mixing or both. For instance, suppose there is no NP in pure decays but some signal for CP violation has been observed. In that case, the formalism dictates that $b_{\lambda}=d_{\lambda}=\delta^{b}_{\lambda}=\delta^{d}_{\lambda}=\phi=\chi=0$, but $\alpha_{NP}\neq0$. This case is very similar to the SM case discussed above. In particular the relations (\ref{SM_1}), (\ref{SM_2}) and (\ref{SM_4}) hold exactly in the same way. This is because the $X$'s do not depend on the (small) NP phase $\alpha_{NP}$ at all, and in the other case the $\alpha_{NP}$ dependence gets canceled in the left hand side of eqn.(\ref{SM_4}). However eqn.(\ref{SM_3}) is modified to \begin{equation} 1+\frac{Y_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+Y_{\lambda \lambda }^{V_1 V_2}}{r^2 \bar{Y}_{\lambda \lambda }^{\bar{V_1} \bar{V_2}}+r^2 \bar{Y}_{\lambda \lambda }^{V_1 V_2}}=\frac{2y}{y+x\tan(\alpha_{SM}+\alpha_{NP})}\label{no_decay_1}\\ \end{equation} Eqn.(\ref{no_decay_1}) along with relations of the type \begin{equation} Y^{V_1V_2}_{\lambda\lambda}+r^2\bar{Y}^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}+Y^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}+r^2\bar{Y}^{V_1V_2}_{\lambda\lambda}=4y\frac{\cos\alpha_{SM}+\alpha_{NP}}{y\cos\alpha_{SM}+\alpha_{NP}-x\sin\alpha_{SM}+\alpha_{NP}}\frac{Y^{V_1V_2}_{\lambda\lambda}+Y^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}{\sqrt{X^{V_1V_2}_{\lambda\lambda}X^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}}. \end{equation} can be simultaneously solved for the determination of $r$ and $\alpha_{NP}$ in this case. We next look into the case where NP is manifested only in pure decays. In this case, $\alpha_{NP}=0$, but $b_{\lambda}$, $d_{\lambda}$, $\delta^{b}_{\lambda}$, $\delta^{d}_{\lambda}$, $\phi$, $\chi$ $\neq0$. For these values of the parameters, we have the following relations among the observables $X$ and $Y$: \begin{equation} \frac{x^2({Y^{V_1V_2}_{\lambda\lambda}}+{\bar{Y}^{V_1V_2}_{\lambda\lambda}})^2+y^2({Y^{V_1V_2}_{\lambda\lambda}}-{\bar{Y}^{V_1V_2}_{\lambda\lambda}})^2}{x^2({Y^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}+{\bar{Y}^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}})^2+y^2({Y^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}-{\bar{Y}^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}})^2}=\frac{\bar{X}^{V_1V_2}_{\lambda\lambda}{X^{V_1V_2}_{\lambda\lambda}}}{{\bar{X}^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}{X^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}}.\label{no_mixing_1} \end{equation} It would be appropriate here to point out relations of the type \begin{eqnarray} &{X^{V_1V_2}_{\lambda\lambda}}-{\bar{X}^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}=8\sin\phi\sin\delta^{ab}_{\lambda}a_{\lambda}b_{\lambda},\label{no_mixing_2}\\ &{X^{\bar{V_1}\bar{V_2}}_{\lambda\lambda}}-{\bar{X}^{V_1V_2}_{\lambda\lambda}}=8\sin\chi\sin\delta^{cd}_{\lambda}c_{\lambda}d_{\lambda}\label{no_mixing_3}. \end{eqnarray} Note that a non zero value of either of the above linear combinations is again a smoking gun signal for NP, since that will mean that the NP weak phases ($\phi$, $\chi$) and the NP weak decay amplitudes ($b_{\lambda}$, $d_{\lambda}$) $\neq0$. Eqns.(\ref{no_mixing_1}), (\ref{no_mixing_2}) and (\ref{no_mixing_3}) hold in the most general case also were NP is present both in pure decays and mixing. The extraction of $\mid\frac{q}{p}\mid$ in the most general scenario can be obtained from relations of the type: \begin{equation} x^2({Y^{V_1V_2}_{\lambda\lambda}}+{r^2\bar{Y}^{V_1V_2}_{\lambda\lambda}})^2+y^2({Y^{V_1V_2}_{\lambda\lambda}}-{r^2\bar{Y}^{V_1V_2}_{\lambda\lambda}})^2=4r^2x^2y^2{\bar{X}^{V_1V_2}_{\lambda\lambda}}{X^{V_1V_2}_{\lambda\lambda}}. \end{equation} Let us now discuss if these observables can be used for the extraction of all the parameters. It is appropriate at this point to look into the case of final states in $D^0$ decays that are CP eigenstates. These include modes like $D^0\rightarrow\phi\omega$, $D^0\rightarrow\rho^0\phi$, $D^0\rightarrow\rho^0\omega$, $D^0\rightarrow\rho\rho$, $D^0\rightarrow\omega\omega$ etc. For these modes, $O^{V_1V_2}=O^{\bar{V_1}\bar{V_2}}$ where $O=\{X, \bar{X}, Y, \bar{Y}\}$ for all combinations of the polarization indices. Therefore the number of observables reduces from 48 to 24. However there are 28 parameters. Discarding $\alpha_{SM}$ (which is precisely known already) and $r$ (since the NP effect, if present, must be very small, it is a reasonable approximation to incorporate the complete effect of NP for the ratio $q/p$ into the phase $\alpha_{NP}$ without changing $r$, which, then, is again precisely known) still leaves 26 parameters to be fitted to 24 observables. Hence, the numerical estimates for all the parameters cannot be obtained for final states that are CP eigenstates without some approximations to reduce the number of parameters. However, relations like (\ref{no_mixing_2}) (eqn.(\ref{no_mixing_2}) and eqn.(\ref{no_mixing_3}) are same in this case) where linear combinations of experimentally measured observables vanish in the absence of new physics can still be devised as tests for the detection of NP effects. For a complete estimation of all the parameters, one has to look at final states which are non CP eigenstates, i.e. modes involving $K^{*0}$ as one of the final states. At present, there is serious dearth of data in the $D\rightarrow VV$ sector. The most well measured mode is $D^0\rightarrow\bar{K^{*0}}\rho^0$ where there are estimates for the total branching fraction, the total transverse wave, S-wave, longitudinal S-wave, P-wave and D-wave branching fractions~\cite{pdg}. The P-wave and longitudinal branching fraction are limits. This gives one six branching fractions which can be written in terms of the theoretical parameters. With the corresponding measurements of $D^0\rightarrow K^{*0}\rho^0$, $\bar{D^0}\rightarrow K^{*0}\rho^0$ and $\bar{D^0}\rightarrow \bar{K^{*0}}\rho^0$ (which are well within the present experimental reach), one has 24 branching fractions written in terms of 27 parameters (discarding $\alpha_{SM}$). It is hence not possible to obtain numerical estimates for all the 27 theoretical parameters with the present data set. However with the measurement of another observable in the future, (say the total longitudinal wave), it is possible to extract numerical values for all the theoretical parameters from a fit to data. With the present data set, one can find numerical estimates for the absolute values of the $A_0$, $A_{\parallel}$ and the $A_{\perp}$ amplitudes (and the corresponding observables $X_{\lambda\lambda}$ from the first relation in (\ref{pol_obs})) and the interference phase between $A_0$ and $A_{\perp}$ for the $D^0\rightarrow\bar{K^{*0}}\rho^0$ mode. We have extracted the amplitudes and the interference phase (which we denote by $\theta$) via a numerical fit using the package HEPfit~\cite{hepfit}. The values of the amplitudes and the phase are displayed in table~\ref{tab1}. \begin{table}[ht] \captionsetup{justification=raggedright, singlelinecheck=false } \centering \begin{tabular}{|c|c|}\hline ~Amplitude~&~Values~\\ \hline\hline $\mid A_0\mid$ &$(4.758\pm3.052)\times10^{-7}$\\\hline $\mid A_\parallel\mid$&$(8.025\pm4.724e)\times10^{-7}$\\\hline $\mid A_\perp\mid$ &$(2.0671\pm0.19355)\times10^{-6}$\\\hline $\theta$ &$(0.00007\pm0.33)^{\circ}$\\\hline \hline \end{tabular} \caption{The absolute values for the longitudinal ($A_0$), parallel transverse ($A_\parallel$) and perpendicular transverse ($A_\perp$) amplitudes and the interference phase between the longitudinal and the perpendicular transverse amplitudes ($\theta$) for $D\rightarrow\overline{K^{*0}}\rho^0$. The values have been extracted from a fit to data. The data is taken from~\cite{pdg}.} \label{tab1} \end{table} \begin{table}[ht] \captionsetup{justification=raggedright, singlelinecheck=false } \centering \begin{tabular}{|c|c|c|}\hline ~Branching ratios~&~Our values~&~Experimental values~\\ \hline\hline S-wave &$(2.79\pm0.51)\%$&$(3.00\pm0.60)\%$\\\hline D-wave &$(2.21\pm0.46)\%$&$(2.10\pm0.60)\%$\\\hline Transverse &$(1.55\pm0.26)\%$&$(1.70\pm0.60)\%$\\\hline Total &$(1.65\pm0.24)\%$&$(1.57\pm0.35)\%$\\\hline \hline \end{tabular} \caption{The total branching ratio and the branching ratios for the various partial waves for the $D\rightarrow\overline{K^{*0}}\rho^0$ mode. The values due to our fit are displayed in the second column and the experimental values are displayed in the last column. All the experimental values are taken from~\cite{pdg}.} \label{tab2} \end{table} The corresponding values thus obtained for the branching ratios are given in table~\ref{tab2}. These values are consistent with experiment within errors. \section{Summary and conclusions\label{summary}} We present a model independent formalism for the detection of NP in $D^0\rightarrow VV$ decays, which involves a time-dependent analysis of the neutral $D$ mesons. This formalism has been previously used to analyze the $B\rightarrow VV$ decays. However the form of the observables change due to the different parametrization of the time dependent $D^0$ decays. We show that it is possible to construct combinations of these observables which are smoking gun signals for NP. We discuss the relations between the observables in the different cases of the SM, NP in decay, NP in mixing and the most general case where NP manifests itself both in mixing and decay. We also extract the absolute values for the longitudinal, the transverse parallel, the transverse perpendicular amplitudes and the interference phase between the longitudinal and the transverse perpendicular amplitudes for the $D^0\rightarrow\overline{K^{*0}}\rho^0$ mode. We find that the interference phase is consistent with zero. \acknowledgements We thank Nita Sinha and Rahul Sinha for fruitful discussions regarding this work.
c7fa8fb99e22c18b02a879a5b5694581b7c02ef4
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:introduction} The cytoplasmic membrane separates the living cell from its extra-cellular surroundings, while other intra-cellular membranes compartmentalize cellular organelles. Biomembranes are constructed from two monolayers (leaflets) in a back-to-back arrangement, and are in general asymmetric in their lipids composition~\cite{Devaux,Meer}. For example, in human red blood cells, the inner cytoplasmic leaflet is composed mostly of phosphatidylethanolamine (PE) and phosphatidylserine (PS), while the outer cytoplasmic leaflet is composed of phosphatidylcholine (PC), sphingomyelin (SM) and a variety of glycolipids~\cite{Opden,Devaux1}. The asymmetric nature of the cell membrane plays a key role in a variety of cellular processes such as endocytosis~\cite{Pomo}, vesicle budding and trafficking~\cite{Dal1}. Furthermore, in living cells, the composition asymmetry is an active and energy-consuming process. It is maintained by several membrane proteins such as flippase and floppase that allow lipids to exchange between the two leaflets with the aid of adenosine triphosphate (ATP)~\cite{Dal}. In artificial multi-component lipid bilayers, the two membrane leaflets can undergo a lateral phase separation. Several authors made the connection between such a phase separation in artificial membranes and existence of small dynamic domains (``rafts") in biological membranes~\cite{KA14}. It should be noted, however, the size of rafts in biological membranes are expected to be in the range of 10--100~nm~\cite{Simons,Munro}. Raft are believed to be enriched mixtures of cholesterol and SM in a liquid-ordered phase (L$_{\rm o}$), embedded in a background of a liquid-disordered phase (L$_{\rm d}$). Despite the lack of an ultimate proof for the existence of rafts, they have been advocated in relation with their potential influence on biological cellular processes. It has been suggested that rafts act as organizing centers for the assembly of signaling molecules, influencing membrane fluidity, and regulating receptor trafficking~\cite{Simons,Munro}. \begin{figure*}[tbh] \begin{center} \includegraphics[scale=0.4]{fig1.eps} \end{center} \caption{\textsf{ Schematic vertical cut through a membrane consisting of several domains. The domains are embedded in a lipid matrix (yellow) and have a line tension acting along the domain rim. Three domain types can be seen and are further explained in Fig.~2: Flat (F), Budded (B) and Dimpled (D). Each domain is formed by two lipids (red and blue) that partition themselves differently on the two domain leaflets. }} \label{fig1} \end{figure*} A theoretical model for domain-induced budding of planar membranes was proposed by Lipowsky~\cite{Lipowsky92} some years ago, and later was extended for closed vesicles~\cite{JL93,JL96}. In the model, the competition between the membrane bending energy and domain line tension leads to a budding transition. More recently, a model describing domain-induced budding in bilayers composed of a binary mixture of lipids was proposed by us~\cite{Wolff}. In particular, we have shown that dimpled domains are formed and remain stable due to an asymmetry between the two compositions of the corresponding domain leaflets, given that the line tension along the domain rim is not too large. The calculations in Ref.~\cite{Wolff} were done for a specific case where the relative concentration between the two lipids on the bilayer domain stays constant, while the lipids are allowed to freely exchange between the two leaflets. In the last decade, however, techniques such as Langmuir-Blodgett or Langmuir-Sch\"{a}efer have enabled a control over the asymmetric composition of artificial membranes~\cite{Kiessling, Cheng}. For example, unsupported bilayers via the Montal-M\"uller method have been used to form asymmetric bilayers in which the composition of each leaflet was independently controlled~\cite{Collins}. A key ingredient in understanding those experiments is the fact that the flip-flop process that exchanges lipids across the leaflets is slower than experimental times. Given these experimental findings, it is worthwhile to consider bilayers where each of the leaflet domain composition (rather than the overall bilayer composition) can be controlled in a separate and independent way. In this paper, we generalize our previous budding model~\cite{Wolff} and extend it to asymmetric two-component lipid domains. Each domain leaflet has a conserved lipid composition that is independent from that of the other leaflet. We consider the possibility of domains curved in the third dimension, which can produce buds as shown in Fig.~\ref{fig1}. In our model, the composition-dependent spontaneous curvature leads to coupling between curvature and lipid composition in each of the domain leaflets~\cite{Leibler,SPA,MS}. Such a coupling leads to rich phase behavior including various phase coexistence regions. For the sake of clarity, we do not take into account any direct interaction between two domains that face each other, although various possibilities have been previously proposed~\cite{May}. In addition, we would like to mention that in a separate set of studies for completely planar membranes, asymmetric bilayers composed of two modulated monolayers (leaflets) were considered~\cite{HKA09,HKA12}. In these works, the static and dynamic properties of concentration fluctuations have been presented together with the related micro-phase separation. Our model may have several experimental implications. We predict that bilayer domains can exist in three states having different equilibrium shapes: fully budded, dimpled, and flat states. Their relative stability depends on controlled system parameters: temperature, degree of compositional asymmetry between the two leaflets and domain size. We find that the dimpled state is the most stable one in some of the parameter range and this is in accord with recent experiments~\cite{RUPK}. Based on the calculated phase diagrams, we anticipate that membranes should exhibit in some parameter range two-phase and three-phase coexistence between different domain states. The outline of this paper is as follows. Section~\ref{sec:model} generalizes our previous budding model~\cite{Wolff}, and the free-energy describing asymmetric two-component lipid domains is discussed. In Sec.~\ref{sec:diagram}, we explain the conditions for various phase equilibria and how to calculate the phase diagrams. We then proceed by presenting the phase diagrams in Sec.~\ref{sec:result}, and discuss the resulting global phase behavior. Finally, a more qualitative discussion is provided in Sec.~\ref{sec:discussion}. \section{Model} \label{sec:model} \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.4]{fig2.eps} \end{center} \caption{\textsf{ A schematic vertical cut through three possible domain shapes: (a) a flat bilayer domain composed of a mixture of A and B lipids (red and blue, respectively), and embedded in an otherwise flat membrane (yellow). The circular flat domain (F) has a radius $L$ and area $S=\pi L^2$. (b) A partial bud (dimpled domain) curved in the out-of-plane direction. The bud (D) of the same area $S$ forms a spherical cap of radius $1/C$, where $C$ is the curvature. (c) A fully-budded domain (B) has a spherical shape of total area $S$, and just touches the flat membrane. The line tension $\gamma$ acts along the line boundary between the domain (red/blue lipids) and the flat membrane matrix (yellow lipids). }} \label{fig2} \end{figure} We consider a membrane consisting of two monolayers (the terms ``monolayer" and ``leaflet" will be used interchangeably in this paper), each composed of an A/B mixture of lipids, which partition themselves asymmetrically between the two leaflets. We assume that the membrane can undergo a lateral phase separation creating domains of different lipid composition. The domains are taken to fully span the two monolayers, but the leaflet compositions in these domains can be different. Hence, the formed lipid domains are, in general, asymmetric. We further assume that the two leaflet compositions are conserved and can be taken as independent from each other. In the following, we discuss the thermodynamic behavior of a system consisting of a large number of such asymmetric domains as shown in Fig.~\ref{fig1}. For this purpose, we explain below the different terms that contribute to the domain free-energy. In Fig.~\ref{fig2}, we show a vertical cut through three possible domain states: flat (F), dimpled (D) and fully-budded (B). The flat circular domain (F) in (a) has an area, $S=\pi L^2$, which is assumed to remain constant during the budding process. For simplicity, we consider in (b) dimpled buds (D) whose shape is a spherical cap of radius $1/C$, and in (c) the extreme case of a completely detached spherical bud (B). The total bending energy of the budded domain is given by the curvature contributions from its two monolayers~\cite{Helfrich73,SafranBook}: \begin{equation} E_{\rm bend} =2 \pi L^2 \kappa \left[ (C -C_0)^2 + (C + C_0)^2 \right], \label{bend} \end{equation} where $\kappa$ is the bending rigidity modulus (assumed to be a constant that is independent of lipid composition) and $C_0$ the monolayer spontaneous curvature. As shown in Fig.~\ref{fig2}, the two monolayer deform with curvatures $+C$ and $-C$, respectively. The second energy contribution is the domain edge energy that is proportional to the perimeter length and to the line tension, $\gamma$~\cite{Lipowsky92}: \begin{equation} E_{\rm edge} = 2 \pi L \gamma\sqrt{1-(LC/2)^2}. \label{edge} \end{equation} Note that in the extreme case, when the domain buds into a complete spherical shape (B) as in Fig.~\ref{fig2}(c), $C=\pm 2/L$ and $E_{\rm edge}=0$. In the above, the strong variation in composition between the domain and its surrounding matrix is effectively taken into account through the line tension, $\gamma$, which is treated as an external control parameter. This situation can be justified for a strong segregation that results in a sharp boundary between the domain and its flat matrix surroundings. As each domain is composed of an A/B mixture, we define $\phi_{\rm A}$ ($\phi_{\rm B}$) as the area fraction (assumed to be equal to the molar fraction) of the ${\rm A}$ lipid ($\rm B$ lipid) in the upper leaflet domain, and similarly, $\psi_{\rm A}$ ($\psi_{\rm B})$ for the lower leaflet domain. We assume that each monolayer is incompressible so that $\phi_{\rm A}+\phi_{\rm B}=1$ and $\psi_{\rm A}+\psi_{\rm B}=1$. Hence, the two relevant order parameters are the relative composition in the upper leaflet: \begin{equation} \phi= \phi_{\rm A}-\phi_{\rm B} , \end{equation} and in the lower leaflet, \begin{equation} \psi= \psi_{\rm A}-\psi_{\rm B} . \label{def} \end{equation} As in any A/B mixture, the possibility of a phase separation can be described by a phenomenological Landau expansion of the free-energy in powers of $\phi$ and $\psi$. This expansion is done separately for each monolayer, and the total contribution to the free-energy is the sum over the two monolayers: \begin{equation} E_{\rm phase} = \pi L^2 \frac{U}{\Xi^2} \left[ \frac{t}{2} (\phi^2 +\psi^2)+ \frac{1}{4} (\phi^4+\psi^4) \right], \label{phase} \end{equation} where $\Xi \equiv \kappa/\gamma$ is the invagination length, $U$ is a parameter that sets the energy scale, and $t \sim (T-T_{\rm c})/T_{\rm c}$ is the reduced temperature ($T_{\rm c}$ being the critical temperature). In Eq.~(\ref{phase}) above, we multiply by the domain area, $\pi L^2$, to obtain the domain free-energy. Hereafter, we will use several dimensionless variables: a rescaled curvature $c \equiv LC$, rescaled spontaneous curvature $c_0 \equiv LC_0$, and rescaled invagination length $\xi\equiv \Xi/L$. The coupling between curvature and composition is taken into account by assuming a linear dependence of the spontaneous curvature $c_0$ on the relative composition in each of the leaflets~\cite{Leibler,SPA,MS}: \begin{equation} c_0(\phi) =\bar{c}_0-\beta \phi, \label{spontaneous1} \end{equation} \begin{equation} c_0(\psi) =\bar{c}_0-\beta \psi, \label{spontaneous2} \end{equation} where $\bar{c}_0$ is the monolayer spontaneous curvature for the symmetric 1:1 composition, $\phi=\psi=0$, and $\beta$ is a coupling parameter that has the same value for the two monolayers. Since $\bar{c}_0$ is a constant, it merely shifts the origin of the chemical potential, and will be dropped out without loss of generality. The total free-energy per domain is then given by the sum of Eqs.~(\ref{bend}), (\ref{edge}) and (\ref{phase}): $E_{\rm tot}=E_{\rm bend}+E_{\rm edge} + E_{\rm phase}$, and its dimensionless form, $\varepsilon=E_{\rm tot}/2\pi \kappa$, is expressed as \begin{align} \varepsilon(\phi, \psi, c)& =(c+\beta\phi)^2+(c-\beta\psi)^2 +\frac{1}{\xi}\sqrt{1-c^2/4} \nonumber \\ &+\frac{1}{\xi^2} \left( \frac{U}{2 \kappa} \right) \left[\frac{t}{2}(\phi^2+\psi^2)+\frac{1}{4}(\phi^4+\psi^4) \right]. \label{eq6} \end{align} We note that Eq.~(\ref{eq6}) depends on three dimensionless parameters: $\beta$, $\xi$, and $U/(2 \kappa)$, while the thermodynamic variables are the reduced temperature $t$ and the three order parameters: $\phi$, $\psi$ and $c$. In the calculations presented hereafter, we set $U/(2 \kappa)=1$ and vary the values of $\beta$ and $\xi$. Within mean-field theory, the equilibrium states and phase transitions are determined by minimizing $\varepsilon$ with respect to $\phi$, $\psi$ and $c$, under the condition that $\phi$ and $\psi$ are conserved order parameters while $c$ is not. From the minimization of $\varepsilon$ with respect to $c$, we obtain the condition \begin{equation} 2(c+\beta\phi)+2(c-\beta\psi)-\frac{c}{4\xi\sqrt{1-c^2/4}}=0. \label{eq7} \end{equation} Then, by substituting the curvature $c=c(\phi, \psi)$ into Eq.~(\ref{eq6}), results in a partially minimized free-energy, $\varepsilon^*$ \begin{equation} \varepsilon^{\ast}(\phi, \psi) = \varepsilon(\phi, \psi, c(\phi, \psi)), \label{epsast} \end{equation} as a function of $\phi$ and $\psi$. Typical experimental values of domain sizes are in the range of $L\simeq 50$--$500$~nm~\cite{Simons}, the bending rigidity $\kappa \simeq 10^{-19} {\rm J}$~\cite{SafranBook}, and the line tension $\gamma \simeq 0.2$--$6.2 \times 10^{-12}$~J/m~\cite{baumgart,tian}. Hence, the scaled invagination length is estimated to be in the range, $\xi \simeq 0.01$--$10$. These values will be used in the next section where we calculate numerically the phase diagrams. \section{Phase equilibria conditions} \label{sec:diagram} In order to obtain various phase coexistence regions, $\varepsilon^{\ast}$ in Eq.~(\ref{epsast}) should be further minimized with respect to the conserved order parameters, $\phi$ and $\psi$. Hence, we consider the following thermodynamical potential \begin{equation} g(\phi, \psi)= \varepsilon^{\ast}(\phi,\psi)-\mu_{\phi}\phi-\mu_{\psi}\psi, \label{grand} \end{equation} where $\mu_{\phi}$ and $\mu_{\psi}$ are the chemical potentials coupled with the A/B relative compositions in the upper and lower domains, respectively. They act as Lagrange multipliers that take into account the conserved $\phi$ and $\psi$ compositions. In general, these two chemical potentials have different values, $\mu_{\phi} \ne \mu_{\psi}$. The special case of $\mu_{\phi} = \mu_{\psi}$, for which only the total relative composition, $\phi + \psi$, is conserved was investigated in our previous work~\cite{Wolff}, while here we deal with a general situation where each of the compositions, $\phi$ and $\psi$, are conserved independently. \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.4]{fig3.eps} \end{center} \caption{\textsf{ (a) Plot of the partially minimized free-energy $\varepsilon^{\ast}(\phi, \psi) $ as a function of $\phi$ and $\psi$ for $\xi=0.25$, $\beta=1$ and $t=-0.2$ (below $T_{\rm c}$). (b) A cut through the free-energy landscape $\varepsilon^{\ast}$ in the particular direction, $\phi+\psi=0$, plotted as a function of $\Phi = (\phi-\psi)/2$. The two P points are cusps at which the slope of $\varepsilon^*$ changes discontinuously, although the Hessian $H$ remains positive (see Appendix). The two Q points correspond to the location at which $H$ vanishes. }} \label{fig3} \end{figure} The thermodynamic equilibrium between the two coexisting phases denoted as `1' and `2' and characterized by $(\phi_1, \psi_1)$ and $(\phi_2, \psi_2)$, satisfies the conditions~\cite{Gibbs} \begin{eqnarray} \left. \partial_\phi g(\phi,\psi)\right|_1&=& \left. \partial_\phi g(\phi,\psi)\right|_2=0,\nonumber\\ \left. \partial_\psi g(\phi, \psi)\right|_1&=& \left. \partial_\psi g(\phi,\psi)\right|_2=0,\nonumber\\ g(\phi_1, \psi_1)&=& g(\phi_2, \psi_2). \label{eq123} \end{eqnarray} Similarly, for a three-phase coexistence between phases `1', `2' and `3', the following set of conditions should be satisfied: \begin{eqnarray} \label{3phase} \left. \partial_\phi g(\phi,\psi)\right|_1&=& \left. \partial_\phi g(\phi, \psi)\right|_2= \left. \partial_\phi g(\phi, \psi)\right|_3=0\nonumber\\ \left. \partial_\psi g(\phi, \psi) \right|_1&=& \left. \partial_\psi g(\phi, \psi) \right|_2= \left. \partial_\psi g(\phi, \psi) \right|_3=0\nonumber\\ g(\phi_1, \psi_1)&=& g(\phi_2, \psi_2)=g(\phi_3, \psi_3). \end{eqnarray} In Fig.~\ref{fig3}(a), we show an example of the partially minimized free-energy, $\varepsilon^{\ast}(\phi,\psi)$, as a function of $\phi$ and $\psi$ at a fixed temperature $t=-0.2$ (below $T_{\rm c}$), and for given values of $\xi$ and $\beta$. In order to have a better view of the free-energy surface, we show in Fig.~\ref{fig3}(b) a cross-section cut of the free-energy surface in the direction of $\Phi = (\phi-\psi)/2$, while keeping $\phi+\psi=0$. Here, we see two singular cusps (points P) where the determinant of the Hessian matrix, $H$ (see Appendix) does not vanish. At these cusps, $H$ changes discontinuously although it remains positive. At points Q, on the other hand, the Hessian $H$ vanishes. More details on the Hessian matrix and determinant, and their relation to the phase stability and spinodal lines are presented in the Appendix. By calculation the Hessian $H$, it is possible to derive the spinodal lines and critical points. Note that in most cases, the critical points and spinodal will not be shown on the phase diagrams, because they are preempted by the first-order phase transition lines and coexistence regions. The phase diagrams are obtained by further minimizing $\varepsilon^{\ast}$, with respect to the two independent variables, $\phi$ and $\psi$. Convex regions of the free energy correspond to single thermodynamical phases. Two-phase coexistence regions correspond to non-convex regions, where we can construct a common tangent plane. The plane touches the free-energy surface at two points that determine the two phases in coexistence. A more special three-phase coexistence region corresponds to a plane that touches the free-energy surface at three points. More details on the numerical procedure of finding the phase diagrams are given below. The numerical computation of the phase diagram is performed using a public-domain software called ``Qhull"~\cite{qhull}. The Qhull software generates initially a fine grid of triangulation, which approximates the free-energy surface, $\varepsilon^{\ast}$. The three-phase coexistence regions correspond to facets with all sides being much larger than the initial discretization. The two-phase coexistence regions are associated with elongated triangles having one short side that is much smaller than the other two longer sides that approximate the tie-lines. Finally, small triangular facets of the free-energy surface are associated with stable one-phase regions. The projection of the triangulated free-energy surface onto the composition plane provides a systematic approximation for the phase diagram. The Qhull results are then used as an initial condition in calculating more precisely the equilibrium phase diagrams, including the various phase coexistence, Eqs.~(\ref{grand})--(\ref{3phase}). \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.5]{fig4.eps} \end{center} \caption{\textsf{ (a) Phase diagram in the ($\phi$, $\psi$) plane for $\beta=1$, $\xi=0.25$ and $t=-0.45$. The corners of the phase diagram indicates the four one-phases: D$_\pm$ and B$_\pm$, while the flat F phase (not drawn) strictly lies only on the diagonal $\phi=\psi$. The black lines represent tie-lines in the two-phase regions, and the two triangles are the three-phase regions (see text for more details). (b) The phase diagram is plotted as in (a) but with a superimposed colored plot for the curvature, $c$. As one crosses the major diagonal, $\phi=\psi$, there is a smooth change from D$_+$ ($c>0$) through the flat F ($c=0$) to the D$_-$ ($c<0$). Furthermore, the curvature also changes smoothly inside the $D_\pm$ phases, but the gradient in orange (D$_-$) and light blue (D$_+$) colors is not shown for clarity. However, the curvature has a jump between B$_-$ ($c=-2$, red) and D$_-$ ($-2< c <0$, orange) regions, as well as between B$_+$ ($c=2$, dark blue) and D$_+$ ($0< c<2$, light blue) ones. }} \label{fig4} \end{figure} \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.5]{fig5.eps} \end{center} \caption{\textsf{ (a) Phase diagram in the ($\phi$, $\psi$) plane as in Fig.~\ref{fig4} but with $\beta=3$, $\xi=0.25$ and $t=-0.45$. (b) The phase diagram is plotted as in (a) but with a superimposed colored plot for the curvature, $c$. }} \label{fig5} \end{figure} \section{Results} \label{sec:result} \subsection{Phase diagrams} \label{sec:diagrams} \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.5]{fig6.eps} \end{center} \caption{\textsf{ (a) Phase diagram in the ($\phi$, $\psi$) plane as in Fig.~\ref{fig4} but with $t=-0.2$, $\xi=0.25$ and $\beta=1$. (b) The phase diagram is plotted as in (a) but with a superimposed colored plot for the curvature, $c$. }} \label{fig6} \end{figure} \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.5]{fig7.eps} \end{center} \caption{\textsf{ (a) Phase diagram in the ($\phi$, $\psi$) plane as in Fig.~\ref{fig4} but with $t=-0.02$, $\xi=0.25$ and $\beta=1$. The red circles correspond to the critical points. (b) The phase diagram is plotted as in (a) but with a superimposed colored plot for the curvature, $c$. }} \label{fig7} \end{figure} Four representative types of phase diagrams are shown in Figs.~\ref{fig4}--\ref{fig7}. In the first two figures, we show the calculated phase diagram for $\beta=1$ (Fig.~\ref{fig4}) and $\beta=3$ (Fig.~\ref{fig5}), while the other parameters are fixed to $\xi=0.25$ and $t=-0.45$. The latter two phase diagrams are for $t=-0.2$ (Fig.~\ref{fig6}) and $t=-0.02$ (Fig.~\ref{fig7}), while keeping $\beta=1$ and $\xi=0.25$. For presentation purposes, in (b) of each figure, we have superimposed the color plot for the curvature $c$ on the phase diagram presented in (a). As denoted above, the various stable phases are the fully-budded phase (B) with curvature $\vert c\vert=2$, the dimpled or partial budded phase (D) with curvature $0<\vert c \vert< 2$, and the flat membrane (F) with zero curvature $c=0$. The subscripts $\pm$ denote whether the bud is curved positively or negatively with respect to the positive normal direction of the planar membrane. Figure~\ref{fig4} includes five homogeneous phases: B$_\pm$, D$_\pm$ and F. The B$_+$ and B$_-$ occupy the regions around the $\phi=-\psi=1$ and $\phi=-\psi=-1$ corners, respectively. On the other hand, the D$_\pm$ occupy the two remaining corners: $\phi=\psi=\pm 1$. The difference between D$_+$ and D$_-$ is only associated with their curvature ($c\lessgtr 0$), which changes continuously. This is further clarified in part (b) of the figure. For all $t$-values, the flat F phase (not seen on the figure) strictly exists only on the $\phi=\psi$ diagonal, but is important for the multi-phase coexistence regions (as discussed below). As one approaches this line from above or below, the D$_\pm$ phases change-over smoothly into the F phase (with $c=0$) on the diagonal line. Moreover, five two-phase coexistence regions are shown in Fig.~\ref{fig4} together with their calculated tie-lines: two B$_+$/D$_+$ and two B$_-$/D$_-$ along the boundaries of the phase diagram, and one B$_+$/B$_-$ along the major diagonal, $\phi=-\psi$. In addition, two three-phase coexistence regions: B$_+$/B$_-$/F can be seen. These are the two triangular regions lying above and below the $\phi=-\psi$ diagonal. Note that the F corner of the three-phase region lies close to the D$_\pm$ corners, but since it lies on the $\phi=\psi$ diagonal, it is identified as the F phase with $c=0$. The three-phase coexistence region between F and B$_\pm$ phases means that each point inside the triangular region is composed of three relative area fractions of the three coexisting phases: the flat (F) and budded (D$_\pm$) phases. The nearly horizontal (D$_-/$B$_-$) or vertical (D$_+/$B$_+$) tie-lines on the boundaries of phase diagram indicate that the two $\phi$ and $\psi$ monolayers are almost decoupled, because either $\phi$ or $\psi$ do not vary along the tie-line. On the other hand, tie-lines that lie along the major diagonal, $\phi=-\psi$, indicate a strong coupling between the $\phi$ and $\psi$ monolayers in the B$_+/$B$_-$ coexistence region. In Fig.~\ref{fig5} with $\beta=3$ and the same temperature as in Fig.~\ref{fig4}, we see that the central binary coexistence region (B$_+$/B$_-$) becomes much larger. On the other hand, the four two-phase coexistence regions of Fig.~\ref{fig4} have shrunken because the extent of the dimpled phase becomes smaller for larger coupling parameter, $\beta$. When $\beta$ gets large values in Eqs.~(\ref{spontaneous1}) and (\ref{spontaneous2}), the composition of each monolayer induces higher curvature that promotes budding. In this case, the central two-phase region, B$_+$/B$_-$, occupies a large fraction of the phase diagram, and most of its tie-lines are parallel to the diagonal $\phi = -\psi$, suggesting a strong coupling between the two monolayers. On the other hand, the two three-phase coexistence regions, B$_+$/B$_-$/F, become smaller in Fig.~\ref{fig5}. Furthermore, we note that in Fig.~\ref{fig5}(b), the regions of the dimpled phases, D$_\pm$, represented by the light blue and orange regions are narrower as compared with those of Fig.~\ref{fig4}(b). It is of interest to explain how one of the monolayers induces a phase transition in the second monolayer. As an illustrative example, Let us consider a point in Figs.~\ref{fig4} and \ref{fig5} with average leaflet composition, $(\phi,\psi)=(0,-0.9)$. At these compositions, the bilayer separates in Fig.~\ref{fig4} into a D$_-$ phase with compositions $(-0.75,-0.9)$ and a B$_-$ phase with $(0.75, -0.9)$. In this weak coupling case, $\beta=1$, the phase separation in one monolayer does not induce any instability leading towards phase separation in the second monolayer, because the tie-line is nearly parallel to the horizontal $\phi$-axis. On the other hand, in Fig.~\ref{fig5} with a larger value of $\beta=3$, the bilayer separates into a B$_-$ phase with $(0.1, -0.9)$ and a B$_+$ phase with $(-0.9, 0.1)$. As the tie-line in this case lies along the major diagonal, $\phi=-\psi$, the two monolayers are influencing each other. Such a situation results from a strong composition-curvature coupling when the parameter $\beta$ is large enough. A more general dependence of the phase diagram on the parameter $\beta$ will be further discussed in the next subsection. Figure~\ref{fig6} is plotted for a higher temperature $t=-0.2$ than in Figs.~\ref{fig4} and {\ref{fig5}, while we fix $\beta=1$ as in Fig.~\ref{fig4}. The dimpled region expands both toward the corners and the middle of the phase diagram. Moreover, there are two new one-phase regions of the dimpled phase (D$_\pm$) resulting in four additional two-phase coexistence regions: two D$_+$/D$_+$ and two D$_-$/D$_-$. The system exhibits a first-order phase transition in composition, while the transition is second-order in curvature. In Fig.~\ref{fig7}, the temperature is increased to $t=-0.02$, while $\beta$ and $\xi$ stay as in Fig.~\ref{fig4}. The chosen $t$-value is higher than in the previous figures, and approaches the critical temperature, $t_{\rm c}=0.04$~\cite{Wolff}. The central region of the phase diagram is dominated by the dimpled phase (D$_\pm$), while the budded regions (B$_\pm$) exist only close to the two corners, with two-phase coexistence regions, B$_+$/D$_+$ and B$_-$/D$_-$. As seen in (b), the regions of the dimpled phases (D$_\pm$) are determined by $\beta$, and do not depend on the temperature $t$ as long as $\xi$ is fixed. Furthermore, there are four critical points appearing in the central region (marked by red circles), for which the system exhibits a second-order phase transition both in curvature and composition. \subsection{Global phase behavior} \label{sec:global} \begin{figure}[tbh] \begin{center} \includegraphics[scale=0.6]{fig8.eps} \end{center} \caption{\textsf{ Schematic phase-behavior plot in the ($1/\xi, \beta$) plane for temperatures (a) $t=-0.25$ and (b) $t=-0.02$. The diagram types, I, II, III and IV, correspond to representative examples as in Figs.~\ref{fig4}, \ref{fig5}, \ref{fig6}, and \ref{fig7}, respectively. }} \label{fig8} \end{figure} Next, we investigate how the different phase-diagram types appear and change as we adjust the system parameters in a global way. In Fig.~\ref{fig8}, we present the global phase behavior in the ($1/\xi, \beta$) plane for $t=-0.25$ in (a) and $t=-0.02$ in (b). We show how the four different types of phase diagrams of Figs.~\ref{fig4}, \ref{fig5}, \ref{fig6}, \ref{fig7}, labelled as I, II, III and IV, respectively, evolve as function of $\xi$ and $\beta$. Figure~\ref{fig8}(a) summarizes the result for $t=-0.25$ and is valid even for lower $t$-values. For $\beta>2$ (strong coupling), the diagram is that of type II in which the dimpled phase (D) region shrinks while the budded phase (B) expands. For $\beta<2$ (weak coupling), on the other hand, the diagram is mostly of type I for which the wide three-phase coexistence occurs. For $\beta<1$ and $1/\xi<4$, type III diagram is found, and contains coexistence regions of the dimpled phases (D$_+$/D$_+$ and D$_-$/D$_-$). In Fig.~\ref{fig8}(b), for $t=-0.02$ that is closer to the critical temperature, type III is replaced by type IV for $\beta<1$ and $1/\xi<5$, while type II also extends to smaller $\beta$ when $1/\xi$ is large enough. When the temperature is higher, the phase transition in curvature from the flat (F) phase to the dimpled (D) one, becomes continuous~\cite{Wolff}. \section{Discussion} \label{sec:discussion} In this paper, we have discussed a phenomenological model that accounts for the budding transition of asymmetric two-component lipid domains, where the two domains have in general different average compositions, represented by $\phi$ and $\psi$. Assuming a linear composition dependence of the spontaneous curvature, we have taken into account a coupling between the local curvature and local lipid composition in each of the two leaflets. We then explored the morphological changes between the flat and budded domains by using a thermodynamic argument. Our free-energy model contains three contributions: bending energy, accounting for domain deformation in the normal direction; line tension along the rim of the budded or flat domain; and a Landau free-energy expansion, which accounts for a phase separation of the two-component lipid domains. We have assumed, in addition, that the domain area remains constant during the budding process. Our model predicts three different states for the domains: fully budded (B), dimpled (D), and flat (F) states. In particular, in some ranges of parameters, the D state is found to be the most stable one, as observed in the experiment~\cite{RUPK}. Within mean-field theory, we have calculated various phase diagrams in terms of two compositions for different temperatures $t$, domain size $1/\xi$, and coupling parameter, $\beta$. The resulting phase behavior is very rich. The calculated phase diagrams include various one-phase, two-phase (e.g., B$_+$/D$_+$, B$_-$/D$_-$ or B$_+$/B$_-$) and three-phase coexistence regions (e.g., B$_+$/B$_-$/F), depending on the curvature-composition parameter $\beta$ as well as the temperature. Finally, four different types of phase diagram morphologies are found, and we have analyzed the global phase behavior in terms of the coupling parameter $\beta$, and the domain size $\xi$. The model analysis suggests that the asymmetry in the lipid composition between the two leaflets can lead to complex morphological and thermodynamic behavior of lipid domains. The most important mechanism that leads to the inter-leaflet correlated phase separation, such as the two-phase coexistence seen in Fig.~\ref{fig5} between two budded phase (B$_+$/B$_-$), is the composition-dependent spontaneous curvature, introduced in Eqs.~(\ref{spontaneous1}) and (\ref{spontaneous2}). With this mechanism, the coupling between domain curvature and its composition is controlled by the parameter $\beta$. Although we did not include any direct interaction between the domains occurring on the two leaflets, the fact that the two domains are in full registry to each other and have opposite curvatures, $\pm C$, results in a strong correlation between the opposing domains. We consider that such an effect generally exists in bio-membranes, even in the absence of specific proteins that maintain the compositional asymmetry between the two leaflets. When the coupling parameter $\beta$ is made larger at a fixed temperature (compare Figs.~\ref{fig4} and \ref{fig5}), the two-phase coexistence region, B$_+$/B$_-$, dramatically expands due to the coupling between the two leaflets. Such a strongly correlated phase separation can be in general observed for lower temperatures. We also note that the stable budded (B) phase occupies the two asymmetric corners of the phase diagram, i.e., $\phi=-\psi=\pm 1$ . Moreover, the region of the dimpled (D) phase increases as the temperature is raised towards the critical temperature from below (compare Figs.~\ref{fig4}, \ref{fig6} and \ref{fig7}). Especially, in Fig.~\ref{fig7}, the dimpled one-phase region dominates the central region of the phase diagram and four different critical points are expected to appear. In the intermediate temperature, as in Fig.~\ref{fig6}, the phase diagram contains several types of two-phase coexistence regions. The present work for asymmetric domains is a generalization of our previous model~\cite{Wolff}, where only the average composition of the domains in the two leaflets is controlled by a single chemical potential. Here, we have considered a more general situation, where each of the two domain composition is controlled in a separate and independent way. This is done by introducing two independent chemical potentials coupled to the two domain compositions, $\phi$ and $\psi$, as in Eq.~(\ref{grand}). Hence, each domain has a conserved lipid composition that is independent from the other one. The resulting phase diagrams show a much richer phase behavior, while the previous results~\cite{Wolff} can be recovered by considering the special case of $\phi+\psi={\rm const.}$ Our assumption that the two opposing domains in different leaflets are correlated to each other is in accord with several experimental observations. For studies of planar but asymmetric composition in the two leaflets, Collins \textit{et al.}~\cite{Collins} reported that in some cases, one leaflet can induce a phase separation in the other leaflet, depending on local lipid composition. Whereas in other cases, the two leaflets do not interact. A similar experimental phenomenon was observed for membranes composed of lipids extracted from biological cells~\cite{Kiessling}. Such situations for asymmetric membranes can be partially described by the different phase behaviors of asymmetric membrane depending on the coupling parameter $\beta$ as shown in Fig.~\ref{fig8}. Moreover, both in our model and in experiments~\cite{Collins}, domain-induced processes take place in lipid membrane without any proteins. These results may suggest a cellular mechanism for regulating protein function by modulating the local lipid composition or inter-leaflet interactions. Although we have mainly discussed the domain-induced budding, our model can be applied to describe the formation of vesicles in mixed amphiphilic systems~\cite{SH}. It was observed in experiment that mixtures of anionic and cationic surfactants in solution form disk-like bilayers in some range of the relative amphiphilic composition. As these disk-shaped bilayers grow in size, they transform into spherical caps and eventually become spherically closed vesicles. In such cases, the spontaneous curvature of bilayer membranes may be induced due to the compositional asymmetry between the two monolayers. Finally, our model suggests that the asymmetry in the lipid composition between the two leaflets leads to a complex behavior of lipid domains even in the absence of any specific enzymes or proteins, which can induce additional coupling between the two leaflets~\cite{Gri04}. The importance of such a pure physical mechanism can be verified in experiments on asymmetric model membranes involving only a lipid mixture. They also can be of relevance to signal transduction~\cite{Simons00}, membrane fusion~\cite{Puri06}, or penetration of viruses into cells~\cite{Chazal03}. We hope that additional experiments will address these issues in the future. \bigskip {\em Acknowledgments.~} J.W.\ acknowledges support from the Service de Coop\'eration Scientifique et Universitaire de l'Ambassade de France en Isra\"el, the French ORT association, and ORT school of Strasbourg. S.K.\ acknowledges support from the Grant-in-Aid for Scientific Research on Innovative Areas ``\textit{Fluctuation and Structure}" (Grant No.\ 25103010) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan, and the Grant-in-Aid for Scientific Research (C) (Grant No.\ 15K05250) from the Japan Society for the Promotion of Science (JSPS). D.A. acknowledges the hospitality of the KITPC and ITP, Beijing, China, and partial support from the Israel Science Foundation (ISF) under Grant No.\ 438/12, the United States--Israel Binational Science Foundation (BSF) under Grant No.\ 2012/060, the ISF-NSFC joint research program under Grant No.\ 885/15, and a CAS President's International Fellowship Initiative (PIFI, China).
2b09ed780446c5d55dc4cdb36329d9db29b7e95f
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The inflationary universe scenario \cite{starobinsky,guth,linde} in which the early universe undergoes a rapid expansion has been generally accepted as a solution to the horizon problem and some other related problems of the standard big-bang cosmology. The origin of the field that drives inflation is still unknown and is subject to speculations. Among many models of inflation a popular class comprise tachyon inflation models \cite{fairbairn,feinstein,shiu1,kofman,sami,shiu2,cline,steer,campo,li,tachyon}. These models are of particular interest as in these models inflation is driven by the tachyon field originating in string theory. The tachyon potential is derived from string theory and has to satisfy some definite properties to describe tachyon condensation and other requirements in string theory. However, Kofman and Linde have shown \cite{kofman} that the slow-roll conditions are not compatible with a string coupling much smaller than one, and the compactification length scale much larger than the Planck length. This leads to the density fluctuations produced during inflation being incompatible with observational constraint on the amplitude of the scalar perturbations. This criticism is based on the string theory motivated values of the parameters in the tachyon potential, i.e., the brane tension and the parameters in the four-dimensional Newton constant obtained via conventional string compactification. Of course, if one relaxes the string theory constraints on the above mentioned parameters, the effective tachyon theory will naturally lead to a type of inflation which will slightly deviate from the conventional inflation based on the canonical scalar field theory. Steer and Vernizzi \cite{steer} have noted a deviation from the standard single field inflation in the second order consistency relations. Based on their analysis they concluded that the tachyon inflation could not be ruled out by the then available observations. It seems like the present observations \cite{planck2015} could perhaps discriminate between different tachyon models and disfavor or rule out some of these models (for a recent discussion on phenomenological constraints imposed by Planck 2015, see, e.g., ref \cite{pirtskhalava}). A simple tachyon model can be analyzed in the framework of the second Randall-Sundrum (RSII) model \cite{randall2}. The RSII model was originally proposed as a possible mechanism for localizing gravity on the 3+1 universe embedded in a 4+1 dimensional spacetime without compactification of the extra dimension. The model is a 4+1 dimensional Anti de Sitter (AdS$_5$) universe containing two 3-branes with opposite tensions separated in the fifth dimension: observers reside on the positive tension brane and the negative tension brane is pushed off to infinity. The Planck mass scale is determined by the curvature of the AdS spacetime rather then by the size of the fifth dimension. The fluctuation of the interbrane distance along the extra dimension implies the existence of the so called {\em radion} -- a massless scalar field that causes a distortion of the bulk geometry. In this regard, a stabilization mechanism of the interbrane distance has been proposed \cite{goldberger} by assuming the presence of scalar fields in the bulk. The stabilization mechanism is relevant for the RSI model where the interbrane distance is kept finite. In RSII model, as the negative tension brane is pushed off to infinity the radion disappears. However, it has been shown by Kim, Tupper, and Viollier \cite{kim} that a disappearance of the radion in RSII is an artifact of linear theory and hence, when going beyond linear theory the radion remains a dynamical field in the RSII model. Moreover, owing to the radion, the distance between branes remains finite in the RSII limit of infinite coordinate bulk even though the coordinate position of the second brane is infinite. The presence of the radion may have interesting physical implications. The radion field has been proposed as an inflaton \cite{barvinsky}. It has been shown \cite{bilic2013b,bilic2013a} that the interaction of the radion with the tachyon can alter the tachyon equation of state: by averaging over large scales the effective equation of state describes the warm dark matter. One of our aims here is to answer the question whether radion could drastically change the early cosmology in the framework of the RSII braneworld model. In this paper we propose a simple tachyon condensate as a model for inflation and we analyze two effects: the coupling of the tachyon with the radion and the modification of the standard cosmology in the RSII scenario. To study these effects we will consider an additional dynamical 3-brane moving in the AdS$_5$ background of the RSII model. The action of the 3+1 dimensional brane in the five dimensional bulk is equivalent to the Dirac-Born-Infeld (DBI) description of the Nambu-Goto 3-brane. \cite{bordemann,jackiw}. It is a simple matter to show that this additional 3-brane behaves effectively as a tachyon with the inverse quartic potential \cite{bilic2013a}. The tachyon model of this kind falls into the class of the power-law tachyon potentials $V(\theta) \propto \theta^{-n}$ with $n>2$ which drive a dark-matter attractor \cite{abramo}: for $\theta \rightarrow \infty$, the pressure tends to unity very quickly with the unpleasant feature of cold dark matter (CDM) domination at the end of inflation. Actually, this problem is imminent for all tachyon models with the ground state at $\theta\rightarrow\infty$ \cite{kofman}. The tachyon field rolls towards its ground state without oscillating about it and the conventional reheating mechanism does not work. Typically these scenarios suffer from a reheating problem, since gravitational particle production is not efficient compared to the standard non-gravitational particle production by an oscillating inflaton field. This will be discussed in more detail at the end of Sec.\ \ref{conditions}. The remainder of the paper is organized as follows. In Sec. \ref{Sec:Randal} we present the RSII model with back-reaction. The system of dynamical equations is presented in a dimensionless form suitable for the calculation of the expansion rate and slow-roll parameters. In Sec. \ref{conditions} the slow-roll approximation and initial conditions for our model are discussed in detail. Numerical results are presented and discussed in Sec. \ref{Sec:numres}. In the concluding section, Sec.\ \ref{conclusions}, we summarize our results and give conclusions. \section{Randall-Sundrum model with back-reaction}\label{Sec:Randal} The bulk spacetime of the extended RSII model which includes the back-reaction of the radion \cite{kim} in Fefferman-Graham coordinates is described by the line element \begin{equation} ds^2_{(5)}=G_{ab} dX^a dX^b=\frac{1}{k^2 z^2}\left[\left(1+ k^2 z^2\eta(x)\right)g^{\mu\nu}dx^\mu dx^\nu -\frac{1}{\left(1+ k^2 z^2\eta(x)\right)^2} dz^2\right] , \label{eq2001} \end{equation} where $k=1/\ell$ is the inverse of the AdS curvature radius $\ell$ and $\eta (x)$ is the radion field. The observer brane is placed at $z=\ell$ and $g_{\mu\nu}$ is the metric on the brane. Consider a dynamical 3-brane moving in the bulk. The brane Lagrangian describes the dynamic of the tachyon field modified by the interaction with the radion. After integrating out the fifth coordinate the total effective action is given by \cite{kim} \begin{equation} S = \int d^4x \sqrt{-g} \left( -\frac{R}{16\pi G}+\frac12 g^{\mu\nu}\Phi_{,\mu}\Phi_{,\nu} \right)+ S_{\rm br} , \label{eq3015} \end{equation} where $\Phi$ is the canonically normalized radion field related to $\eta$ as \begin{equation} \eta = \sinh^2 \left( \sqrt{\frac{4\pi G}{3}} \Phi \right). \end{equation} The brane action $S_{\rm br}$ is derived in Ref.\ \cite{bilic2013a} in terms of the induced metric or the ``pull back" of the bulk space-time metric $G_{ab}$ to the brane, \begin{equation} g^{(\rm ind)}_{\mu\nu}=G_{ab} \frac{\partial X^a}{\partial x^\mu} \frac{\partial X^b}{\partial x^\nu} \, . \label{eq0002} \end{equation} The action takes the form \begin{equation} S_{\rm br} = -\sigma \int d^4x \sqrt{- \det g^{\rm ind}_{\mu \nu}} =- \int d^{4}x \sqrt{-g} \, \frac{\sigma}{k^4\Theta^4}(1+k^2\Theta^2 \eta)^2 \sqrt{1-\frac{g^{\mu\nu}\Theta_{,\mu}\Theta_{,\nu}}{(1+k^2\Theta^2\eta)^3}}, \label{eq2006} \end{equation} where $\sigma$ denotes the brane tension and $\Theta$ is the tachyon field. In the absence of the radion ($\Phi = 0$) the brane action is just the tachyon condensate with the inverse quartic potential \begin{equation} S_{\rm br}^{(0)} = - \int d^{4}x \sqrt{-g} \frac{\lambda}{\Theta^4} \sqrt{1-g^{\mu\nu}\Theta_{,\mu}\Theta_{,\nu}}, \label{tachS} \end{equation} where \begin{equation} \lambda = \frac{\sigma}{k^4}. \end{equation} The combined brane-radion Lagrangian reads \begin{equation} {\cal{L}} = \frac12 g^{\mu\nu}\Phi_{,\mu}\Phi_{,\nu} -\frac{\lambda\psi^2}{\Theta^4}\sqrt{1-\frac{g^{\mu\nu}\Theta_{,\mu}\Theta_{,\nu}}{\psi^3}}, \label{lagrangian} \end{equation} where \begin{equation} \psi=1+k^2 \Theta^2 \eta . \end{equation} Hence, in our model, the tachyon is as usual the Dirac-Born-Infeld type scalar field whereas the radion is a canonical scalar. In the following we will assume the spatially flat FRW spacetime on the observer brane with four dimensional line element in the standard form \begin{equation} ds^2=g_{\mu\nu}dx^\mu dx^\nu=dt^2-a^2(t)(dr^2+r^2 d\Omega^2) \label{eq0012} \end{equation} The treatment of our system in a cosmological context is conveniently performed in the Hamiltonian formalism. For this purpose we first define the conjugate momentum fields as \begin{equation} \Pi_\Phi^\mu= \frac{\partial{\cal{L}}}{\partial\Phi_{,\mu}}, \label{eq2115} \quad \Pi_\Theta^\mu= \frac{\partial{\cal{L}}}{\partial\Theta_{,\mu}}. \end{equation} In the cosmological context $\Pi_\Phi^\mu$ and $\Pi_\theta^\mu$ are time-like so we may also define their magnitudes as \begin{equation} \Pi_\Phi=\sqrt{g_{\mu\nu}\Pi_\Phi^\mu\Pi_\Phi^\nu} , \hspace{1cm} \Pi_\theta=\sqrt{g_{\mu\nu}\Pi_\theta^\mu\Pi_\theta^\nu}. \label{eq2118} \end{equation} The Hamiltonian density may be derived from the stress tensor corresponding to the Lagrangian (\ref{lagrangian}) or by the Legendre transformation. Either way one finds \cite{bilic2013a} \begin{equation} {\cal{H}} = \frac{1}{2}\Pi_{\Phi}^2+\frac{\lambda\psi^2}{\Theta^4}\sqrt{1+\Pi_{\Theta}^2\Theta^8/(\lambda^2\psi)}. \label{totH} \end{equation} For later use we may also need the Hamiltonian density in terms of $\Phi_{,\mu}$ and $\Theta_{,\mu}$, \begin{equation} {\cal{H}} = \frac12 g^{\mu\nu}\Phi_{,\mu}\Phi_{,\nu} +\frac{\lambda\psi^2}{\Theta^4}\left(1-\frac{g^{\mu\nu}\Theta_{,\mu}\Theta_{,\nu}}{\psi^3}\right)^{-1/2}. \label{eq014} \end{equation} Next, we can write Hamilton's equations in the form \begin{eqnarray} \dot{\Phi} = \frac{\partial{\cal{H}}}{\partial\Pi_\Phi},\label{eqHam1}\\ \dot{\Theta} = \frac{\partial{\cal{H}}}{\partial\Pi_\Theta},\label{eqHam2} \\ \dot{\Pi}_\Phi +3H\Pi_\Phi=-\frac{\partial{\cal{H}}}{\partial\Phi},\label{eqHam3}\\ \dot{\Pi}_\Theta + 3H\Pi_\Theta=-\frac{\partial{\cal{H}}}{\partial\Theta}. \label{eqHam4} \end{eqnarray} In the spatially flat Randall-Sundrum cosmology the Hubble expansion rate $H$ is related to the Hamiltonian via the modified Friedmann equation \cite{maartens} \begin{eqnarray} H\equiv\frac{\dot{a}}{a}=\sqrt{\frac{8 \pi G}{3} \mathcal{H}\left(1+ \frac{2 \pi G}{3k^2} \mathcal{H}\right) }. \label{scale_a} \end{eqnarray} In addition, we will make use of the energy-momentum conservation equation \begin{equation} \dot{\mathcal{H}}+3H(\mathcal{H}+\mathcal{L})=0 , \label{3200} \end{equation} which, combined with the time derivative of (\ref{scale_a}), yields the second Friedmann equation in the form \begin{equation} \dot{H}=-4\pi G(\mathcal{H}+\mathcal{L})\left(1+ \frac{4 \pi G}{3k^2} \mathcal{H}\right). \label{eq3222} \end{equation} Thus, the Friedman equations are modified in the RSII cosmology. As far as we know the effects of these modifications on tachyon inflation were first studied by Bento, Bertolami and Sen \cite{bento}. To solve the system of equations (\ref{eqHam1})-(\ref{scale_a}), it is convenient to rescale the time as $t=\tau/k$ and express the system in terms of dimensionless quantities. Besides, we can eliminate the coupling constant $\lambda$ from the equations by appropriately rescaling the fields $\Phi$ and $\Theta$ and their conjugate fields $\Pi_{\Phi}$ and $\Pi_{\Theta}$. To this end we introduce the dimensionless functions \begin{eqnarray} h = H/k, \quad \phi = \Phi/(k \sqrt{\lambda}),\quad \pi_\phi = \Pi_\Phi/(k^2\sqrt{\lambda})),\qua \theta=k \Theta, \quad \pi_\theta = \Pi_{\Theta}/(k^4 \lambda), \label{eq002} \end{eqnarray} and rescale the Lagrangian and Hamiltonian to obtain the dimensionless pressure and energy density: \begin{equation} \bar{p}\equiv \frac{\mathcal{L}}{k^4\lambda} =\frac{1}{2}\pi_\phi^2-\frac{\psi^2}{\theta^4} \frac{1}{\sqrt{1+\theta^8\pi_\theta^2/\psi}} , \label{eq504} \end{equation} \begin{equation} \bar{\rho}\equiv \frac{\mathcal{H}}{k^4\lambda}=\frac{1}{2}\pi_\phi^2+\frac{\psi^2}{\theta^4} \sqrt{1+\theta^8\pi_\theta^2/\psi}. \label{eq505} \end{equation} Following Steer and Vernizzi \cite{steer} we also introduce a combined dimensionless coupling \begin{equation} \kappa^2=8\pi\lambda G k^2 . \label{eq102} \end{equation} Then, from (\ref{eqHam1})-(\ref{scale_a}) we obtain the following set of equations \begin{equation} \label{sysRT1} \dot \phi=\pi_\phi, \end{equation} \begin{equation} \label{sysRT2} \dot \theta=\frac{\theta^4\psi\pi_{\theta}} {\sqrt{1+\theta^8\pi_{\theta}^2/\psi}}, \end{equation} \begin{equation} \label{sysRT3} \dot \pi_\phi=-3h\pi_\phi -\frac{\psi}{2\theta^2}\,\frac{4+3\theta^8\pi_{\theta}^2/\psi} {\sqrt{1+\theta^8\pi_{\theta}^2/\psi}}\eta', \end{equation} \begin{equation} \label{sysRT4} \dot \pi_{\theta}=-3h{{\pi }_{\theta}} +\frac{\psi}{\theta^5}\,\frac{4-3\theta^{10}\eta\pi_{\theta}^2/\psi} {\sqrt{1+\theta^8\pi_{\theta}^2/\psi}}, \end{equation} where \begin{eqnarray} \label{h} h\equiv \frac{\dot{a}}{a}=\sqrt{\frac{\kappa^2}{3}\bar{\rho}\left(1+\frac{\kappa^2}{12}\bar{\rho} \right)}, \end{eqnarray} \begin{equation} \psi=1+\theta^2 \eta, \end{equation} \begin{equation} \eta = \sinh^2 \left( \sqrt{\frac{\kappa^2}{6}} \phi \right) , \end{equation} \begin{equation} \eta^{\prime}=\frac{d\eta}{d\phi} = \sqrt{\frac{\kappa^2}{6}} \sinh \left( \sqrt{\frac{2 \kappa^2}{3}} \phi \right). \label{eq012} \end{equation} In addition to Eqs.\ (\ref{sysRT1})-(\ref{h}) we can solve in parallel the second Friedman equation \begin{eqnarray} \dot{h}=-\frac{\kappa^2}{2}(\bar{\rho}+\bar{p})\left(1+\frac{\kappa^2}{6}\bar{\rho} \right), \label{h2} \end{eqnarray} and \begin{equation} \dot{N}=h \label{n} \end{equation} where $N$ is the number of e-folds as a function of $\tau$. In Eqs.\ (\ref{sysRT1})-(\ref{n}) and from now on the overdot denotes a derivative with respect to $\tau$. Obviously, the explicit dependence on $\lambda$ in Eqs.\ (\ref{sysRT1})-(\ref{eq012}) is eliminated and the only remaining free parameter is $\kappa$. Equations (\ref{sysRT1}) and (\ref{sysRT2}) can be used to express the pressure and energy density in terms of $\dot{\phi}$ and $\dot{\theta}$. One thus finds \begin{equation} \bar{p}=\frac{1}{2}\dot{\phi}^2-\frac{\psi^2}{\theta^4} \sqrt{1-\dot{\theta}^2/\psi^3} , \label{eq015} \end{equation} \begin{equation} \bar{\rho}=\frac{1}{2}\dot{\phi}^2+\frac{\psi^2}{\theta^4} \frac{1}{\sqrt{1-\dot{\theta}^2/\psi^3}}. \label{eq016} \end{equation} The functional dependence of $h$ on $\tau$ will be used to calculate the slow-roll parameters and the number of e-folds. The slow-roll parameters are defined as \cite{steer,schwarz} \begin{equation} \epsilon_{i} \equiv \frac{d\ln| \epsilon_{i-1}|}{Hdt},\qquad i \geq 1, \end{equation} where \begin{equation} \epsilon_0 \equiv \frac{H_*}{H}. \end{equation} and $H_*$ is the Hubble rate at an arbitrarily chosen time. Using the previously defined dimensionless Hubble rate $h$, the first two parameters can be written as \begin{equation}\label{eps1} \epsilon_1 = -\frac{\dot{h}}{h^2}, \end{equation} \begin{equation}\label{eps2} \epsilon_2 = 2\epsilon_1+\frac{\ddot{h}}{h\dot{h}} . \end{equation} The conditions for a slow-roll regime are satisfied when $\epsilon_1 < 1$ and $\epsilon_2 < 1$, and inflation ends when any of them exceeds unity. The effect of the radion and the tachyon can be seen if we compare the slow-roll parameters for the full model with those for the model with inverse quartic tachyon potential. The number of e-folds is defined through Eq.\ (\ref{n}) as \begin{equation} N=\int_{\tau_{\rm i}}^{\tau_{\rm f}}d\tau h, \end{equation} where the subscripts ${\rm i}$ and ${\rm f}$ denote the beginning and the end of inflation, respectively. The slow-roll parameters are related to observable quantities, in particular to the tensor-to-scalar ratio $r$ and the scalar spectral index $n_{\rm s}$ defined by \begin{equation} r=\frac{\mathcal{P}_{\rm T}}{\mathcal{P}_{{\rm S}}}, \label{eq3005} \end{equation} \begin{equation} n_{\rm s}= \frac{d\ln \mathcal{P}_{{\rm S}}}{d\ln q}, \label{eq3006} \end{equation} where $\mathcal{P}_{{\rm S}}$ and $\mathcal{P}_{\rm T}$ are the power spectra of scalar and tensor perturbations, respectively, evaluated at the horizon, i.e., for a wave-number satisfying $q=aH$. Calculation of the spectra proceeds by identifying the proper canonical field and imposing quantization of the quadratic action for the near free field. The procedure for a general k-inflation is described in \cite{garriga} and applied to the tachyon fluid in Refs. \cite{steer,frolov}. It turns out that at the lowest order in $\epsilon_1$ and $\epsilon_2$, the power spectra may be expressed in the same way as in the standard tachyon inflation \cite{steer}: \begin{equation} \mathcal{P}_{\rm T} \simeq [1-2(1+C)\epsilon_1]\frac{16GH^2}{\pi}, \label{eq3007} \end{equation} \begin{equation} \mathcal{P}_{{\rm S}} \simeq [1-2(1+C-\alpha)\epsilon_1-C \epsilon_2]\frac{GH^2}{\pi\epsilon_1 }, \label{eq3008} \end{equation} where $C=-2+\ln 2 +\gamma\simeq -0.72$, and $\alpha$ is a parameter related to the speed of sound expanded in $\epsilon_1$: \begin{equation} c_{\rm s}=1-2\alpha\epsilon_1 +O(\epsilon_1^2). \label{eq3009} \end{equation} The parameter $\alpha$ will be calculated in the next section using the relation between $\epsilon_1$ and the speed of sound in the slow-roll approximation. \section{Conditions for tachyon inflation} \label{conditions} \subsection{Inverse quartic potential} To solve the system of equations (\ref{sysRT1})-(\ref{sysRT4}) we need to choose initial conditions relevant for inflation. To this end we first solve a simpler case in which the radion is absent. The model is described by the pure tachyon Lagrangian with the inverse quartic potential \cite{bpu9} \begin{equation} {\cal{L}} = -\frac{\lambda}{\Theta^4}\sqrt{1-g^{\mu\nu}\Theta_{,\mu}\Theta_{,\nu}} \,, \label{eq000} \end{equation} with the corresponding Hamiltonian \begin{equation} {\cal{H}} =\frac{\lambda}{\Theta^4}\sqrt{1+\Pi_{\Theta}^2\Theta^8/(\lambda^2)}. \label{eq001} \end{equation} As before, we rescale the time as $t=\tau/k$, absorb the coupling constant $\lambda$ into the conjugate field $\Pi_{\Theta}$ and introduce the dimensionless functions as in (\ref{eq002}). Then, using (\ref{eqHam2}), (\ref{eqHam4}) and (\ref{eq001}) we obtain the Hamilton equations in the form \begin{equation} \dot \theta=\frac{\theta^4\pi_{\theta}} {\sqrt{1+\theta^8\pi_{\theta}^2}} \label{eq003} \end{equation} \begin{equation} \dot \pi_\theta=-3h\pi_\theta +\frac{4}{\theta^5 \sqrt{1+\theta^8\pi_\theta^2}} \label{eq004} \end{equation} The dimensionless pressure and energy density are given by \begin{equation} \bar{p}= -\frac{1}{\theta^4\sqrt{1+\theta^8\pi_{\theta}^2}}= -\frac{1}{\theta^4}\sqrt{1-\dot{\theta}^2}, \label{eq008} \end{equation} \begin{equation} \bar{\rho}= \frac{1}{\theta^4}\sqrt{1+\theta^8\pi_{\theta}^2}= \frac{1}{\theta^4}\frac{1}{\sqrt{1-\dot{\theta}^2}}, \label{eq2008} \end{equation} Using this we also find a simple expression for the sound speed \begin{equation} c_{\rm s}^2\equiv \left.\frac{\partial p}{\partial\rho}\right|_{\theta} =\frac{1}{1+\theta^8\pi_{\theta}^2}= 1-\dot{\theta}^2. \label{eq2009} \end{equation} The Friedmann equations are given by Eq.\ (\ref{h}) and (\ref{h2}) with (\ref{eq008}) and (\ref{eq2008}). Hence, as before, the coupling $\lambda$ drops out and the only remaining parameter is $\kappa$ defined in (\ref{eq102}). The system of equations (\ref{eq003})-(\ref{eq004}) can be solved numerically for a chosen set of initial conditions. Once the solution for $H$ is found the slow-roll parameters can be easily calculated. \subsection{Slow-roll approximation and initial conditions} \label{initial} To find appropriate initial value of the field $\theta$, we first consider the pure tachyon model in the slow-roll approximation. Tachyon inflation is based upon the slow evolution of $\theta$ with the slow-roll conditions \cite{steer} \begin{equation} \dot{\theta}\simeq \theta^4\pi_\theta \ll 1, \quad \dot\pi_\theta \ll 3h \pi_\theta, \end{equation} so that in the slow-roll approximation we may neglect the factors $(1-\dot{\theta}^2)^{1/2}$. Then, during inflation we have \begin{equation} h\simeq \frac{\kappa}{\sqrt3 \theta^2}\left(1+\frac{\kappa^2}{12\theta^4}\right)^{1/2}, \label{eq1007} \end{equation} \begin{equation} \dot{\theta} \simeq \frac{4}{3h\theta} \simeq \frac{4\theta }{\sqrt3 \kappa} \left(1+\frac{\kappa^2}{12\theta^4}\right)^{-1/2} , \label{eq1008} \end{equation} \begin{equation} \ddot{\theta} \simeq \frac{4\dot{\theta} }{\sqrt3 \kappa} \left(1+\frac{\kappa^2}{12\theta^4}\right)^{-3/2} \left(1+\frac{\kappa^2}{4\theta^4}\right), \label{eq1009} \end{equation} and using (\ref{h2}) we also find \begin{equation} \dot{h}=-\frac{\kappa^2 \dot{\theta}^2}{2\theta^4}\left(1+\frac{\kappa^2}{6\theta^4}\right). \end{equation} As a consequence, the slow-roll parameters (\ref{eps1}) and (\ref{eps2}) can be approximated by \begin{eqnarray} \epsilon_1&\simeq&\frac32 \dot{\theta}^2\left(1+\frac{\kappa^2}{6\theta^4}\right) \left(1+\frac{\kappa^2}{12\theta^4}\right)^{-1} \nonumber \\ &\simeq&\frac{8\theta^2}{\kappa^2}\left(1+\frac{\kappa^2}{6\theta^4}\right) \left(1+\frac{\kappa^2}{12\theta^4}\right)^{-2}, \label{eq009} \end{eqnarray} \begin{eqnarray} \epsilon_2 &\simeq& 2\frac{\ddot{\theta}}{h\dot{\theta}} -\dot{\theta}^2\frac{\kappa^2}{4\theta^4} \left(1+\frac{\kappa^2}{6\theta^4}\right)^{-1}\left(1+\frac{\kappa^2}{12\theta^4}\right)^{-1} \nonumber \\ &\simeq& \frac{8\theta^2}{\kappa^2} \left(1+\frac{\kappa^2}{12\theta^4}\right)^{-2}\left[ 1+\frac{\kappa^2}{4\theta^4}-\frac{\kappa^2}{6\theta^4}\left(1+\frac{\kappa^2}{6\theta^4}\right)^{-1} \right]. \label{eq010} \end{eqnarray} In the slow roll regime we have $\kappa^2/\theta^4\gg 1$, so the corrections due to the RSII modification in Eqs.\ (\ref{eq1007})-(\ref{eq010}) will dominate over unity and we find \begin{equation} h\simeq \frac16 \frac{\kappa^2}{\theta^4}, \quad\quad \dot{\theta}\simeq 8 \frac{\theta^3 }{ \kappa^2} , \quad\quad \ddot{\theta}\simeq 24 \frac{\theta^2}{\kappa^2}\dot{\theta}, \label{eq021} \end{equation} \begin{equation} \epsilon_1\simeq 3 \dot{\theta}^2\simeq 192 \frac{\theta^6}{\kappa^4}, \quad\quad \epsilon_2\simeq 288 \frac{\theta^6}{\kappa^4} \simeq \frac32\epsilon_1 . \label{eq018} \end{equation} In contrast, if we disregarded the RSII corrections we would obtain the usual slow-roll equations of tachyon inflation \cite{steer,campo,li,garousi} for the potential $V=\lambda/ \theta^{4}$ \begin{equation} h\simeq \frac{1}{\sqrt{3}} \frac{\kappa}{\theta^2}, \quad\quad \dot{\theta}\simeq \frac{4}{\sqrt{3}} \frac{\theta}{ \kappa}, \quad\quad \ddot{\theta}\simeq \frac{4}{\sqrt{3}} \frac{\dot{\theta}}{\kappa}, \label{eq0021} \end{equation} \begin{equation} \epsilon_1\simeq \frac32 \dot{\theta}^2\simeq 8 \frac{\theta^2}{\kappa^2}, \quad\quad \epsilon_2\simeq \epsilon_1. \label{eq0018} \end{equation} Hence, in the slow-roll regime the tachyon inflation in the RSII modified cosmology proceeds in a quite distinct way compared with that in the standard FRW cosmology. However, close to and at the end of inflation we have $\kappa^2/\theta_{\rm f}^4\ll 1$ and we can neglect the RSII cosmology corrections. Hence, the expressions (\ref{eq0021}) and (\ref{eq0018}) can be used at the end of inflation where we find \begin{equation} \epsilon_1(\theta_{\rm f})\simeq \epsilon_2(\theta_{\rm f})\simeq \frac{8\theta_{\rm f}^2}{\kappa^2} \simeq 1, \label{eq017} \end{equation} and \begin{equation} h(\theta_{\rm f})\simeq \frac{8}{\sqrt3 \kappa}. \label{eq2007} \end{equation} In the slow-roll approximation the number of e-folds is given by \begin{eqnarray}\label{inteq} N \simeq\frac{\kappa^2}{4} \int_{\theta_0}^{\theta_{\rm{f}}} \frac{d\theta}{\theta^3} \left(1+\frac{\kappa^2}{12\theta^4}\right) \simeq \frac{\kappa^2}{8\theta_0^2}\left(1+\frac{\kappa^2}{36\theta_0^4}\right)-1 \simeq \frac23 \frac{1}{\epsilon_1(\theta_0)}-1, \end{eqnarray} where we have exploited $\kappa^2/(36\theta_0^4)\gg 1 $ at the beginning and $\kappa^2/(36\theta_{\rm f}^4)\ll 1 $ at the end of inflation together with the condition (\ref{eq017}). For comparison, in the standard tachyon inflation described by (\ref{eq0021}) and (\ref{eq0018}) we would obtain \begin{equation} N_{\rm st.tach} \simeq\frac{\kappa^2}{8\theta_0^2}-1 \simeq \frac{1}{\epsilon_1(\theta_0)}-1. \label{eq319} \end{equation} For example, the choice $\kappa^2=5$ and $\theta_0 = 0.25$ leads to $N_{\rm st.tach} \simeq 9 $ whereas in RSII cosmology with the same parameters one finds $N \simeq 365$. At this point it is convenient to estimate a phenomenologically acceptable range of the couplings $\lambda$ and $\kappa$. Although the evolution equations do not depend on $\lambda$, its approximate value is needed for choosing appropriate initial conditions for the radion field. The value of $\lambda$ may be estimated using the observational constraint on the amplitude of scalar perturbations. The approximate expression \begin{equation} \mathcal{P}_{\rm S}\simeq \frac{GH^2}{\pi\epsilon_1}, \end{equation} which follows from (\ref{eq3008}), is to be compared with the power spectrum amplitude $A_s$ measured by Planck \cite{planck2015}: \begin{equation} A_s\simeq 2.2\times 10^{-9}. \end{equation} Hence, we must make sure that the condition \begin{equation} \frac{H}{M_{\rm P}} \lesssim \sqrt{\pi A_s}\simeq 8.31\times 10^{-5} \end{equation} is satisfied close to and at the end of inflation (where $\epsilon_1 \lesssim 1$). (See also Ref.\ \cite{thomas}.) Here we define the Planck mass as $M_{\rm P}=G^{-1/2}$. According to Eq.\ (\ref{eq2007}) in the slow-roll approximation near the end of inflation we have \begin{equation} \frac{H}{M_{\rm P}}\simeq \frac{8}{\sqrt{3}} \frac{k}{\kappa M_{\rm P}}= \sqrt{\frac{8 }{3\pi}}\frac{k^2}{\sqrt{\sigma}} , \label{eq03} \end{equation} yielding \begin{equation} \frac{k}{\sigma^{1/4}}\lesssim 10^{-2}, \label{eq01} \end{equation} or \begin{equation} \lambda\gtrsim 10^{8}. \label{eq02} \end{equation} To estimate $\kappa$ note first that the tension of a D$p$-brane is given by \cite{johnson} \begin{equation} \sigma =\frac{1}{(2\pi)^p \alpha^{\prime (p+1)/2} g_{\rm s}}, \end{equation} where $g_{\rm s}$ is the string coupling constant and $1/(2\pi\alpha^{\prime})$ is the string tension. Using this for $p=3$ from (\ref{eq01}) we find a constraint \begin{equation} g_{\rm s}\lesssim 4\times 10^{-11}\left(\frac{M_{\rm s}}{k}\right)^4 , \end{equation} where $M_{\rm s}=1/\sqrt{\alpha^{\prime}}$. Using this constraint we can choose $k$ and $M_{\rm s}$ such that the scale hierarchy \begin{equation} H < k < M_{\rm s} < M_{\rm P} \end{equation} is satisfied. With this in mind we can give an order of magnitude estimate for an acceptable range of values of our free parameter $\kappa$. Requiring $k\gtrsim H$, from (\ref{eq03}) it follows \begin{equation} \kappa\gtrsim 8/\sqrt{3}, \end{equation} so we can safely choose $\kappa$ to vary in the range $1<\kappa<20$. Equations (\ref{eq018}) and (\ref{inteq}) can be used to estimate the value of the ratio $\kappa^2/\theta_0^3$ by fitting the model to the observational parameters $r$ and $n_{\rm s}$ defined by (\ref{eq3005}) and (\ref{eq3006}). First, using (\ref{eq2009}) and (\ref{eq018}) we can express the sound speed in terms of $\epsilon_1$: \begin{equation} c_s \simeq1-\frac16 \epsilon_1, \end{equation} and comparing this with (\ref{eq3009}) we find $\alpha=1/12$. In contrast, in the standard tachyon inflation the value $\alpha=1/6$ is obtained \cite{steer}. Using (\ref{eq3007}) and (\ref{eq3008}) we find up to the second order \begin{equation} r=16 \epsilon_1\left[1- \frac16 \epsilon_1+C\epsilon_2 \right], \label{eq005} \end{equation} \begin{equation} n_{\rm s}=1-2 \epsilon_1 - \epsilon_2 -\left[2 \epsilon_1^2+\left(2C +\frac{17}{6} \right)\epsilon_1\epsilon_2 +C\epsilon_2\epsilon_3\right], \label{eq006} \end{equation} where it is understood that $\epsilon_1$ and $\epsilon_2$ take their values at or close to the beginning of inflation. Using these equations at linear order with $\epsilon_2\simeq 3\epsilon_1/2$ and (\ref{inteq}), we find the approximate relations \begin{equation} r=\frac{32}{3}\frac{1}{N+1} , \label{eq501} \end{equation} \begin{equation} n_{\rm s}=1-\frac73 \frac{1}{N+1}, \label{eq502} \end{equation} \begin{equation} r=\frac{32}{7}(1-n_{\rm s}). \label{eq503} \end{equation} Had we used the usual relation $\epsilon_2\simeq \epsilon_1$ and (\ref{eq319}) we would have obtained the standard tachyon-inflation relations \cite{steer,li} \begin{equation} r=\frac{16}{N+1}, \end{equation} \begin{equation} n_{\rm s}=1-\frac{3}{N+1}, \end{equation} \begin{equation} r=\frac{16}{3} (1-n_{\rm s}). \label{eq5004} \end{equation} Comparing $r$ and $n_{\rm s}$ with the latest observations we can fix the parameters $\epsilon_1$, $\epsilon_2$, and $N$. Then, from (\ref{inteq}) we can determine the ratio $\kappa^2/\theta_0^3$ and for a chosen set of values of $\kappa$ we find the corresponding $\theta_0$. These values of $\kappa$ and the corresponding initial values $\theta_0$ as initial conditions are then used to solve the system of equation (\ref{eq003})-(\ref{eq004}) for the pure tachyon and (\ref{sysRT1})-(\ref{sysRT4}) for the tachyon-radion system. There is no a priori reason to restrict possible initial values of the radion field $\Phi$ so we can choose a range of initial values based on the natural scale dictated by observations, i.e., the scale between $H$ and $M_{\rm P}$. Hence, a natural initial value for $\Phi$ would be of the order of $k$ or a few orders larger, say in the range 10 to 1000 $k$. However, according to (\ref{eq002}) the dimensionless radion field $\phi$ is rescaled with respect to $\Phi/k$ by a factor of $1/\sqrt{\lambda}\simeq 10^{-4}$ so we can choose the initial value $\phi_0=\phi(0)$ in the range from 0.001 to 0.5. As we have mentioned in Introduction our tachyon model suffers from the so called reheating problem \cite{kofman}. A possible way out is provided by string theory. String theory D-branes couple to the (pull-back of) antisymmetric tensor field $\mathcal{B}$ that combines the Kalb-Ramond and electromagnetic fields. In this way there exist a natural coupling between the tachyon and the electromagnetic field. Therefore, this interaction could serve as a possible reheating mechanism at the end of inflation. In a tachyon model based on brane-antibrane annihilation resulting in a time dependent tachyon condensate it has been shown \cite{cline} that a coupling of massless fields to the time dependent tachyon condensate could yield a reheating efficient enough to overcome the above mention problem of a CDM dominance. Another possible way out of the reheating problem could be the so called {\em warm inflation} \cite{berera}. Warm inflation is an alternative inflation scenario with no need for a reheating period. In warm inflation, dissipative effects are included during inflation, so that radiation is produced in parallel with the inflationary expansion and inflation ends when the universe heats up to become radiation dominated. This scenario has been successfully applied to tachyon inflation models \cite{herrera,motaharfar} and, in principle, should also work for our model. This requires further investigation which goes beyond the scope of the present paper. \section{Numerical results}\label{Sec:numres} The system of equations (\ref{sysRT1})-(\ref{sysRT4}) is evolved numerically starting from $\tau=0$ up to some large $\tau$ of the order of 100. The initial values $\theta_0$ and $\phi_0$ are chosen as described in the previous section and the initial conjugate momenta are taken to be $\pi_{\theta 0}=\pi_{\phi 0}=0$. The function $N(\tau)$ is solved simultaneously using (\ref{n}) with $N(0)=0$. The time evolution of the slow roll parameters $\epsilon_1$ and $\epsilon_2$ are obtained using (\ref{eps1}) and (\ref{eps2}). The inflation ends at a point $\tau_{\rm f}$ at which $\epsilon_1(\tau_{\rm f})=1$. The beginning of inflation at $\tau_i$ is then found by requiring $N(\tau_{\rm f})-N(\tau_{\rm i})=N$. The results for $\kappa=2$, $\theta_0=0.25$ and $\phi_0=0.4$ are presented in Fig.\ \ref{fig1} together with the results calculated for the system (\ref{eq003})-(\ref{eq004}) within the standard FRW cosmology with the same initial $\theta_0=0.25$ and $\pi_{\theta 0}=0$. \begin{figure}[!ht] \includegraphics[width=0.45\textwidth,trim= 0 0cm 0 0cm]{fig1-comp.eps} \hspace{0.02\textwidth} \includegraphics[width=0.45\textwidth]{fig2-comp.eps} \\ \vspace{0.02\textwidth} \includegraphics[width=0.45\textwidth]{fig4-comp.eps} \hspace{0.02\textwidth} \includegraphics[width=0.45\textwidth]{fig5-comp.eps} \caption{Time evolution of the tachyon $\theta$ (top left), the radion $\phi$ (top right) fields and of the slow-roll parameters $\epsilon_1$ (bottom left) and $\epsilon_2$ (bottom right) for $\kappa=2$, $\phi_0=0.4$, $\pi_{\theta 0}=\pi_{\phi 0}=0$, $\theta_0=0.25$, in the tachyon-inflation model with inverse quartic potential in the RSII cosmology with radion (full black line) and without radion (dotted red line). The dashed blue line represents the corresponding results for the tachyon-inflation model with inverse quartic potential in the standard FRW cosmology and no radion.} \label{fig1} \end{figure} To calculate the quantities $n_s$ and $r$ we proceed as follows. For a chosen pair of $(N, \kappa)$ we first find the initial value $\theta_0$ from Eq.\ (\ref{inteq}). Then, for a chosen set of initial values $\phi_0$, $\pi_{\theta 0}$, $\pi_{\phi 0}$ (at $\tau=0$) we find the corresponding $\phi_{\rm i}$, $\pi_{\theta \rm i}$, $\pi_{\phi \rm i}$ at $\tau=\tau_{\rm i}$ and calculate $h$, $\dot{h}$ and $\ddot{h}$ using equations (\ref{sysRT1})-(\ref{h}) with (\ref{eq504}) and (\ref{eq505}). From this we find $\epsilon_1(\theta_{\rm{i}})$ and $\epsilon_2(\theta_{\rm{i}})$ using the defining expressions (\ref{eps1}) and (\ref{eps2}) and calculate the parameters $r$ and $n_{\rm s}$ from (\ref{eq005}) and (\ref{eq006}), respectively. In Fig.\ \ref{fig2} we present $n_s$ and $r$ as functions of $\kappa$ and $N$ for a fixed $\phi_0=0.01$. \begin{figure}[ht!] \begin{center} \includegraphics[width=0.45\textwidth,trim= 0 0cm 0 0cm]{plot3d-r.eps} \hspace{0.02\textwidth} \includegraphics[width=0.45\textwidth]{plot3d-n.eps} \caption{The quantities $r$ (left) and $n_{\rm s}$ (right) as functions of the parameters $N$ and $\kappa$ as indicated on the horizontal axes for initial $\phi_0=0.01$. The initial $\theta_0$ varies with $N$ and $\kappa$ according to Eq.\ (\ref{inteq}).} \label{fig2} \end{center} \end{figure} In Fig.\ \ref{fig3} we present the $n_s-r$ diagram with 10000 points superimposed on the observational constraints taken from the Planck Collaboration 2015 \cite{planck2015}. Each point in the diagram corresponds to a set $(N, \kappa, \phi_0)$ chosen randomly in the range $60 \leq N \leq 120$, $1\leq \kappa \leq 12$ and $0\leq \phi_0 \leq 0.5$. \begin{figure}[ht!] \centering \includegraphics[width=0.5\textwidth,trim= 0 0cm 0 0cm]{fig6.eps} \caption{$r$ versus $n_{\rm s}$ diagram with observational constraints from Ref.\ \cite{planck2015}. The dots represent the calculation in the tachyon-radion model for various $N$, $\kappa$ and $\phi_0$ chosen randomly in the range $60 \leq N \leq 120$, $1\leq \kappa \leq 12$ and $0\leq \phi_0 \leq 0.5$. The full line represents the slow-roll approximation (Eq.\ (\ref{eq503})) of the RSII model with no radion. The dashed line represents the slow-roll approximation (Eq.\ (\ref{eq5004})) of the standard tachyon model with inverse quartic potential.} \label{fig3} \end{figure} To obtain a more favorable distribution of points within the 2$\sigma$ area measured by the Planck collaboration we have used the distribution histograms of the number of e-folds $N$ and the parameters $\kappa$ and $\phi_0$. In this way we have been guided to restrict $N$ to range between 85 and 110, $\kappa$ between 1 and 8, and $\phi_0$ between 0 and 0.5. The outcome of these constraints is presented in Fig.\ \ref{fig4}. \begin{figure}[ht!] \centering \includegraphics[width=0.5\textwidth,height=0.5\textwidth]{fig6a.eps} \caption{Same as in Fig.\ \ref{fig3} for $N$, $\kappa$, and $\phi_0$ chosen randomly in the intervals $85 \leq N \leq 110$, $1\leq \kappa \leq 8$ and $0\leq \phi_0 \leq 0.5$. The inner and outer dash-dotted lines denote the Planck contours of the one and two $\sigma$ constraints, respectively. } \label{fig4} \end{figure} In this figure one notices basically two bands: the dominant one almost parallel to the $r$ axis with most data points concentrated close to the line corresponding to the RSII model prediction without radion and the less dens one almost parallel to the $n_s$ axis, with $r$ between 0.1 and 0.13. From a quite large data set of numerical results we find that smaller values of $\kappa$ give a better agreement with observational data from the Planck mission. For example, if we fix $\kappa=2$ and vary $N$ in a wider range $N$ $60\leq N \leq 120$) and $\phi_0$ in a bit narrower range $0 \leq\phi_0 \leq 0.25$) we obtain an interesting splitting of points into three disconnected clusters (Fig.\ \ref{fig5}). This splitting is a clear manifestation of the nonlinearity of Eqs.\ (\ref{sysRT1})-(\ref{sysRT4}). \begin{figure}[ht!] \centering \includegraphics[width=0.5\textwidth,height=0.5\textwidth]{fig6c.eps} \caption{Same as in Fig.\ \ref{fig4} for a fixed $\kappa=2$ and $N$ and $\phi_0$ chosen randomly in the intervals $60 \leq N \leq 120$ and $0\leq \phi_0 \leq 0.25$.} \label{fig5} \end{figure} The most pronounced cluster of numerical points fits in quite well within the 2$\sigma$ regions of constraints given by the Planck collaboration. A still better agreement between our model and the observational data constraints is obtained for a very narrow range of parameters as shown in Figs.\ \ref{fig6} and \ref{fig7}. \begin{figure}[ht!] \centering \includegraphics[width=0.5\textwidth,height=0.5\textwidth]{fig6d.eps} \caption{Same as in Fig.\ \ref{fig5} for $N$ and $\phi_0$ chosen randomly in the intervals $110 \leq N \leq 120$ and $0.1\leq \phi_0 \leq 0.2$.} \label{fig6} \end{figure} \begin{figure}[ht!] \centering \includegraphics[width=0.5\textwidth,height=0.5\textwidth]{fig6e.eps} \caption{Same as in Fig.\ \ref{fig3} for fixed $\kappa=1.25$ and $\phi_0=0.05$ and $N$ chosen randomly in the interval $115 \leq N \leq 120$.} \label{fig7} \end{figure} The best fit is obtained for $\kappa=1.25$, $\phi_0=0.05$ and $115 \leq N \leq 120$. \section{Summary and conclusions} \label{conclusions} We have investigated a model of inflation based on the dynamics of a D3-brane in the AdS$_5$ bulk of the RSII model. The bulk metric is extended to include the back reaction of the radion excitations. We have shown that the slow-roll equations of the tachyon inflation are quite distinct to those of the standard tachyon inflation with the same potential. In particular, the departure of the sound speed from unity equals $c_{\rm s}-1\simeq -\epsilon_1/6$ in contrast to the standard result $c_{\rm s}-1\simeq -\epsilon_1/3$. The $n_{\rm s}-r$ relation in our model is substantially different from the standard one and is closer to the best observational value, as shown in Fig.\ \ref{fig6}. Note that the largest concentration of numerical results, represented by the points in the plot in Fig.\ \ref{fig6}, are within 2$\sigma$ of the {\em Planck} TT+lowP and {\em Planck} TT+lowP+lensing\-+BAO+JLA+H$_0$ (red and light blue shaded regions, respectively). Clearly, the agreement with observations is not ideal and it is fair to say that the present model is disfavored but not excluded. However, one should bare in mind that the model is based on the brane dynamics which results in a definite potential with one free parameter only. In contrast, the majority of other tachyon potentials discussed in the literature are chosen arbitrarily and adjusted so that the results agree with observations. In this work we have analyzed the simplest tachyon model that stems from the dynamics of a D3-brane in an AdS$_5$ bulk yielding basically an inverse quartic potential. In principle, the same mechanism could lead to a more general tachyon potential if the AdS$_5$ background metric is deformed by the presence of matter in the bulk, e.g., in the form of a minimally coupled scalar field with an arbitrary self-interaction potential. This will be the subject of our investigation in the future. \section*{Acknowledgments} This work has been supported by ICTP - SEENET-MTP project PRJ-09 Cosmology and Strings. The work of N.\ Bili\'c has been supported by the Croatian Science Foundation under the project (IP-2014-09-9582). D.\ Dimitrijevic, G.\ Djordjevic and M.\ Milosevic acknowledge support provided by the Serbian Ministry for Education, Science and Technological Development under the projects No 176021 and No 174020. G.\ Djordjevic would like to thank CERN-Theoretical Physics Department for kind hospitality during the finalization of this paper.
27ec94dfda5ba095429a284c3719a2b214bab085
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction}\label{sec1} Optimization problems are prevalent and have held great importance throughout history in engineering applications and scientific endeavors. For instance, many problems in the field of artificial intelligence (AI) can be viewed as optimization problems. Accordingly, generic \emph{local} optimization methods, such as hill climbing and the gradient method, have been successfully adopted to solve AI problems since early research on the topic \cite{kir70,gul94,dei11}. On the other hand, the application of \emph{global} optimization to AI problems has been studied much less despite its practical importance. This is mainly due to the lack of necessary computational power in the past and the absence of a practical global optimization method with a strong theoretical basis. Of these two obstacles, the former is becoming less serious today, as evidenced by a number of studies on global optimization in the past two decades \cite{hor90,ryo96,he04,rio13}. The aim of this paper is to partially address the latter obstacle. The inherent difficulty of the global optimization problem has led to two distinct research directions: development of heuristics without theoretically guaranteed performance and advancement of theoretically supported methods regardless of its difficulty. A degree of practical success has resulted from heuristic approaches such as simulated annealing, genetic algorithms (for a brief introduction on the context of AI, see \citeR{rus09}), and swarm-based optimization (for an interesting example of a recent study, see \citeR{dal09}). Although these methods are heuristics without strong theoretical supports, they became very popular partly because their optimization mechanisms aesthetically mimic nature's physical or biological optimization mechanism. On the other hand, the Lipschitzian approach to global optimization aims to accomplish the global optimization task in a theoretically supported manner. Despite its early successes in theoretical viewpoints \cite{shu72,mla86,pin86,han91}, the early studies were based on an assumption that is impractical in most applications: the Lipschitz constant, which is the bound of the slope on the objective function, is known. The relaxation of this crucial assumption resulted in the well-known DIRECT algorithm \cite{jon93} that has worked well in practice, yet guarantees only consistency property. Recently, the Simultaneous Optimistic Optimization (SOO) algorithm \cite{mun11} achieved the guarantee of a finite-time error bound without knowledge of the Lipschitz constant. However, the practical performance of the algorithm is unclear. In this paper, we propose a generic global optimization algorithm that is aimed to achieve both a satisfactory performance in practice and a finite-loss bound as the theoretical basis without strong additional assumption% \footnote{In this paper, we use the term ``strong additional assumption'' to indicate the assumption that the tight Lipschitz constant is known and/or the main assumption of many Bayesian optimization methods that the objective function is a sample from Gaussian process with some known kernel and hyperparameters.} (Section~\ref{sec2}), and apply it to an AI planning problem (Section~\ref{sec3}). For AI planning problem, we aim at solving real-world engineering problem with a long planning horizon and with continuous state/action space. As an illustration of the advantage of our method, we present the preliminary results of an application study conducted on accident management of a nuclear power plant as well. Note that the optimization problems discussed in this paper are practically relevant yet inherently difficult to scale up for higher dimensions, i.e., NP-complete \cite{murty1987some}. Accordingly, we discuss possible extensions of our algorithm for higher dimensional problems, with an experimental illustration with a 1000-dimensional problem. \section{Global Optimization on Black-Box Function}\label{sec2} The goal of global optimization is to solve the following very general problem: \begin{gather*} {\max}_x f(x)\\ \text{subject to }x\in \Omega \end{gather*} where $f$ is the objective function defined on the domain $\Omega\subseteq \mathbb{R}^D$. Since the performance of our proposed algorithm is independent of the scale of $\Omega$, we consider the problem with the rescaled domain $\Omega' = [0, 1]^D$. Further, in this paper, we focus on a deterministic function~% $f$. For global optimization, the performance of an algorithm can be assessed by the loss $r_n$, which is given by \[ r_n=\max_{x\in\Omega'} f(x) - f(x^+(n)). \] Here, $x^+(n)$ is the best input vector found by the algorithm after $n$ trials (more precisely, we define $n$ to denote the total number of divisions in the next section). A minimal assumption that allows us to solve this problem is that the objective function~$f$ can be evaluated at all points of $\Omega'$ in an arbitrary order. In most applications, this assumption is easily satisfied, for example, by having a simulator of the world dynamics or an experimental procedure that defines $f$ itself. In the former case, $x$ corresponds to the input vector for a simulator $f$, and only the ability to arbitrarily change the input and run the simulator satisfies the assumption. A possible additional assumption is that the gradient of function $f$ can be evaluated. Although this assumption may produce some effective methods, it limits the applicability in terms of real-world applications. Therefore, we assume the existence of a simulator or a method to evaluate $f$, but not an access to the gradient of $f$. The methods in this scope are often said to be \emph{derivative-free} and the objective function is said to be a \emph{black-box} function. However, if no further assumption is made, this very general problem is proven to be intractable. More specifically, any number of function evaluations cannot guarantee getting close to the optimal (maximum) value of $f$ \cite{dix78}. This is because the solution may exist in an arbitrary high and narrow peak, which makes it impossible to relate the optimal solution to the evaluations of $f$ at any other points. One of the simplest additional assumptions to restore the tractability would be that the slope of $f$ is bounded. The form of this assumption studied the most is Lipschitz continuity for~$f$: \begin{equation}\label{eq1} \abs{f(x_1)-f(x_2)}\le b\norm{x_1-x_2},\quad \forall x_1,x_2 \in \Omega', \end{equation} where $b > 0$ is a constant, called the \emph{Lipschitz constant}, and $\norm{\,\cdot\,}$ denotes the Euclidean norm. The global optimization with this assumption is referred to as \emph{Lipschitz optimization}, and has been studied for a long time. The best-known algorithm in the early days of its history was the Shubert algorithm \cite{shu72}, or equivalently the Piyavskii algorithm \cite{piy67} as the same algorithm was independently developed. Based on the assumption that the Lipschitz constant is known, it creates an upper bound function over the objective function and then chooses a point of $\Omega'$ that has the highest upper bound at each iteration. For problems with higher dimension $D\ge 2$, finding the point with the highest upper bound becomes difficult and many algorithms have been proposed to tackle the problem \cite{may84,mla86}. These algorithms successfully provided finite-loss bounds. However appealing from a theoretical point of view, a practical concern was soon raised regarding the assumption that the Lipschitz constant is known. In many applications, such as with a complex physics simulator as an objective function $f$, the Lipschitz constant is indeed unknown. Some researchers aimed to relax this somewhat impractical assumption by proposing procedures to estimate the Lipschitz constant during the optimization process \cite{str73,kva03}. Similarly, the Bayesian optimization method with upper confidence bounds \cite{bro09} \textit{estimates} the objective function and its upper confidence bounds with a certain model assumption, avoiding the prior knowledge on the Lipschitz constant. Unfortunately, this approach, including the Bayesian optimization method, results in mere heuristics unless the several additional assumptions hold. The most notable of these assumptions are that an algorithm can maintain the overestimate of the upper bound and that finding the point with the highest upper bound can be done in a timely manner. As noted by \citeA{han95}, it is unclear if this approach provides any advantage, considering that other successful heuristics are already available. This argument still applies to this day to relatively recent algorithms such as those by \citeA{kva03} and \citeA{bub11}. Instead of trying to estimate the unknown Lipschitz constant, the well-known DIRECT algorithm \cite{jon93} deals with the unknowns by simultaneously considering all the possible Lipschitz constants, $b$: $0 < b < \infty$. Over the past decade, there have been many successful applications of the DIRECT algorithm, even in large-scale engineering problems (\citeR{car01,he04}; \citeR{zwo05}). Although it works well in many practical problems, the DIRECT algorithm only guarantees consistency property, $\lim_{n\to\infty}r_n=0$ \cite{jon93,mun13}. The SOO algorithm \cite{mun11} expands the DIRECT algorithm and solves its major issues, including its weak theoretical basis. That is, the SOO algorithm not only guarantees the finite-time loss bound without knowledge of the slope's bound, but also employs a weaker assumption. In contrast to the Lipschitz continuity assumption used by the DIRECT algorithm (Equation~\eqref{eq1}), the SOO algorithm only requires the local smoothness assumption described below. \begin{assumption}[Local smoothness]\label{ass1} The decreasing rate of the objective function $f$ around at least one global optimal solution $\{x^*\in\Omega' : f(x^*)=\sup_{x\in\Omega'} f(x)\}$ is bounded by a~semi-metric $\ell$, for any $x\in\Omega'$ as \[ f(x^*)-f(x)\le\ell(x,x^*). \] \end{assumption} Here, semi-metric is a generalization of metric in that it does not have to satisfy the triangle inequality. For instance, $\ell(x,x^*) = b\norm{x^*-x}$ is a metric and a semi-metric. On the other hand, whenever $\alpha > 1$ or $p < 1$, $\ell(x,x^*) = b\norm{x^*-x}^\alpha_p$ is not a metric but only a semi-metric since it does not satisfy the triangle inequality. This assumption is much weaker than the assumption described by Equation~\eqref{eq1} for two reasons. First, Assumption~\ref{ass1} requires smoothness (or continuity) only at the global optima, while Equation~\eqref{eq1} does so for any points in the whole input domain, $\Omega'$. Second, while Lipschitz continuity assumption in Equation~\eqref{eq1} requires the smoothness to be defined by a metric, Assumption~\ref{ass1} allows a semi-metric to be used. To the best of our knowledge, the SOO algorithm is the only algorithm that provides a finite-loss bound with this very weak assumption. Summarizing the above, while the DIRECT algorithm has been successful in practice, concern about its weak theoretical basis led to the recent development of its generalized version, the SOO algorithm. We further generalize the SOO algorithm to increase the practicality and strengthen the theoretical basis at the same time. This paper adopts a very weak assumption, Assumption~\ref{ass1}, to maintain the generality and the wide applicability. \section{Locally Oriented Global Optimization (LOGO) Algorithm} In this section, we modify the SOO algorithm \cite{mun11} to accelerate the convergence while guaranteeing theoretical loss bounds. The new algorithm with this modification, the LOGO (Locally Oriented Global Optimization) algorithm, requires no additional assumption. To use the LOGO algorithm, one needs no prior knowledge of the objective function~% $f$; it may leverage prior knowledge if it is available. The algorithm uses two parameters, $h_\mathit{max} (n)$ and $w$, as inputs where $h_\mathit{max} (n)\in[1,\infty)$ and $w\in\mathbb{Z}^+$. $h_\mathit{max} (n)$ and~$w$ act in part to balance the local and global search. $h_\mathit{max} (n)$ biases the search towards a global search whereas $w$ orients the search toward the local area. The case with $w = 1$ or $4$ (top or bottom diagrams) in Figure~\ref{fig1} illustrates the functionality of the LOGO algorithm in a simple $2$-dimensional objective function. In this view, the LOGO algorithm is a generalization of the SOO algorithm with the local orientation parameter $w$ in that SOO is a special case of LOGO with a fixed parameter $w = 1$. \subsection{Predecessor: SOO Algorithm}\label{sec2.1} \begin{figure}[t] \small \includegraphics[width=0.92\textwidth]{fig1} \caption{Illustration of SOO ($w = 1$) and LOGO ($w = 1$ or $4$) at the end of each iteration}\label{fig1} \end{figure} Before we discuss our algorithm in detail, we briefly describe its direct predecessor, the SOO algorithm% \footnote{We describe SOO with the simple division procedure that LOGO uses. SOO itself does not specify a division procedure. } \cite{mun11}. The top diagrams in Figure~\ref{fig1} (the scenario with \(w=1\)) illustrates the functionality of the SOO algorithm in a simple $2$-dimensional objective function. As illustrated in Figure~\ref{fig1}, the SOO algorithm employs hierarchical partitioning to maintain hyperintervals, each center of which is the evaluation point of the objective function $f$. That is, in Figure~\ref{fig1}, each rectangle represents the hyperintervals at the end of each iteration of the algorithm. Let $\psi_h$ be a set of rectangles of the same size that are divided $h$ times. The algorithm uses a parameter, $h_\mathit{max} (n)$, to limit the size of rectangle so as to be not overly small (and hence restrict the greediness of the search). In order to select and refine intervals that are likely to contain a global optimizer, the algorithm executes the following procedure: \begin{enumerate}[(i)] \item\label{alg1-1} Initialize $\psi_0 = \{\Omega'\}$ and for all \(i>0\), $\psi_i = \{\varnothing\}$ \vspace{-7pt} \item\label{alg1i} Set $h = 0$ \vspace{-7pt} \item\label{alg1ii} Select the interval with the maximum center value among the intervals in the set $\psi_h$ \vspace{-7pt} \item\label{alg1iii}\label{alg1iv} If the interval selected by \eqref{alg1ii} has a center value greater than that of any \textit{larger} interval (i.e., intervals in \(\psi_l\) for all \(l<h\)), divide it and adds the new intervals to \(\psi_{h+1}\). Otherwise, reject the interval and skip this step. \vspace{-0pt} \item\label{alg1v} Set $h = h + 1$ \vspace{-7pt} \item\label{alg1vi} Repeat \eqref{alg1ii}--\eqref{alg1v} until no smaller interval exists (i.e., until $\psi_l=\{\varnothing\}$ for all \(l\ge h\)) or $h > h_\mathit{max}(n)$ \vspace{-7pt} \item\label{alg10} Delete all the intervals already divided in \eqref{alg1iii} from \(\psi \) and repeat \eqref{alg1i}--\eqref{alg1vi} \end{enumerate} We now explain this procedure using the example in Figure~\ref{fig1}. For brevity, we use the term, ``iteration'', to refer to the iteration of step \eqref{alg1i}--\eqref{alg10}. In Figure~\ref{fig1}, the center point is shown as a (black) dot in each rectangle and each rectangle with a (red) circle around a (black) dot is the one that was divided (into three smaller rectangles) during an iteration. At the beginning of the first iteration, there is only one rectangle in $\psi_0$, which is the entire search domain $\Omega'$. Thus, step \eqref{alg1ii} selects this rectangle and step \eqref{alg1iv} divides it, resulting in the leftmost diagram with $N = 3$ (the rectangle with the center point with a red circle is the one divided during the first iteration and the other two are created as a result). At the beginning of the second iteration, there are three rectangles in $\psi_1$ (i.e., the three rectangles in the leftmost diagram with $N = 3$) but none in $\psi_0$ (because step \eqref{alg10} in the previous iteration deleted the interval in $\psi_0$). Hence, steps \eqref{alg1ii}--\eqref{alg1iv} are not executed for $\psi_0$ and we begin with $\psi_1$. Step \eqref{alg1ii} selects the top rectangle from the three rectangles because it has the maximum center point among these. Step \eqref{alg1iii} divides it because there is no larger interval, resulting in the second diagram on the top (labeled with $w = 1$). Iteration 2 continues by conducting steps \eqref{alg1ii}--\eqref{alg1iv} for $\psi_2$ because there are three smaller rectangles in $\psi_2$. Step \eqref{alg1ii} selects the center rectangle on the top (in the second diagram on the top labeled with $w = 1$). However, step \eqref{alg1iii} rejects it because its center value is not greater than that of the larger rectangle in \(\psi_l\) with \(l<h=2\). There is no smaller rectangle in \(\psi\) and iteration 2 ends. At the beginning of iteration 3, there are two rectangles in $\psi_1$ and three rectangles in $\psi_2$ (as shown in the second diagram on the top labeled with $w = 1$). Iteration 3 begins by conducting steps \eqref{alg1ii}--\eqref{alg1iv} for $\psi_1$. Steps \eqref{alg1ii}--\eqref{alg1iii} select and divide the top rectangle. For rectangles in $\psi_2$, steps \eqref{alg1ii}--\eqref{alg1iii} select and divides the middle rectangle. Here, the middle rectangle was rejected in iteration 2 because of a larger rectangle with a larger center value that existed in iteration 2. However, that larger rectangle no longer exists in iteration 3 due to step \eqref{alg10} in the end of iteration 2, and hence it is not rejected. The result is the third diagram (on the top labeled with $w = 1)$. Iteration 3 continues for the newly created rectangles in $\psi_3$. It halts, however, for the same reason as iteration 2. \subsection{ Description of LOGO }\label{sec2.2} Let $\Psi_k$ be the superset that is the union of the $w$ sets as $\Psi_k=\psi_{kw}\cup\psi_{kw+1}\cup\dots\cup\psi_{kw+w-1}$ for $k=0,1,2,\dots$. Then, similar to the SOO algorithm, the LOGO algorithm conducts the following procedure to select and refine the intervals that are likely to contain a global optimizer: \begin{enumerate}[(i)] \item\label{alg1-1} Initialize $\psi_0 = \{\Omega'\}$ and for all \(i>0\), $\psi_i = \{\varnothing\}$ \vspace{-7pt} \item\label{alg2i} Set $k = 0$ \vspace{-7pt} \item\label{alg2ii} Select the interval with the maximum center value among the intervals in the superset~$\Psi_k$ \vspace{-7pt} \item\label{alg2iii}\label{alg2iv} If the interval selected by \eqref{alg1ii} has a center value greater than that of any \textit{larger} interval (i.e., intervals in $\Psi_l$ with $l<k$), divide it and adds the new intervals to \(\psi\). Otherwise, reject the interval and skip this step. \vspace{-7pt} \item\label{alg2v} Set $k = k + 1$ \vspace{-7pt} \item\label{alg2vi} Repeat \eqref{alg2ii}--\eqref{alg2v} until no smaller interval exists (i.e., until $\Psi_l=\{\varnothing\}$ for all $l\ge k$) or $k>\floor{h_\mathit{max}(n)/w}$. \item\label{alg10} Delete all the intervals already divided in \eqref{alg1iii} from \(\psi \) and repeat \eqref{alg1i}--\eqref{alg1vi} \end{enumerate} When compared with the SOO algorithm, the above steps are identical except that LOGO processes the superset $\Psi_k$ instead of set $\psi_h$. The superset $\Psi_k$ is reduced to $\psi_h$ with $k = h$ when $w = 1$ and thus LOGO is reduced to SOO. We now explain this procedure using the example in Figure~\ref{fig1}. With $w = 1$, the LOGO algorithm functions in the same fashion as the SOO algorithm. See the last paragraph in the previous section for the explanation as to how the SOO and LOGO algorithms function in this example. For the case with $w = 4$, the difference arises during iteration 3 when compared to the case with $w = 1$. At the beginning of iteration 3, there are two sets $\psi_1$ and $\psi_2$ (i.e., there are two sizes of rectangles in the second diagram on the bottom with $w = 4$). However, there is only one superset consisting of the two sets $\Psi_0=\psi_0\cup\psi_{0+1}\cup\psi_{0+4-1}$. Therefore, step \eqref{alg2ii}--\eqref{alg2iv} is conducted only for $k = 0$ and the LOGO algorithm divides only the one rectangle with the highest center value among those in $\Psi_0$. Consequently, the algorithm has one additional iteration (iteration 4) using the same number of function evaluations (\(N=9\)) as the case with $w = 1$. It can be seen that as $w$ increases, the algorithm is more biased to the local search and, in this example, this strategy turns out to be beneficial as the algorithm divides the rectangle near the global optima more when $w = 4$ than when $w = 1$. The pseudocode for the LOGO algorithm is provided in Algorithm~\ref{alg1}. Steps \eqref{alg2i}, \eqref{alg2v}, and \eqref{alg2vi} correspond to the for-loop in lines 10--19. Steps \eqref{alg2ii}--\eqref{alg2iii} correspond to line 11 and line 12--14, respectively. We use following notation. Each hyperrectangle, $\omega\subseteq\Omega'$, is coupled with a function value at its center point $f(c_\omega)$, where $c$ indicates the center point of the rectangle. As explained earlier, we use $h$ to denote the number of divisions and the index of the set as in $\psi_h$. We define $\omega_{h,i}$ to be the $i^\text{th}$ element of a set $\psi_h$ (i.e., $\omega_{h,i}\in\psi_h$). Let $x_{h,i}$ and $c_{h,i}$ be an arbitrary point and the center point in the rectangle $\omega_{h,i}$, respectively. We denote \(val[\omega_{h,i}]\) to indicate a stored function value of the center point in the rectangle \(\omega_{h,i}\). As it can be seen in line 14, this paper considers a simple division procedure with a rescaled domain \(\Omega'\). If we have prior knowledge about the domain of the function, we should leverage the information. For example, we could map the original input space to another so that we can obtain a better \(\ell\) based on the theoretical results in Section 4, or we could employ a more elaborate division procedure based on the prior knowledge. \begin{algorithm}[t] \caption{LOGO algorithm}\label{alg1} \begin{algorithmic}[1] \makeatletter \addtocounter{ALG@line}{-1} \makeatother \State {\bf Inputs (problem):} an objective function $f\colon x\in\mathbb{R}^D\to\mathbb{R}$, the search domain $\Omega$:~$x\in\nobreak \Omega$ \State {\bf Inputs (parameter):} the search depth function $h_\mathit{max}\colon\mathbb{Z}^+\to[1,\infty)$, the local weight $w\in\mathbb{Z}^+$, stopping condition \State Define the set $\psi_h$ as a set of hyperrectangles divided $h$ times \State Define the superset $\Psi_k$ as the union of the $w$ sets: $\Psi_k=\psi_{kw}\cup\psi_{kw+1}\cup\dots\cup\psi_{kw+w-1}$ \State Normalize the domain $\Omega$ to $\Omega' = [0, 1]^D$ \State Initialize the variables \begin{tabular}[t]{l} the set of hyperrectangles: $\psi_h=\{\varnothing\}$, $h=0,1,2,\dots$,\\ the current maximum index of the set: $h_\mathit{upper}=0$\\ the number of total divisions: $n=1$ \end{tabular} \State Adds the initial hyperrectangle $\Omega'$ to the set: $\psi_0\leftarrow\psi_0\cup\{\Omega'\}$ (i.e., $\omega_{0,0}=\Omega'$) \State Evaluate the function $f$ at the center point of $\Omega'$, $c_{0,0}$: $\mathit{val}[\omega_{0,0}]\leftarrow f(c_{0,0})$ \For{iteration ${}= 1, 2, 3,\dots$} \State $\mathit{val}_\mathit{max}\leftarrow-\infty$, $h_\mathit{plus}\leftarrow h_\mathit{upper}$ \For{$k=0,1,2,\dots,\max(\floor{\min(h_\mathit{max}(n),h_\mathit{upper})/w},h_\mathit{plus})$} \State \underline{Select}\ a hyperrectangle to be divided: $(h,i)\in\arg\max_{h,i}\mathit{val}[\omega_{h,i}]$ for $h,i:\omega_{h,i}\in\Psi_k$ \If{$\mathit{val}[\omega_{h,i}]>\mathit{val}_\mathit{max}$} \State $\mathit{val}_\mathit{max}\leftarrow\mathit{val}[\omega_{h,i}]$, $h_\mathit{plus}\leftarrow0$, $h_\mathit{upper}\leftarrow\max(h_\mathit{upper},h+1)$, $n\leftarrow n+1$ \State \underline{Divide}\ this hyperrectangle $\omega_{h,i}$ along the longest coordinate direction \Statex\qquad\qquad\quad\begin{tabular}[t]{l} - three smaller hyperrectangles are created $\rightarrow$ $\omega_\mathit{left}$, $\omega_\mathit{center}$, $\omega_\mathit{right}$\\ - $\mathit{val}[\omega_\mathit{center}]\leftarrow\mathit{val}[\omega_{h,i}]$ \end{tabular} \State \underline{Evaluate}\ the function $f$ at the center points of the two new hyperrectangles: \begin{center} $ \mathit{val}[\omega_\mathit{left}]\leftarrow f(c_{\omega_\mathit{left}}),\quad \mathit{val}[\omega_\mathit{right}]\leftarrow f(c_{\omega_\mathit{right}}) $ \end{center} \State \underline{\smash{Group}}\ the new hyperrectangles into the set $h + 1$ and remove the original rectangle: \begin{center} $ \psi_{h+1}\leftarrow\psi_{h+1} \cup\{\omega_\mathit{center},\omega_\mathit{left},\omega_\mathit{right}\},\quad \psi_h\leftarrow\psi_h\setminus \omega_{h,i} $ \end{center} \EndIf \State {\bf if} stopping condition is met {\bf then Return} $(h,i)=\arg\max_{h,i}\mathit{val}[\omega_{h,i}]$ \EndFor \EndFor \end{algorithmic} \end{algorithm} Having discussed how the LOGO algorithm functions, we now consider the reason why the algorithm might work well. The key mechanism of the DIRECT and SOO algorithms is to divide all the hyperintervals with potentially highest upper bounds w.r.t.\ unknown smoothness at each iteration. The idea behind the LOGO algorithm is to reduce the number of divisions per iteration by biasing the search toward the local area with the concept of the supersets. Intuitively, this can be beneficial for two reasons. First, by reducing the number of divisions per iteration, more information can be utilized when selecting intervals to divide. For example, one may simultaneously divide five or ten intervals per iteration. In the former, when selecting the sixth to the tenth interval to divide, one can leverage information gathered by the previous five divisions (evaluations), whereas the latter makes it impossible. Because the selection of intervals depends on the information, which in turn provides the new information to the next selection, the minor difference in availability of the information may make the two sequences of the search very different in the long run. Second, by biasing the search toward the local area, the algorithm likely converges faster in a certain type of problem. In many practical problems, we do not aim to find a position of global optima, but a position with a function value close to global optima. In that case, the local bias is likely beneficial unless there are too many local optima, the value of which is far from that of global optima. Even though our local bias strategy is motivated to solve the problem of the impractically slow convergence rate of most global optimization methods, the algorithm maintains guaranteed loss bounds w.r.t.\ global optima, as discussed below. \section{Theoretical Results: Finite-Time Loss Analysis}\label{sec2.3} We first derive the loss bounds for the LOGO algorithm when it uses any division strategy that satisfies certain assumptions. Then, we provide the loss bounds for the algorithm with the concrete division strategy provided in Algorithm~\ref{alg1} and with the parameter values that we use in the rest of this paper. The motivation in the first part is to extend the existing framework of the theoretical analysis and thus produce the basis for future work. The second part is to prove that the LOGO algorithm maintains finite-time loss bounds for the parameter settings that we actually use in the experiments. \subsection{Analysis for General Division Method}\label{sec2.3.1} In this section, we generalize the result obtained by \citeA{mun13} in that the previous result is now seen as a special case of the new result when $w = 1$. The previous work provided the loss bound of the SOO algorithm with any division process that satisfied the following two assumptions, which we adopt in this section. \begin{assA}[Decreasing diameter]\label{assA1} There exists a function $\delta(h)>0$ such that for any hyperinterval $\omega_{h,i}\subseteq\Omega'$, we have $\delta(h)\ge\sup_{x_{h,i}}\ell(x_{h,i},c_{h,i})$, while $\delta(h-1)\ge\delta(h)$ holds for $h\ge 1$. \end{assA} \begin{assA}[Well-shaped cell]\label{assA2} There exists a constant $\nu > 0$ such that any hyperinterval $\omega_{h,i}$ contains a $\ell$-ball of radius $\nu\delta(h)$ centered in $\omega_{h,i}$. \end{assA} Intuitively, Assumption~\ref{assA1} states that the unknown local smoothness $\ell$ is upper-bounded by a monotonically decreasing function of $h$. This assumption ensures that each division does not increase the upper bound, $\delta(h)$. Assumption~\ref{assA2} ensures that every interval covers at least a certain amount of space in order to relate the number of intervals to the unknown smoothness $\ell$ (because $\ell$ is defined in terms of space). To present our analysis, we need to define the relevant terms and variables. We define $\epsilon$-optimal set $X_\epsilon$ as \[ X_\epsilon:=\{x\in\Omega':f(x)+\epsilon\ge f(x^*)\}. \] That is, the set of $\epsilon$-optimal set $X_\epsilon$ is the set of input vectors whose function value is at least $\epsilon$-close to the value of the global optima. In order to bound the number of hyperintervals relevant to the $\epsilon$-optimal set $X_\epsilon$, we define \emph{near-optimality dimension} as follows. \begin{definition}[Near-optimality dimension]\label{def1} The near-optimality dimension is the smallest $d\ge 0$ such that there exists $C > 0$, for all $\epsilon > 0$, the maximum number of disjoint $\ell$-balls of radius $\nu\epsilon$ centered in the $\epsilon$-optimal set $X_\epsilon$ is less than or equal to $C\epsilon^{-d}$. \end{definition} The \emph{near-optimality dimension} was introduced by \citeA{mun11} and is closely related to a previous measure used by \citeA{kle08}. The value of the near-optimality dimension $d$ depends on the objective function $f$, the semi-metric $\ell$ and the division strategy (i.e., the constant $\nu$ in Assumption~\ref{assA2}). If we consider a semi-metric~$\ell$ that satisfies Assumptions~\ref{ass1}, \ref{assA1}, and \ref{assA2}, then the value of $d$ depends only on such a semi-metric~$\ell$ and the division strategy. In Theorem~\ref{the2}, we show that the division strategy of the LOGO algorithm can let $d = 0$ for a general class of semi-metric $\ell$. Now that we have defined the relevant terms and variables used in previous work, we introduce new concepts to advance our analysis. First, we define the set of $\delta$-optimal hyperinterval $\psi_h(w)^*$ as \[ \psi_h(w)^*:=\{\omega_{h,i}\subseteq \Omega': f(c_{\omega_{h,i}}) + \delta(h-w+1)\ge f(x^*)\}. \] The $\delta$-optimal hyperinterval $\psi_h(w)^*$ is used to relate the hyperintervals to $\epsilon$-optimal set $X_\epsilon$. Indeed, the $\delta$-optimal hyperinterval $\psi_h(w)^*$ is almost identical to the $\delta(h-w+1)$-optimal set $X_{\delta(h-w+1)}$ ($\epsilon$-optimal set $X_\epsilon$ with $\epsilon$ being $\delta(h-w+1)$), except that $\psi_h(w)^*$ only considers the hyperintervals and the values of their center points while $X_{\delta(h-w+1)}$ is about the whole input vector space. In order to relate $\psi_h(w)^*$ to $\psi_h(1)^*$, we define \emph{$\ell$-ball ratio} as follows. \begin{definition}[$\ell$-ball ratio]\label{def2} For every $h$ and $w$, the $\ell$-ball ratio is the smallest $\lambda_h(w)>0$ such that the volume of a $\ell$-ball of radius $\delta(h-w+1)$ is no more than the volume of $\lambda_h(w)$ disjoint $\ell$-balls of radius $\delta(h)$. \end{definition} In the following lemma, we bound the maximum cardinality of $\psi_h(w)^*$. We use $\abs{\psi_h(w)^*}$ to denote the cardinality. \begin{lemma}\label{lem1} Let $d$ be the near-optimality dimension and $C$ denote the corresponding constant in Definition~\ref{def1}. Let $\lambda_h(w)$ be the $\ell$-ball ratio in Definition~\ref{def2}. Then, the $\delta$-optimal hyperinterval is bounded as \[ \abs{\psi_h(w)^*}\le C\lambda_h(w)\delta(h-w+1)^{-d}. \] \end{lemma} \begin{proof} The proof follows the definition of $\epsilon$-optimal set $X_\epsilon$, Definition~\ref{def1}, Definition~\ref{def2}, and Assumption \ref{assA2}. From the definition of $\epsilon$-optimal space $X_\epsilon$, we can write $\delta(h-w+1)$-optimal set as \[ X_{\delta(h-w+1)}=\{x\in\Omega':f(x)+\delta(h-w+1)\ge f(x^*)\}. \] The definition of the near-optimality dimension (Definition~\ref{def1}) implies that at most $C\delta(h-w+1)^{-d}$ centers of disjoint $\ell$-balls of radius $\nu\delta(h-w+1)$ exist within space $X_{\delta(h-w+1)}$. Then, from the definition of the $\ell$-ball ratio (Definition~\ref{def2}), the space of $C\delta(h-w+1)^{-d}$ disjoint $\ell$-balls of radius $\nu\delta(h-w+1)$ is covered by at most $C\lambda_h(w)\delta(h-w+1)^{-d}$ disjoint $\ell$-balls of radius $\nu\delta(h)$. Notice that the set of space covered by $C\delta(h-w+1)^{-d}$ disjoint $\ell$-balls of radius $\nu\delta(h-w+1)$ is a superset of $X_{\delta(h-w+1)}$. Therefore, we can deduce that there are at most $C\lambda_h(w)\delta(h-w+1)^{-d}$ centers of disjoint $\ell$-balls of radius $\nu\delta(h)$ within $X_{\delta(h-w+1)}$. Now, recall the definition of the $h$-$w$-optimal interval, \[ \psi_h(w)^*:=\{\omega_{h,i}\subseteq\Omega':f(c_{\omega_{h,i}}) + \delta(h-w+1)\ge f(x^*)\} \] and notice that the number of intervals is equal to the number of centers $c_{\omega_{h,i}}$ that satisfy the condition $f(c_{\omega_{h,i}})+\delta(h-w+1)\ge f(x^*)$. Assumption \ref{assA2} causes this number to be equivalent to the number of centers of disjoint $\ell$-balls with radius $\nu\delta(h)$, which we showed to be upper bounded by $C\lambda_h(w)\delta(h-w+1)^{-d}$. \end{proof} Next, we bound the maximum size of \emph{the optimal hyperinterval}, which contains a global optimizer $x^*$. In the following analysis, we use the concept of the set and superset of hyperintervals. Recall that set $\psi_h$ contains all the hyperintervals that have been divided $h$ times thus far, and superset $\Psi_k$ is a union of the $w$ sets, given as $\Psi_k=\psi_{kw}\cup\psi_{kw+1}\cup\dots\cup\psi_{kw+w-1}$ for $k=0,1,2,\dots$. We say that a hyperinterval is \emph{dominated} by other intervals when the hyperinterval is not divided because its center value is at most that of other hyperintervals in any set. \begin{lemma}\label{lem2} Let $k_n^*$ be the highest integer such that the optimal hyperinterval, which contains a global optimizer $x^*$, belongs to the superset $\Psi_{k_n^*}$ after $n$ total divisions (i.e., $k_n^*\le n$ determines the size of the optimal hyperinterval, and hence the loss of the algorithm). Then, $k_n^*$ is lower bounded as $k_n^*\ge K$ with any $K$ that satisfies $0\le K\le\floor{h_\mathit{max}(n)/w}$ and \[ n\ge \bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^K\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}\biggr). \] \end{lemma} \begin{proof} Let $\tau(\Psi_k)$ be the number of divisions, with which the algorithm further divides the optimal hyperinterval in superset $\Psi_k$ and places it into $\Psi_{k+1}$. In the example in Figure~\ref{fig1} with $w = 1$, the optimal hyperinterval is initially the whole domain $\Omega'\subseteq\Psi_0$. It is divided with the first division and the optimal hyperinterval is placed into $\Psi_1$. Therefore, $\tau(\Psi_0)=1$. Similarly, $\tau(\Psi_1)=2$. A division of non-optimal interval occurs before that of the optimal one for $\tau(\Psi_2)$ and hence $\tau(\Psi_2)=4$. In other words, $\tau(\Psi_k)$ is the time when the optimal hyperinterval in superset $\Psi_k$ is further divided and escapes the superset $\Psi_k$, entering into $\Psi_{k+1}$. Let $c_{kw+l,i^*}$ be the center point of the optimal hyperinterval in a set $\psi_{kw+l}\subseteq\Psi_k$. We prove the statement by showing that the quantity $\tau(\Psi_k)-\tau(\Psi_{k-1})$ is bounded by the number of $\delta$-optimal hyperintervals $\psi_h(w)^*$. To do so, let us consider the possible hyperintervals to be divided during the time $[\tau(\Psi_{k-1}),\tau(\Psi_k)-1]$. For the hyperintervals in the set $\psi_{kw}$, the ones that can possibly further be divided during this time must satisfy $f(c_{kw,i})\ge f(c_{kw,i^*})\ge f(x^*) -\delta(kw)$. The first inequality is due to the fact that the algorithm does not divide an interval that has center value less than the maximum center value of an existing interval for each set, and there exists $f(c_{kw,i^*})$ during the time $[\tau(\Psi_{k-1}),\tau(\Psi_k)-1]$. The second inequality follows Assumption~\ref{ass1} and the definition of the optimal interval. Then, from the definition of $\psi_h(w)^*$, the hyperintervals that can possibly be divided during this time belong to $\psi_{kw}(1)^*\subseteq\Psi_k$. In addition to set $\psi_{kw}$, in superset $\Psi_k$, there are sets $\psi_{kw+l}$ with $l\colon w - 1 \ge l \ge 1$. For these sets, we have $f(c_{kw+l,i})\ge f(c_{lw+l,i^*})\ge f(x^*)-\delta(kw)$ with similar deductions. Here, notice that during the time $[\tau(\Psi_{k-1}),\tau(\Psi_k)-1]$, we can be sure that the center value in the superset is lower bounded by $f(c_{kw,i^*})$ instead of $f(c_{kw+l,i^*})$. In addition, we have $\delta(kw)=\delta(kw+l-l)$. Thus, we can conclude that the hyperintervals in set $\psi_{kw+l}$ that can be divided during time $[\tau(\Psi_{k-1}),\tau(\Psi_k)-1]$ belongs to $\psi_{kw+l}(l+1)^*$ where $(w - 1)\ge l\ge 1$. No hyperinterval in superset $k$ may be divided at iteration since the intervals can be dominated by those in other supersets. In this case, we have $f(c_{jw+l,i})\ge f(c_{kw,i^*})\ge f(x^*)-\delta(kw)$ for some $j < k$ and $l\ge 0$. With similar deductions, it is easy to see $f(x^*)-\delta(kw)\ge f(x^*)-\delta(jw+l)$. Thus, the hyperintervals in a superset $\Psi_j$ with $j < k$ that can dominate those in superset $\Psi_k$ during $[\tau(\Psi_{k-1}),\tau(\Psi_k)-1]$ belongs to $\psi_{jw+l}(1)^*$. Putting the above results together and noting that the algorithm divides at most $\floor{h_\mathit{max}(n)/w} + 1$ intervals during any iteration ($h_\mathit{plus}$ plays its role only when the algorithm divides at most one interval), we have \[ \tau(\Psi_k)-\tau(\Psi_{k-1}) \le \biggl(\bbfloor{\frac{h_\mathit{max}(n)}{w}}+1\biggr) \biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}+\sum_{j=1}^{k-1}\sum_{l=0}^{w-1}\abs{\psi_{jw+l}(1)^*}\biggr). \] Then, \[ \sum_{k=1}^{k_n^*}\tau(\Psi_k)-\tau(\Psi_{k-1}) \le \bbfloor{\frac{h_\mathit{max}(n)+w}{w}} \sum_{k=1}^{k_n^*}\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}\biggr) \] since the last term for a superset $\Psi_j$ with $j\le k-1$ in the previous inequality contains only the optimal intervals that are subsets of the optimal intervals covered by the new summation $\sum_{k=1}^{k_n^*}$. If $k_n^*\ge\floor{h_\mathit{max}(n)/w}$, then the statement always holds true for any $0\,{\le}\, K\,{\le}\, \floor{h_\mathit{max}(n)/w}$ since $k_n^*\ge\floor{h_\mathit{max}(n)/w}\ge K$. Accordingly, we assume $k_n^*<\floor{h_\mathit{max}(n)/w}$ in the following. Since $\tau(\Psi_0)$ is upper bounded by the term in the previous summation on the right hand of the above inequality with $k = 0$, \[ \tau(\Psi_{k_n^*+1})\le \bbfloor{\frac{h_\mathit{max}(n)+w}{w}} \sum_{k=0}^{k_n^*+1}\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}\biggr). \] By the definition of $k_n^*$, we have $n<\tau(\Psi_{k_n^*+1})$. Therefore, for any $K\le\floor{h_\mathit{max}(n)/w}$ such that \begin{multline*} \bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^{K}\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}\biggr)\\ \le n < \bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^{k_n^*+1}\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}\biggr), \end{multline*} we have $k_n^*\ge K$. \end{proof} With Lemmas~\ref{lem1} and \ref{lem2}, we can now present the main result in this section that provides the finite-time loss bound of the LOGO algorithm. \begin{theorem}\label{the1} Let $\ell$ be a semi-metric such that Assumptions~\ref{ass1}, \ref{assA1}, and \ref{assA2} are satisfied. Let $h(n)$ be the smallest integer $h$ such that \[ n\le C\bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^{\floor{h/w}}\biggl(\delta(kw)^{-d}+\sum_{l=1}^{w-1}\lambda_{kw+l}(l+1)\delta(kw)^{-d}\biggr). \] Then, the loss of the LOGO algorithm is bounded as \[ r_n\le\delta\bigl(\min(w\floor{h(n)/w}-w,w\floor{h_\mathit{max}(n)/w})\bigr). \] \end{theorem} \begin{proof} From Lemma~\ref{lem1} and the definition of $h(n)$, \begin{align*} n&>C\bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^{\floor{h(n)/w}-1}\biggl(\delta(kw)^{-d} +\sum_{l=1}^{w-1}\lambda_{kw+l}(l+1)\delta(kw+l-l)^{-d}\biggr)\\ &\ge \bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^{\floor{h(n)/w}-1}\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw+l}(l+1)^*}\biggr). \end{align*} Therefore, we set $K$ as $K=\floor{h(n)/w}-1$ in the following to apply the result of Lemma~\ref{lem2}. Then, it follows that $k_n^*\ge K(n)$ when $K<\floor{h_\mathit{max}(n)/w}$. Here, the number of divisions that an interval in the superset $\Psi_K$ is at least $Kw = w\floor{h(n)/w}-w$. Therefore, from Assumptions~\ref{ass1}, \ref{assA1}, and \ref{assA2}, we can deduce that $r_n\le\delta(w\floor{h(n)/w}-w)$. When $K\ge\floor{h_\mathit{max}(n)/w}$, we have $\floor{(h_\mathit{upper})/w}\ge k_n^*\ge\floor{h_\mathit{max}(n)/w}$. Thus, in this case, we have $k_n^*$ being equal to at least $\floor{h_\mathit{max}(n)/w}$. From Assumptions~\ref{ass1}, \ref{assA1}, and \ref{assA2}, we can similarly deduce that $r_n\le\delta(w\floor{h_\mathit{max}(n)/w})$. \end{proof} The loss bound stated by Theorem~\ref{the1} applies to the LOGO algorithm with any division strategy that satisfies Assumptions~\ref{assA1} and \ref{assA2}. We add the following assumption about the division process to derive more concrete forms of the loss bound. \begin{assA}[Decreasing diameter revisit]\label{assA3} The decreasing diameter defined in Assumption~\ref{ass1} can be written as $\delta(h) = c\gamma^{h/D}$ for some $c > 0$ and $\gamma < 1$, and accordingly the corresponding $\ell$-ball ratio is $\lambda_h(w)=(\delta(h-w+1)/\delta(h))^D$. \end{assA} Assumption~\ref{assA3} is similar to an assumption made by \citeA{mun13}, which is that $\delta(h) =\nobreak c\gamma^h$. In contrast to the previous assumption, our assumption explicitly reflects the fact that the size of a hyperinterval decreases at a slower rate for higher dimensional problems. For the LOGO algorithm, the validity of Assumptions~\ref{assA1}, \ref{assA2}, and \ref{assA3} is confirmed in the next section. We now present the finite-loss bound for the LOGO algorithm in the case of the general division strategy with the above additional assumption and with $d = 0$. \begin{corollary}\label{cor1} Let $\ell$ be a semi-metric such that Assumptions~\ref{ass1}, \ref{assA1}, \ref{assA2}, and \ref{assA3} are satisfied. If the near-optimality dimension $d = 0$ and $h_\mathit{max}(n)$ is set to $\sqrt{n}-w$, then the loss of the LOGO algorithm is bounded for all $n$ as \[ r_n\le c\exp\Biggl(-\min\biggl(\sqrt{n}\frac{w}{C}\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}-2, \sqrt{n}-w\biggr)\frac{w}{D}\ln\frac1\gamma\Biggr). \] \end{corollary} \begin{proof} Based on the definition of $h(n)$ in Theorem~\ref{the1}, we first relate $h(n)$ to $n$ as \begin{align*} n&\le C\frac{h_\mathit{max}(n)+w}{w}\sum_{k=0}^{\floor{h(n)/w}}\biggl(\delta(kw)^{-d}+\sum_{l=1}^{w-1}\lambda_{kw+l}(l+1)\delta(kw+l-l)^{-d}\biggr)\\ &=C\frac{h_\mathit{max}(n)+w}{w}\sum_{k=0}^{\floor{h(n)/w}}\biggl(1+\sum_{l=1}^{w-1}\gamma^{-w}\biggr) \le C\frac{h_\mathit{max}(n)+w}{w}\biggl(\bbfloor{\frac{h(n)}w}+1\biggr)\sum_{l=0}^{w-1}\gamma^{-w}. \end{align*} The first line follows the definition of $h(n)$, and the second line is due to $d = 0$ and Assumption~\ref{assA3}. By algebraic manipulation, \[ \bbfloor{\frac{h(n)}{w}}\ge \frac{n}C\frac{w}{h_\mathit{max}(n)+w}\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}-1. \] Here, we use $h_\mathit{max}(n)=\sqrt{n}-w$, and hence \[ \bbfloor{\frac{h(n)}{w}}\ge \sqrt{n}\frac{w}C\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}-1. \] By substituting these results into the statement of Theorem~\ref{the1}, \[ r_n\le \delta\Biggl(\min\biggl(\sqrt{n}\frac{w^2}C\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}-2w, w\sqrt{n}-w^2\biggr)\Biggr). \] From Assumption~\ref{assA3}, $\delta(h) = c\gamma^{h/D}$. By using $\delta(h) = c\gamma^{h/D}$ in the above inequality, we have the statement of this corollary. \end{proof} \begin{figure} \centering \hair2pt \labellist \pinlabel $\gamma$ [r] at 0 75 \pinlabel $w$ [t] at 100 0 \endlabellist \includegraphics[width=0.4\textwidth]{fig2} \vskip5mm \caption{Effect of local bias $w$ on loss bound in the case of $d = 0 :=w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$}\label{fig2} \end{figure} Corollary~\ref{cor1} shows that the LOGO algorithm guarantees an exponential bound on the loss in terms of $\sqrt{n}$ (a stretched exponential bound in terms of $n$). The loss bound in Corollary~\ref{cor1} becomes almost identical to that of the SOO algorithm with $w = 1$. Accordingly, we illustrate the effect of $w$, when $n$ is large enough to let us focus on the coefficient of $\sqrt{n}$, in Figure~\ref{fig2}. The (red) bold line with label 1 indicates the area where $w$ has no effect on the bound. The area with lines having labels greater than one is where $w$ improves the bound, and the area with labels less than one is where $w$ diminishes the bound. More concretely, in the figure, we consider the ratio of the coefficient of $\sqrt{n}$ in the loss bound with the various value of $w$ to that with $w = 1$. The ratio is $w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$ or $w$, depending on which element of the min in the bound is smaller. Since $w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$ is at most $w$ (in the domain we consider), we plotted $w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$ to avoid overestimating the benefit of $w$. Thus, this is a rather pessimistic illustration of the advantage of our generalization regarding $w$. For instance, if the second element of the min in the bound is smaller and $n$ is large enough, increasing $w$ always improves the bound, regardless of the values in Figure~\ref{fig2}. The next corollary presents the finite-loss bound for the LOGO algorithm in the case of $d\neq 0$. \begin{corollary}\label{cor2} Let $\ell$ be a semi-metric such that Assumptions~\ref{ass1}, \ref{assA1}, \ref{assA2}, and \ref{assA3} are satisfied. If the near-optimality dimension $d > 0$ and $h_\mathit{max}(n)$ is set to be $\Theta((\ln n)^{c_1})-w$ for some $c_1 > 1$, the loss of the LOGO algorithm is bounded as \[ r_n\le \tilde{O}\Biggl(n^{-1/d}\biggl(w^2(\gamma^{wd/D}-\gamma^{2wd/D})\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}\biggr)^{\!-1/d}\,\Biggr). \] \end{corollary} \begin{proof} In the same way as the first step in the proof of Corollary~\ref{cor1}, except for $d > 0$, \[ n\le Cc^{-d}\frac{h_\mathit{max}(n)+w}{w}\sum_{k=0}^{\floor{h(n)/w}}\sum_{l=0}^{w-1}\gamma^{-l-kwd/D}. \] The reason why we could not bound the loss in a similar rate as in the case $d = 0$ is that the last summation term $\sum_{l=0}^{w-1}$ is no longer independent of $k$. Since \[ \sum_{l=0}^{w-1}\gamma^{-l-kwd/D} = \gamma^{-kwd/D}\frac{\gamma^{-w}-1}{\gamma^{-1}-1},\quad \sum_{k=0}^{\floor{h(n)/w}}\gamma^{-kwd/D} = \frac{\gamma^{-(\floor{h(n)/w}+1)wd/D}-1}{\gamma^{-wd/D}-1}, \] with algebraic manipulation, \[ c^{-d}(\gamma^{-(\floor{h(n)/w}+1)wd/D}-1) \ge \frac{n}C\frac{w}{h_\mathit{max}(n)+w}\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}(\gamma^{-wd/D}-1). \] Therefore, \[ c\gamma^{(w\floor{h(n)/w}-w)/D} \ge \Biggl(\frac{n}C\frac{w}{h_\mathit{max}(n)+w}\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}(\gamma^{-wd/D}-1)\gamma^{2wd/D}\Biggr)^{\!-1/d}. \] From Theorem~\ref{the1} and Assumption~\ref{assA3}, \[ r_n\le\max\Biggl(\biggl(\frac{n}C\frac{w}{h_\mathit{max}(n)+w} \biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1} (\gamma^{wd/D}-\gamma^{2wd/D})\biggr)^{\!-1/d},c\gamma^{(w\floor{h_\mathit{max}(n)/w}-w)/D}\Biggr). \] For $h_\mathit{max}(n)=\Theta((\ln n)^{c_1})-w$ and for a sufficiently large $n$, the first element of the previous max becomes larger than the second one, and its order is equivalent to the one in the statement. \end{proof} We derived the loss bound for the SOO algorithm with Assumption~\ref{assA3} in the case of $d\neq 0$ as well. The SOO version of the loss bound is $r_n\le \tilde{O}(n^{-1/d}(\gamma^{d/D}-\gamma^{2d/D})^{-1/d})$, which is equivalent to the loss bound of the LOGO algorithm with $w = 1$ in Corollary~\ref{cor2}. In Figure~\ref{fig3}, we thereby illustrate the effect of $w$ on the loss bound in the $\tilde{O}$ form. In the figure, we plotted the ratio of the elements inside $\tilde{O}$ of the loss bounds. From Figure~\ref{fig2} and Figure~\ref{fig3}, we can infer that the loss bound is improved with $w > 1$ if $\gamma$ is large and $d$ is small (when $n$ is sufficiently large). Intuitively, this makes sense, since there are more of the different yet similar sizes of hyperintervals w.r.t.\ $\ell$ if $\gamma$ is larger and $d$ is smaller. In that case, dividing all the hyperintervals in the marginally different sizes would be redundant and a waste of computational resources. Note that our discussion here is limited to the loss bound that we have now, which may be tightened in future work. We would then see the different effects of $w$ on such tightened bounds. \begin{figure} \captionsetup[subfigure]{labelformat=empty} \centering \begin{subfigure}[b]{0.296\textwidth} \hair2pt \labellist \pinlabel $\gamma$ [r] at 12 55 \pinlabel $w$ [t] at 75 -5 \endlabellist \includegraphics[width=1.13\textwidth]{fig3a}\hfill \vskip5mm \caption{(a) $d=0.01$} \end{subfigure}\hfill \begin{subfigure}[b]{0.31\textwidth} \labellist \pinlabel $\gamma$ [r] at 5 55 \pinlabel $w$ [t] at 75 -5 \endlabellist \includegraphics[width=1.01\textwidth]{fig3b}\hfill \vskip5mm \caption{(b) $d=0.5$} \end{subfigure}\hfill \begin{subfigure}[b]{0.31\textwidth} \labellist \pinlabel $\gamma$ [r] at 2 55 \pinlabel $w$ [t] at 70 -5 \endlabellist \includegraphics[width=0.98\textwidth]{fig3c} \vskip5mm \caption{(c) $d=1.0$} \end{subfigure} \caption{Effect of local bias $w$ on loss bound in the case of $d\neq 0 :=(w^2(\gamma^{wd/D}-\gamma^{2wd/D}(\frac{\gamma^{-w}-1}{\gamma^{-1}-1})^{-1})^{-1/d}/(\gamma^{d/D}-\gamma^{2d/D})^{-1/d}$}\label{fig3} \end{figure} \subsection{Basis of Practical Usage}\label{sec2.3.2} In this section, we derive the loss bound of the LOGO algorithm with the concrete division strategy presented in Section~\ref{sec2.1}. The purpose of this section is to analyze the LOGO algorithm with the division process and the parameter settings that are actually used in the rest of this paper. The results of this section are directly applicable to our experiments. In this section, we discard Assumptions~\ref{assA1}, \ref{assA2}, and \ref{assA3}. We consider the following assumption to present the loss bound in a concrete form. \begin{assB}\label{assB1} There exists a semi-metric $\ell$ such that that it satisfies Assumption~\ref{ass1} and both of the following conditions hold: \begin{itemize} \item there exist $b > 0$, $\alpha > 0$ and $p\ge1$ such that for all \(x,y \in \Omega'\), $\ell(x,y)=b\norm{x-y}^\alpha_p$ \item there exist \(\theta \in (0,1)\) such that for all \(x \in \Omega'\), $ f({ x^ * }) \ge f( x) + \theta \ell \left( { x,{ x^ * }} \right)$. \end{itemize} \end{assB} First, we state that the loss bound of the algorithm with the practical division process and parameter settings decreases at a stretched exponential rate. \begin{theorem}[worst-case analysis]\label{the2} Let $\ell$ be a semi-metric such that Assumptions~\ref{ass1} and \ref{assB1} are satisfied. The loss of the LOGO algorithm is bounded as \[ r_n\le c\exp\Biggl(-\min\biggl(\sqrt{n}\frac{w}{w'C}\biggl(\frac{\gamma^{-w}-1}{\gamma^{-1}-1}\biggr)^{\!-1}-2, w'\sqrt{n}-w\biggr)\frac{w}{D}\ln\frac1{\gamma}\Biggr) \] where $\gamma=3^{-\alpha}$ and $c=b3^\alpha D^{\alpha/p}$. Here, $w'=1$ if we set the parameter as $h_\mathit{max}(n)=\sqrt{n}-w$. On the other hand, $w'=w$ if we set the parameter as $h_\mathit{max}(n)=w\sqrt{n}-w$. \end{theorem} \begin{proof} From Assumption \ref{assB1} and the division strategy, \[ {\sup}_{x_{h,i}}\ell(x_{h,i},c_{h,i})\le b(3^{-\floor{h/D}}D^{1/p})^\alpha = bD^{\alpha/p}3^{-\alpha\floor{h/D}} \] which corresponds to the \emph{diagonal} length of each rectangle, while $3^{-\alpha\floor{h/D}}$ corresponds to the length of the \emph{longest} side. This quantity is upper bounded by $3^{-\alpha h/D +\alpha}$. Thus, we consider the case where $\delta(h)=b3^{\alpha} D^{\alpha/p}3^{-\alpha h/D}$, which satisfies Assumption~\ref{assA1}. Also, Assumption~\ref{assA3} is satisfied for $\delta(h)$ with $\gamma=3^{-\alpha}$ and $c=b3^\alpha D^{\alpha/p}$. Every rectangle contains at least a $\ell$-ball of radius corresponding to the length of the \emph{shortest} side for the rectangle. Consequently, we have at least a $\ell$-ball of radius $\nu\delta(h)=b3^{-\alpha}3^{-\alpha h/D}$ for any rectangle where $\nu=3^{-2\alpha}D^{-\alpha/p}$, which satisfies Assumption~\ref{assA2}. From Assumption~\ref{assB1}, the volume $V$ of a $\ell$-ball of radius $\nu\delta(h)$ is proportional to $(\nu\delta(h))^D$ as the following: $V^p_D(\nu\delta(h))=(2\nu\delta(h)\Gamma(1+1/p))^D / \Gamma(1+D/p)$. Therefore, Assumption~\ref{assA3} is satisfied for the $\ell$-ball ratio $\lambda_h(w)$. In addition, the \emph{$\delta(h)$-optimal set} $X_{\delta(h)}$ is covered by a $\ell$-ball of radius $\delta(h)$ by Assumption \ref{assB1}, and thereby contains at most $(\delta(h)/\nu\delta(h))^D = \nu^{-D}$ disjoint $\ell$-balls of radius $\nu\delta(h)$. Hence, the number of the $\ell$-balls does not depend on $\delta(h)$, which means $d = 0$. Now that we have satisfied Assumptions~\ref{assA1}, \ref{assA2}, and \ref{assA3} with $\gamma=3^{-\alpha}$, $c=b3^\alpha D^{\alpha/p}$, and $d = 0$, we obtain the statement by following the proof of Corollary~\ref{cor1}. \end{proof} Regarding the effect of local orientation $w$, Theorem~\ref{the2} presents the worst-case analysis. Recall that $w$ is introduced in this paper to restore the practicality of global optimization methods. Thus, focusing on the worst case is likely too pessimistic. To mitigate this problem, we present the following optimistic analysis. \begin{theorem}[best-case analysis in terms of $w$]\label{the3} Let $\ell$ be a semi-metric such that Assumptions~\ref{ass1} and \ref{assB1} are satisfied. For $1\le l\le w$, let $\omega_{h+l-1,i'}$ be any hyperinterval that may dominate other intervals in the set $\psi_h$ during the algorithm's execution. Assume that $\omega_{h+l-1,i'}\subseteq\psi_h(1)^*$. Then, the loss of the LOGO algorithm is bounded as \[ r_n\le c\exp\biggl(-\min\biggl(\sqrt{n}\frac1{w'C}-2, w'\sqrt{n}-w\biggr)\frac{w}D\ln\frac1{\gamma}\biggr) \] where $\gamma=3^{-\alpha}$ and $c=b3^\alpha D^{\alpha/p}$. Here, $w'=1$ if we set the parameter as $h_\mathit{max}(n)=\sqrt{n}-w$. On the other hand, $w'=w$ if we set the parameter as $h_\mathit{max}(n)=w\sqrt{n}-w$. \end{theorem} \begin{proof} The statement of Lemma~\ref{lem2} is modified as \[ n\ge\bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^K\biggl(\abs{\psi_{kw}(1)^*}+\sum_{l=1}^{w-1}\abs{\psi_{kw}(1)^*}\biggr). \] The statement of Theorem~\ref{the1} is modified as \begin{align*} n&\le C\bbfloor{\frac{h_\mathit{max}(n)+w}{w}}\sum_{k=0}^{\floor{h/w}}(w\delta(kw)^{-d}),\\ r_n&\le\delta\bigl(\min(w\floor{h(n)/w}-w,w\floor{h_\mathit{max}(n)/w})\bigr). \end{align*} Then, we can follow the proof of Theorem~\ref{the2} and Corollary~\ref{cor1}, obtaining \[ \bbfloor{\frac{h(n)}{w}}\ge\frac{n}C\frac{1}{h_\mathit{max}(n)+w}-1. \] With $h_\mathit{max}(n)=\sqrt{n}-w$, from the modified statement of Theorem~\ref{the1}, we obtain the statement of this theorem. \end{proof} As Theorem~\ref{the3} makes a strong assumption to eliminate the negative effect of the local orientation in the bound, increasing $w$ always improves the loss bound in the theorem when $n$ is sufficiently large. This may seem to be overly optimistic, but we show an instance of this case in our experiment. \begin{figure}[t!] \captionsetup[subfigure]{labelformat=empty} \centering \begin{subfigure}[b]{0.45\textwidth} \hair2pt \labellist \pinlabel $\alpha$ [r] at 18 90 \pinlabel $w$ [t] at 115 0 \endlabellist \includegraphics[width=.98\textwidth]{fig4a} \vskip5mm \caption{(a) Pessimistic $w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$} \end{subfigure} \hfill \begin{subfigure}[b]{0.45\textwidth} \hair2pt \labellist \pinlabel $\alpha$ [r] at 1 90 \pinlabel $w$ [t] at 100 0 \endlabellist \includegraphics[width=0.9\textwidth]{fig4b} \vskip5mm \caption{(b) Optimistic $w^2$} \end{subfigure} \caption{Effect of local bias $w$ on loss bound with practical setting. The real effect would exist somewhere in-between.}\label{fig4} \end{figure} More realistically, the effect of $w$ with large $n$ would exist somewhere between the left and the right diagrams in Figure~\ref{fig4}. As in the previous figures, the (red) bold line with label~$1$ is where $w$ has no effect on the bound, the area with labels greater than one is where $w$ improves the bound, and the area with labels less than one is where $w$ diminishes the bound. The left diagram shows the effect of $w$ in the worst case of Theorem~\ref{the2} by plotting $w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$ with $\gamma=3^{-\alpha}$. The reason why plotting $w^2(\gamma^{-1}-1)/(\gamma^{-w}-1)$ represents the worst case is discussed in the previous section. The right diagram presents the effect of $w$ in the best case of Theorem~\ref{the2} or Theorem~\ref{the3} by simply plotting $w^2$. Notice that in both Theorem~\ref{the2} and Theorem~\ref{the3}, the best scenario for the effect of $w$ is when we use $h_\mathit{max}(n)=w\sqrt{n}-w$ and the second element of the min dominates the bound. In this case, the coefficient of $\sqrt{n}$ is $w^2$, which is the effect of $w$ on the bound when $n$ is large enough to ignore the other term. In conclusion, we showed that the LOGO algorithm provides a stretched exponential bound on the loss with the algorithm's division strategy, which is likely more practical than the one used in the analysis of the SOO algorithm, and with the parameter setting $h_\mathit{max}(n)=\sqrt{n}-w$ or $h_\mathit{max}(n)=w\sqrt{n}-w$. We also discussed how the local bias $w$ affects the loss bound. Based on these results, we use the LOGO algorithm in the following experiments. \section{Experimental Results}\label{sec2.4} In this section, we test the LOGO algorithm with a series of experiments. In the main part of the experiments, we compared the LOGO algorithm with its direct predecessor, the SOO algorithm \cite{mun11} and its latest powerful variant, the Bayesian Multi-Scale Optimistic Optimization (BaMSOO) algorithm \cite{wan14}. The BaMSOO algorithm combines the SOO algorithm with a Gaussian Process (GP) to leverage the GP's estimation of the upper confidence bound. It was shown to outperform the traditional Bayesian Optimization method that uses a GP and the DIRECT algorithm \cite{wan14}. Accordingly, we omitted the comparison with the traditional Bayesian Optimization method. We also compare LOGO with popular heuristics, simulated annealing (SA) and genetic algorithm (GA) (see \citeR{rus09} for a brief introduction). In the experiments, we rescaled the domains to the $[0, 1]^D$ hypercube. We used the same division process for SOO, BaMSOO and LOGO, which is the one presented in Section~\ref{sec2.2} and proven to provide stretched exponential bounds on the loss in Section~\ref{sec2.3.2}. Previous algorithms have also been used with this division process in experiments \cite{jon93,gab01,mun13,wan14}. For the SOO and LOGO algorithms, we set $h_\mathit{max}(n) =w\sqrt{n}-w$. This setting guarantees a stretched exponential bound for LOGO, as proven in Section~\ref{sec2.3.2}, and for SOO \cite{mun13}. For the LOGO algorithm, we used a simple adaptive procedure to set the parameter $w$. Let $f(x_i^+)$ be the best value found thus far in the end of iteration $i$. Let $W = \{3, 4, 5, 6, 8, 30\}$. The algorithm begins with $w = W_1 = 3$. At the end of iteration $i$, the algorithm set $w = W_k$ with $k=\min(j+1,6)$ if $f(x_i^+)\ge f(x_{i-1}^+)$, and $k=\max(j-1,1)$ otherwise, where $W_j$ is the previous parameter value $w$ before this adjustment occurs. Intuitively, this adaptive procedure is to encourage the algorithm to be locally biased when it seems to be making progress, forcing it to explore a more global region when this does not seem to be the case. Although the values in the set $W = \{3, 4, 5, 6, 8, 30\}$ are arbitrary, this simple setting was used in all the experiments in this paper, including the real-world application in Section~\ref{sec3.4}. The results demonstrate the robustness of this setting. As discussed later, a future work would be to replace this simple adaptive mechanism to improve the performance of the proposed algorithm. For the BaMSOO algorithm, the previous work of \citeA{wan14} used a pair of a good kernel and hyperparameters that were handpicked for each test function. In our experiments, we assumed that such a handpicking procedure was unavailable, which is typically the case in practice. We tested several pairs of a kernel and hyperparameters; however, none of the pairs performed robustly well for all the test functions (e.g., one pair performed well for one test function, although not others). Thus, we used the empirical Bayes method to adaptively update the hyperparameters\footnote{We implemented BaMSOO by ourselves to use the empirical Bayes method, which was not done in the original implementation. The original implementation of BaMSOO was not available for us as well. }. We selected the isotropic Matern kernel with $\nu=5/2$, which is given by $\kappa(x,x')=g(\sqrt{5\norm{x-x'}^2/l})$, where the function $g$ is defined to be $g(z)=\sigma^2(1+z+z^3/3)$. The hyperparameters were initialized to $\sigma=1$ and $l=0.25$. We updated the hyperparameters every iteration until $1{,}000$ function evaluations were executed and then per $1{,}000$ iterations afterward (to reduce the computational cost). For SA and GA, we used the same settings as those of the Matlab standard subroutines \texttt{simulannealbnd} and \texttt{ga}, except that we specified the domain bounds. \begin{table} \centering \begin{scriptsize} \renewcommand{\tabcolsep}{1.45pt} \renewcommand{\arraystretch}{1.4} \begin{tabular}{|l|c|c|c|c|c|c|c|c|c|c|c|} \hline \multicolumn{1}{|c|}{\multirow{2}{*}{$f$}} & \multirow{2}{*}{$D$} & \multirow{2}{*}{$\Omega$} & \multicolumn{3}{|c|}{SOO} & \multicolumn{3}{|c|}{BaMSOO} & \multicolumn{3}{|c|}{LOGO}\\ \cline{4-12} &&&$N$ & $\mathit{Time}\,(s)$ & $\mathit{Error}$ & $N$ & $\mathit{Time}\,(s)$ & $\mathit{Error}$ & $N$ & $\mathit{Time}\,(s)$ & $\mathit{Error}$\\ \hline Sin~1 & $1$ & $[0,1]$ & $57$ & $5.3\,\mathrm{E}{-}02$ & $2.3\,\mathrm{E}{-}06$ & $30$ & $2.0\,\mathrm{E}{+}00$ & $2.3\,\mathrm{E}{-}06$ & $17$ & $3.9\,\mathrm{E}{-}02$ & $2.3\,\mathrm{E}{-}06$\\ \hline Sin~2 & $2$ & $[0,1]^2$ & $271$ & $1.7\,\mathrm{E}{-}01$ & $4.6\,\mathrm{E}{-}06$ & $181$ & $7.5\,\mathrm{E}{+}00$ & $4.6\,\mathrm{E}{-}06$ & $45$ & $5.4\,\mathrm{E}{-}02$ & $4.6\,\mathrm{E}{-}06$\\ \hline Peaks & $2$ & $[-3,3]^2$ & $141$ & $1.0\,\mathrm{E}{-}01$ & $9.0\,\mathrm{E}{-}05$ & $37$ & $3.5\,\mathrm{E}{+}00$ & $9.0\,\mathrm{E}{-}05$ & $35$ & $6.1\,\mathrm{E}{-}02$ & $9.0\,\mathrm{E}{-}05$\\ \hline Branin & $2$ & $[-5,10]\times[0,15]$ & $339$ & $2.1\,\mathrm{E}{-}01$ & $9.0\,\mathrm{E}{-}05$ & $121$ & $8.1\,\mathrm{E}{+}00$ & $9.0\,\mathrm{E}{-}05$ & $85$ & $7.0\,\mathrm{E}{-}02$ & $8.7\,\mathrm{E}{-}05$\\ \hline Rosenbrock~2 & $2$ & $[-5,10]^2$ & $491$ & $3.1\,\mathrm{E}{-}01$ & $9.7\,\mathrm{E}{-}06$ & \cellcolor{gray!50}${>}4000$ & \cellcolor{gray!50}$5.8\,\mathrm{E}{+}04$ & \cellcolor{gray!50}$5.5\,\mathrm{E}{-}03$ & $137$ & $1.3\,\mathrm{E}{-}01$ & $9.7\,\mathrm{E}{-}06$\\ \hline Hartman~3 & $3$ & $[0,1]^3$ & $359$ & $2.3\,\mathrm{E}{-}01$ & $7.91\,\mathrm{E}{-}05$ & $126$ & $8.9\,\mathrm{E}{+}00$ & $7.9\,\mathrm{E}{-}05$ & $65$ & $7.1\,\mathrm{E}{-}02$ & $5.1\,\mathrm{E}{-}05$\\ \hline Shekel~5 & $4$ & $[0,10]^4$ & $1101$ & $6.6\,\mathrm{E}{-}01$ & $8.4\,\mathrm{E}{-}05$ & $316$ & $3.1\,\mathrm{E}{+}01$ & $8.4\,\mathrm{E}{-}05$ & $157$ & $1.2\,\mathrm{E}{-}01$ & $8.4\,\mathrm{E}{-}05$\\ \hline Shekel~7 & $4$ & $[0,10]^4$ & $1117$ & $7.1\,\mathrm{E}{-}01$ & $9.4\,\mathrm{E}{-}05$ & $95$ & $1.2\,\mathrm{E}{+}01$ & $9.4\,\mathrm{E}{-}05$ & $157$ & $1.2\,\mathrm{E}{-}01$ & $9.4\,\mathrm{E}{-}05$\\ \hline Shekel~10 & $4$ & $[0,10]^4$ & $1117$ & $6.4\,\mathrm{E}{-}0.1$ & $9.68\,\mathrm{E}{-}05$ & \cellcolor{gray!50}${>}4000$ & \cellcolor{gray!50}$4.5\,\mathrm{E}{+}04$ & \cellcolor{gray!50}$8.1\,\mathrm{E}{+}00$ & $197$ & $1.5\,\mathrm{E}{-}01$ & $9.7\,\mathrm{E}{-}05$\\ \hline Hartman~6 & $6$ & $[0,1]^6$ & $1759$ & $1.2\,\mathrm{E}{+}00$ & $7.51\,\mathrm{E}{-}05$ & \cellcolor{gray!50}${>}4000$ & \cellcolor{gray!50}$4.0\,\mathrm{E}{+}04$ & \cellcolor{gray!50}$2.3\,\mathrm{E}{-}03$ & $161$ & $1.3\,\mathrm{E}{-}01$ & $6.8\,\mathrm{E}{-}05$\\ \hline Rosenbrock~10 & $10$ & $[-5,10]^{10}$ & \cellcolor{gray!50}${>}8000$ & \cellcolor{gray!50}$7.8\,\mathrm{E}{+}00$ & \cellcolor{gray!50}$3.83\,\mathrm{E}{-}03$ & \cellcolor{gray!50}${>}8000$ & \cellcolor{gray!50}$5.8\,\mathrm{E}{+}04$ & \cellcolor{gray!50}$9.6\,\mathrm{E}{+}00$ & $1793$ & $1.7\,\mathrm{E}{+}00$ & $4.8\,\mathrm{E}{-}05$\\ \hline \end{tabular} \end{scriptsize} \caption{Performance comparison in terms of the number of evaluations ($N$) and CPU time ($\mathit{Time}$) to achieve $\mathit{Error} < 10^{-4}$. The grayed cells indicate the experiments where we could not achieve $\mathit{Error} < 10^{-4}$ even with a large number of function evaluations ($4000$ or $8000$).}\label{tab1} \end{table} Table~\ref{tab1} shows the results of the comparison with $11$ test functions in terms of the number of evaluations and CPU time to achieve a small error. The first two test functions, Sin~1 and Sin~2, were used to test the SOO algorithm \cite{mun13}, and have the form $f(x) = (\sin(13x)\sin(27x) + 1) / 2$ and $f (x_1, x_2) = f (x_1) f (x_2)$ respectively. The form of the third function, Peaks, is given in Equation (16) and illustrated in Figure 2 of \citeauthor{mcd07}'s paper \citeyear{mcd07}. The rest of the test functions are common benchmarks in global optimization literature; \citeauthor{sur13} present detailed information about the functions \citeyear{sur13}. In the table, $\mathit{Time}\,(s)$ indicates CPU time in second and $\mathit{Error}$ is defined as \[ \mathit{Error}= \begin{cases} \abs{(f(x^*)-f(x^+))/f(x^*)} & \text{if }f(x^*)\neq0,\\ \abs{f(x^*)-f(x^+)} & \text{otherwise}.\\ \end{cases} \] In the table, $N = 2n$ is the number of function evaluations needed to achieve $\mathit{Error} < 10^{-4}$, where $n$ is the total number of divisions and is the one used as the main measure in the analyses in the previous sections. Here, $N$ is equal to $2n$ because of the adopted division process. Thus, the lower the value of $N$ becomes, the better the algorithm's performance is. We continued iterations until $4000$ function evaluations for all the functions with dimensionality less than $10$, and $8000$ for the function with dimensionality equal to $10$. As can be seen in Table~\ref{tab1}, the LOGO algorithm outperformed the other algorithms. The superior performance of the LOGO algorithm with the small number of function evaluations is attributable to its focusing on the promising area discovered during the search. Conversely, the SOO algorithm continues to search the global domain and tends to be similar to a uniform grid search. The BaMSOO algorithm also follows the tendency toward a grid search as it is based on the SOO algorithm. The BaMSOO algorithm chooses where to divide based on the SOO algorithm; however, it omits function evaluations when the upper coincidence bound estimated by the GP indicates that the evaluation is not likely to be beneficial. Although this mechanism of the BaMSOO algorithm seems to be beneficial to reduce the number of function evaluations in some cases, it has two serious disadvantages. The first disadvantage is computational time due to the use of GP. Notice that it requires $O(N^3)$ every time to re-compute the upper confidence bound.% \footnote{Although there are several methods to mitigate its computational burden by approximation, the effect of the approximation on the performance of the BaMSOO algorithm is unknown and left to a future work.} A more serious disadvantage is the possibility of not determining a solution at all. From Table~\ref{tab1}, we can see that BaMSOO improves the performance of SOO in $7/11$ cases; however, it \emph{severely} degrades the performance in $4/11$ cases. Moreover, not only may BaMSOO reduce the performance but also it may not guarantee convergence even in the limit \emph{in practice}. The BaMSOO algorithm reduces the number of function evaluations by relying on the estimation of the upper confidence bound. However, the estimation can be wrong, and if it is wrong, it may never explore the region where the global optimizer exists. Notice that these limitations are not unique to BaMSOO but also apply to many GP-based optimization methods. In terms of the first limitation (computational time), BaMSOO is a significant improvement when compared to other traditional GP-based optimization methods \cite{wan14}. Although the LOGO algorithm has a bias toward local search, it maintains a strong theoretical guarantee, similar to the SOO algorithm, as proven in the previous sections. In terms of theoretical guarantee, the SOO algorithm and the LOGO algorithm share a similar rate on the loss bound and base their analyses on the same set of assumptions that hold in practice. On the other hand, the BaMSOO algorithm has a worse rate on the loss bound (an asymptotic loss of the order $n^{-(1-\epsilon)/d})$ and its bound only applies to a restricted class of a metric $\ell$ (the Euclidean norm to the power $\alpha = \{1,2\}$). It also requires several additional assumptions to guarantee the bound. Some of the additional assumptions would be impractical, particularly the assumption of the objective function being always well-captured by the GP with a chosen kernel and hyperparameters. As discussed above, this assumption would cause BaMSOO to not only lose the loss bound but also the consistency guarantee (i.e., convergence in the limit) in practice. Figure~\ref{fig5} presents the performance comparison for each number of function evaluations and Figure~\ref{fig6} plots the corresponding computational time. In both figures, a lower plotted value along the vertical axis indicates improved algorithm performance. For SA and GA, each figure shows the mean over 10 runs. We report the mean of the standard deviation over time in the following. For SA, it was 1.19 (Sin 1), 1.32 (Sin 2), 0.854 (Peaks), 0.077 (Branin), 1.06 (Rosenbrock 2), 0.956 (Hartman 3), 0.412 (Shekel 5), 0.721 (Shekel 7), 1.38 (Shekel 10), 0.520 (Hartman 6), and 0.489 (Rosenbrock 10). For GA, it was 0.921 (Sin 1), 0.399 (Sin 2), 0.526 (Peaks), 0.045 (Branin), 1.27 (Rosenbrock 2), 0.493 (Hartman 3), 0.216 (Shekel 5), 0.242 (Shekel 7), 1.19 (Shekel 10), 0.994 (Hartman 6), and 0.181 (Rosenbrock 10). \begin{figure}[t] \centering \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5a} \caption{Sin 1} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5b} \caption{Sin 2} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5c} \caption{Peaks} \end{subfigure} \vskip3mm \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5d} \caption{Branin} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5e} \caption{Rosenbrock 2} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5f} \caption{Hartman 3} \end{subfigure} \vskip3mm \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5g} \caption{Shekel 5} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5h} \caption{Shekel 7} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5i} \caption{Shekel 10} \end{subfigure} \vskip3mm \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5j} \caption{Hartman 6} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig5k} \caption{Rosenbrock 10} \end{subfigure}\hfill \rule{0.32\textwidth}{0pt} \caption{Performance comparison: the number of evaluations $N$ vs.\ the log error computed as $\log_{10}\abs{f(x^*)-f(x^+)}$. $f(x^*)$ indicates the true optimal value of the objective function and $f(x^+)$ is the best value determined by each algorithm.}\label{fig5} \end{figure} \begin{figure}[t] \centering \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6a} \caption{Sin 1} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6b} \caption{Sin 2} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6c} \caption{Peaks} \end{subfigure} \vskip3mm \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6d} \caption{Branin} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6e} \caption{Rosenbrock 2} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6f} \caption{Hartman 3} \end{subfigure} \vskip3mm \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6g} \caption{Shekel 5} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6h} \caption{Shekel 7} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6i} \caption{Shekel 10} \end{subfigure} \vskip3mm \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6j} \caption{Hartman 6} \end{subfigure}\hfill \begin{subfigure}[b]{0.32\textwidth} \includegraphics[width=\textwidth]{fig6k} \caption{Rosenbrock 10} \end{subfigure}\hfill \rule{0.32\textwidth}{0pt} \caption{CPU time comparison: CPU time required to achieve the performance indicated in Figure~\ref{fig5}}\label{fig6} \end{figure} As illustrated in Figure~\ref{fig5}, the LOGO algorithm generally delivered improved performance compared to the other algorithms. A particularly impressive result for the LOGO algorithm was its robustness for the more challenging functions, Shekel~10 and Rosenbrok~10. The function Shekel~$m$ has $m$ local optimizers and the slope of the surface generally becomes larger as $m$ increases. Therefore, Shekel~10 and Rosenbrok~10, which have $10$-dimensionality, are generally more difficult functions when compared with the others in our experiment. Indeed, only the LOGO algorithm achieved acceptable performance on these. From Figure~\ref{fig6}, we can see that the LOGO algorithm and the SOO algorithm were fast. The LOGO algorithm was often marginally slower than the SOO algorithm owing to the additional computation required to maintain the supersets. The reason why the BaMSOO algorithm required a large computational cost at some horizontal axis points is that it continued skipping to conduct the function evaluations (because the evaluations were judged to be not beneficial based on GP). This is an effective mechanism of BaMSOO to avoid wasteful function evaluations; however, one must be careful to make sure that the function evaluations are costly, relative to this mechanism. In summary, compared to the BaMSOO algorithm, the LOGO algorithm was faster and considerably simpler (in both implementation and parameter selection) and had stronger theoretical bases while delivering superior performance in the experiments. When compared with the SOO algorithm, the LOGO algorithm decreased the theoretical convergence rate in the worst case analysis, but exhibited significant improvements in the experiments. \begin{figure} \centering \hair2pt \begin{subfigure}[b]{0.49\textwidth} \includegraphics[width=\textwidth]{fig7b} \caption{ Scalability: a 1000-dimensional function} \label{fig7a} \end{subfigure} \labellist \pinlabel \rotatebox{90}{\scriptsize Log Distance to Optimal \normalsize} [r] at 12 180 \pinlabel {$N$} [t] at 270 25 \endlabellist \begin{subfigure}[b]{0.49\textwidth} \includegraphics[width=\textwidth]{fig7} \caption{Sensitivity to local bias parameter $w$ }\label{fig7b} \end{subfigure} \caption{On the current possible limitations of LOGO}\label{fig7} \end{figure} Now that we have confirmed the advantages of the LOGO algorithm, we discuss its possible limitations: scalability and parameter sensitivity. The scalability for high dimensions is a challenge for non-convex optimization in general as the search space grows exponentially in space. However, we may achieve the scalability by leveraging additional structures of the objective function that are present for some applications. For example, \citeauthor{kawaguchi2016deep} \citeyear{kawaguchi2016deep} showed an instance of deep learning models, in which the objective function has such an additional structure: the nonexistence of poor local minima. As an illustration, we combine LOGO with a random embedding method, REMBO \cite{wang2013bayesian}, to account for another structure: a low effective dimensionality. In Figure \ref{fig7} (a), we report the algorithms' performances for a 1000 dimensional function: Sin 2 embedded in 1000 dimensions in the same manner described in Section 4.1 in the previous study \cite{wang2013bayesian}. Another possible limitation of LOGO is the sensitivity of its performance to the free parameter $w$. Even though we provided theoretical analysis and insight on the effect of the parameter value in the previous section, it is yet unclear how to set $w$ in a principle manner. We illustrate this current limitation in Figure~\ref{fig7} (b). The result labeled with ``adaptive w'' indicates the result with the fixed adaptive mechanisms of \(w\) that we use in all the other experiments except ones in Figure 7 (b) and 8. In the illustration, we use the Branin function because the experiment conducted with it clearly illustrated the limitation. As can be seen in the figure, the performance in the early stage is always improved as $w$ increases because the algorithm finds a local optimum faster with higher $w$. However, if $w$ is too large, such as $w = 20$ in the figure, the algorithm gets stuck at the local optimum for a long time. Thus, the best value (or sweet spot) exists between too large and too small values of $w$. In the results of this experiment, it can be seen that the choice of $w = 2$ is the best, which finds the global optima with high precision within only $200$ function evaluations. However, this limitation would not be a serious problem in practice for the following four reasons. First, a similar limitation exists, to the best of our knowledge, for any algorithms that are successfully used in practice (e.g., simulated annealing, genetic algorithm, swarm-based optimization, the DIRECT algorithm, and Bayesian optimization). Second, unlike any other previous algorithm, the finite-time loss bound always applies even for a bad choice of $w$. Third, we demonstrated in the previous experiments that a very simple adaptive rule may suffice to produce a good result. Also, future work may further mitigate this limitation by developing different methods to adaptively determine the value of $w$. Also, another possibility would be to conduct optimization over \(w\) with a cheaper surrogate model. Finally, the limitation may not apply to some of the target objective functions at all. For the fourth and final reason, recall that we speculated in the algorithm's analysis that increasing $w$ would always have beneficial effects in some problems, as illustrated in Figure~\ref{fig4}. Clearly, any problems within the scope of local optimization fall into this category. In Figure~\ref{fig8}, we show a rather unobvious instance of such problems, and thus an example, for which the limitation of the parameter sensitivity does not apply. As can be seen in the diagram on the left in Figure~\ref{fig8}, this test function has many local optima, only one of which is the global optimum. Nevertheless, as in the diagram on the right, the performance of the LOGO algorithm improves as $w$ increases, with no harmful effect. \begin{figure}[ht] \centering \small\hair2pt \labellist \pinlabel $f$ [r] at -200 600 \pinlabel $x_1$ [tr] at 0 300 \pinlabel $x_2$ [t] at 400 330 \endlabellist \includegraphics[width=0.47\textwidth]{fig8a}\hfill \small\hair2pt \labellist \pinlabel \rotatebox{90}{Log Distance to Optimal} [r] at -160 420 \pinlabel {Local Bias Parameter: $w$} [t] at 230 142 \endlabellist \includegraphics[width=0.44\textwidth]{fig8b} \vskip3mm \caption{An example of problems where increasing $w$ is always better. The diagram on the left shows the objective function, and the diagram on the right presents the performance at $N = 50$ and $100$ for each $w$.}\label{fig8} \end{figure} \section{Planning with LOGO via Policy Search}\label{sec3} We now apply the LOGO algorithm to planning, which is an important area in the field of AI. The goal of our planning problem is to find an action sequence that maximizes the total return over the infinite discounted horizon or finite horizon (unlike classical planning problem, we do not consider constraints that specify the \emph{goal} state). In this paper, we discuss the formulations for the case of the infinite discounted horizon, but all the arguments are applicable to the case of a finite horizon with straightforward modifications. We consider the case where the state/action space is continuous, the planning horizon is long, and the transition and reward functions are known and deterministic. The planning problem can be formulated as follows. Let $S\subset\mathbb{R}^{D_S}$ be a set of states, $A\subset\mathbb{R}^{D_A}$ be a set of actions, $T\colon\mathbb{R}^{D_S} \to \mathbb{R}^{D_S}$ be a transition function, $R\colon\mathbb{R}^{D_S}\times\mathbb{R}^{D_A}\to\mathbb{R}$ be a return or reward function, and $\gamma\le 1$ be a discount factor. A planner considers to take an action $a \in A$ in a state $s \in S$, which triggers a transition to another state based on the transition function $T$, while receiving a return based on the reward function $R$. The discount factor $\gamma$ discounts the future rewards to fulfill either or both of the following two roles: accounting for the relative importance of immediate rewards compared to future rewards, and obviating the need to think ahead toward the infinite horizon. An action sequence can be represented by a policy $\pi$ that maps the state space to the action space: $\pi\colon\mathbb{R}^{D_S} \to \mathbb{R}^{D_A}$. The value of an action sequence or a policy $\pi$, $V^\pi$, is the sum of the rewards over the infinite discounted horizon, which is \[ V^\pi(s_0)=\sum_{j=0}^\infty\gamma^t R(s_j,\pi(s_j)). \] The value of a policy can be also written with a recursive form as \begin{equation}\label{eq3} V^\pi(s)=R(s,\pi(s)) + \gamma V^\pi\bigl(T(s,\pi(s))\bigr). \end{equation} Here, we are interested in finding the optimal policy $\pi^*$. In the dynamic programing approach, we can compute the optimal policy, by solving the following Bellman's optimality equation: \begin{equation}\label{eq4} V^*(s)=\max_a R(s,a) + \gamma V^*(T(s,a)) \end{equation} where $V^*$ is the value of the optimal policy. In Equation~\eqref{eq4}, the optimal policy $\pi^*$ is the set of the actions defined by the max. A major problem with this approach is that the efficiency of the computation depends on the size of the state space. In a real-world application, the state space is usually very large or continuous, which often makes it impractical to solve Equation~\eqref{eq4}. A successful approach to avoid the state size dependency is to focus only on the state space that is reachable from the current state within the planning time horizon. In this way, even with an infinitely large state space, a planner only needs to consider a finitely sized subset of the space. This approach is called \emph{local planning}. Unlike local optimization vs.\ global optimization, the optimal solution of local planning is indeed globally optimal, given the initial state. It is called local planning because it does not cover all the states and its solution changes for different initial states. Accordingly, as the initial state changes, a planner may need to conduct re-planning. A natural way to solve local planning is to use tree search methods, which construct a tree rooted in an initial state toward the future possible states in the depth of the planning horizon. This tree search can be conducted using any traditional search method, including both uninformed search (e.g., breadth-first and depth-first search) and informed (heuristic) search (e.g., $\rm A^*$ search). Also, recent studies have developed several tree-based algorithms that are specialized to local planning. Among those, the SOO algorithm, the direct predecessor of the LOGO algorithm, was applied to local planning with the tree search approach \cite{bus13}. Most of the new algorithms, for example, HOLOP \cite{bub10,wei12}, operate with stochastic transition functions. However efficient these proposed algorithms are, the search space in the tree search approach grows exponentially in the planning time horizon, $H$. Therefore, local planning with the tree search approach would not work well with a very long time horizon. In some applications, a small $H$ is justified, but in other applications, it is not. If an application problem requires a long time tradeoff between immediate and future rewards, then the tree search approach would be impractical. Here, we are motivated to solve such a real-world application, and therefore need another approach. In this paper, we consider policy search \cite{dei13} as an effective alternative to solve the planning problem with continuous state/action space and with a long time horizon. Policy search is a form of local planning. Thus, like the tree search approach, it operates even with infinitely large or continuous state space. In addition, unlike the tree search approach, policy search significantly reduces the search space by naturally integrating the domain-specific expert knowledge into the structure of the policy. More concretely, the search space of policy search is a set of policies $\{\pi_x : x \in \Omega\}$, which are parameterized by a vector $x$ in $\Omega\subset \mathbb{R}^D$. Therefore, the search space is no longer dependent on the planning time horizon $H$, the state space $S$, nor the action space $A$, but only on the parameter space $\Omega$. Here, parameter space $\Omega$ can be determined by expert knowledge, which can significantly reduce the search space. We use the regret $r_m$ as the measure of the policy search algorithm's performance: \[ r_m = V^{\pi_x^*}(s_0) - V^{\pi_x^+(m)}(s_0) \] where $\pi_x^*$ is the optimal policy in the given set of policies $\{\pi_x : x \in \Omega\}$, and $\pi_x^+(m)$ is the best policy found by an algorithm after the $m$ steps of planning. An evaluation of each policy takes $mH$ steps if we consider a fixed planning horizon $H$. Here, $\pi_x^*$ may differ from the optimal policy $\pi^*$ when $\pi^*$ is not covered in the set $\{\pi_x : x \in\Omega\}$. The policy search approach is usually adopted with gradient methods \cite{bax01,kob13,wei14}. While a gradient method is fast, it converges to local optima \cite{sut99}. Further, it has been observed that it may result in a mere random walk when large plateaus exist in the surface of the policy space \cite{hei08}. Clearly, these problems can be resolved using global optimization methods at the cost of scalability \cite{bro09,aza14}. Unlike previous policy search methods, our method guarantees finite-time regret bounds w.r.t.\ global optima in $\{\pi_x : x \in\Omega\}$ without strong additional assumption, and provides a practically useful convergence speed. \subsection{LOGO-OP Algorithm: Leverage (Unknown) Smoothness in Both Policy Space and Planning Horizon}\label{sec3.1} In this section, we present a simple modification of the LOGO algorithm to leverage not only the unknown smoothness in policy space but also the known smoothness over the planning horizon. The former is accomplished by the direct application of the LOGO algorithm to policy search, and the latter is what the modification in this section aims to do without losing the advantage of the original LOGO algorithm. We call the modified version, Locally Oriented Global Optimization with Optimism in Planning horizon (LOGO-OP). As a result of this modification, we add a new free parameter $L$. The pseudocode for the LOGO-OP algorithm is provided in Algorithm~\ref{alg2}. By comparing Algorithms~\ref{alg1} and \ref{alg2}, it can be seen that the LOGO-OP algorithm functions in the same manner as the LOGO algorithm, except for line 15 (the function evaluation or, equivalently, the policy evaluation in the policy search) and line 20. Notice that the LOGO algorithm is directly applicable to the policy search by considering $V$ to be $f$ in Algorithm~\ref{alg1}. While the LOGO algorithm does not assume the structure of the function $f$, the LOGO-OP algorithm functions with and exploits the given structure of the value function $V$ (i.e., MDP model). The algorithm functions as follows. The policy evaluation is performed for each policy $\pi_x$ with a parameter $x$ specified by each of the two new hyperrectangles (from line 15-1 to \hbox{15-11}). Given the initial condition $s_0 \in S$, the transition function $T$, the reward function~$R$, a~discount factor $\gamma \le 1$, and the policy $\pi_x$, the algorithm computes the value of the policy as in Equation~\eqref{eq3} (from line 15-2 to line 15-10, except line 15-6). The main modification appears in line 15-6 where the algorithm leverages the known smoothness over the planning horizon. Remember that \emph{the unknown smoothness in policy space} (or input space $x$) is specified as $f(x^*)-f(x)\le \ell(x,x^*)$ (from Assumption~\ref{ass1}) and thus it infers the upper bound of the value of a policy \emph{that is not yet evaluated but similar (close in policy space w.r.t.\ $\ell$) to already evaluated polices}. Conversely, \emph{the known smoothness over the planning horizon} renders the upper bound on the value of a policy \emph{while the particular policy is being evaluated}. That is, the known smoothness over the planning horizon can be written as \[ \sum_{j=0}^\infty \gamma^j R(s_j,\pi_x(s_j)) - \sum_{j=0}^t \gamma^j R(s_j,\pi_x(s_j))\le \frac{\gamma^{t+1}}{1-\gamma}R_\mathit{max} \] where $0\le t\le\infty$ is a arbitrary point in the planning horizon as in line 15-3 and $R_\mathit{max}$ is the maximum reward. This known smoothness is due to the definition of $R_\mathit{max}$ and the sum of a geometric series. In the case of the finite horizon with $H$, we have the same formula with $(\gamma^t / (1-\gamma))R_\mathit{max}$ being replaced by $(H-t)R_\mathit{max}$. In line 15-6, unlike the original LOGO algorithm, the LOGO-OP algorithm terminates the evaluation of a policy when the continuation of evaluating the policy is judged to be a misuse of the computational resources based on the known smoothness over the planning horizon. Concretely, it terminates the evaluation of a policy when the upper bound of the value of the policy becomes less than $(V^+ - L)$, where $V^+$ is the value of the best policy found thus far and $L$ is the algorithm's parameter. When the upper bound of the value of policy becomes less than $V^+$, the planner can know that the policy is not the best policy. Thus, it is tempting to simply terminate the policy evaluation with this criterion. However, the essence of the LOGO algorithm is the utilization of the unknown smoothness embedded in the surface of the value function in the policy space. In other words, the algorithm makes use of the result of each policy evaluation, whether the policy is the best one or not. Any interruption of the policy evaluation changes the shape of the surface of the value function, which interferes with the mechanism of the LOGO algorithm. Nevertheless, the some degree of the interruption is likely to be beneficial since our goal is to find the optimal policy instead of surface analysis. \begin{algorithm}[t] \caption{LOGO-OP algorithm}\label{alg2} \expandafter\patchcmd\csname\string\algorithmic\endcsname {\labelwidth 1.2em}{\labelwidth6mm}{}{} \begin{algorithmic}[1] \makeatletter \addtocounter{ALG@line}{-1} \makeatother \State {\bf Inputs (problem):} the initial condition $s_0 \in S$, the transition function $T$, the reward function $R$, a discount factor $\gamma \le 1$ with convergence criteria (or finite horizon $H$), the policy space $\pi_x: x \in \Omega\subset \mathbb{R}^D$. \State {\bf Inputs (parameter):} the search depth function $h_\mathit{max}\colon\mathbb{Z}^+\to[1,\infty)$, the local weight $w\in\mathbb{Z}^+$, stopping condition, the maximum reward $R_\mathit{max}$, a parameter $L$. \let\algln\alglinenumber \algrenewcommand{\alglinenumber}[1]{\footnotesize 2--5:} \State lines 2--5 are exactly the same as lines 2--5 in Algorithm~\ref{alg1} \makeatletter \addtocounter{ALG@line}{3} \makeatother \algrenewcommand{\alglinenumber}[1]{\footnotesize #1:} \State Adds the initial hyperrectangle $\Omega'$ to the set: $\psi_0\leftarrow\psi_0\cup\{\Omega'\}$ (i.e., $\omega_{0,0} = \Omega'$) \State Evaluate the value function $V$ at the center point of $\Omega'$, $c_{0,0}$: $\mathit{val}[\omega_{0,0}]\leftarrow V(c_{0,0})$, $V^+\leftarrow\mathit{val}[\omega_{0,0}]$ \For{iteration ${}= 1, 2, 3,\dots$} \State $\mathit{val}_\mathit{max}\leftarrow -\infty$, $h_\mathit{plus}\leftarrow h_\mathit{upper}$ \For{$k=0,1,2,\dots,\max(\floor{\min(h_\mathit{max}(n),h_\mathit{upper})/w},h_\mathit{plus})$} \State \underline{Select}\ a hyperrectangle to be divided: $(h,i)\in\arg\max_{h,i}\mathit{val}[\omega_{h,i}]$ for $h,i : \omega_{h,i}\in\Psi_k$ \If{$\mathit{val}[\omega_{h,i}]>\mathit{val}_\mathit{max}$} \State $\mathit{val}_\mathit{max}\leftarrow\mathit{val}[\omega_{h,i}]$, $h_\mathit{plus}\leftarrow 0$, $h_\mathit{upper}\leftarrow\max(h_\mathit{upper},h+1)$, $n\leftarrow n+1$ \State \underline{Divide}\ this hyperrectangle $\omega_{h,i}$ along the longest coordinate direction \Statex \qquad\qquad\quad\begin{tabular}[t]{l} - three smaller hyperrectangles are created $\rightarrow$ $\omega_\mathit{left}$, $\omega_\mathit{center}$, $\omega_\mathit{right}$\\ - $\mathit{val}[\omega_\mathit{center}]\leftarrow\mathit{val}[\omega_{h,i}]$ \end{tabular} \color{blue!30!black} \State \underline{Evaluate}\ the value function $V$ at the center points of the two new hyperrectangles: \newcounter{foo} \makeatletter \setcounter{foo}{\value{ALG@line}} \setcounter{ALG@line}{0} \algrenewcommand{\alglinenumber}[1]{\footnotesize\hskip1.3cm \llap{\arabic{foo}--\arabic{ALG@line}}:} \makeatother \For{each policy $\pi_x$ corresponding $c_{\omega_\mathit{left}}$ and $c_{\omega_\mathit{right}}$} \State $z_1\leftarrow0$, $z_2\leftarrow1$, $s\leftarrow s_0$ \For{$t=0,1,2,\dots$,} \State $z_1\leftarrow z_1+z_2R(s,\pi_x(s))$ \State $z_2\leftarrow \gamma z_2$, $s\leftarrow T(s,\pi_x(s))$ , \State {\bf if} $z_1 + (\gamma^{t+1} / (1-\gamma))R_\mathit{max} < (V^+ - L)$ {\bf then Exit loop} \State {\bf if} convergence criteria is met {\bf then Exit loop} \EndFor \State save $z_1$ as the value of the corresponding rectangle \State $\mathit{val}[\omega_\mathit{left}]\leftarrow z_1$ or $\mathit{val}[\omega_\mathit{right}]\leftarrow z_1$ \EndFor \State $V^+\leftarrow\max(V^+,\mathit{val}[\omega_\mathit{left}],\mathit{val}[\omega_\mathit{center}],\mathit{val}[\omega_\mathit{right}])$ \makeatletter \setcounter{ALG@line}{\value{foo}} \makeatother \algrenewcommand{\alglinenumber}[1]{\footnotesize #1:} \color{black} \State \underline{\smash{Group}}\ the new hyperrectangles into the set $h + 1$ and remove the original rectangle: \begin{center} $ \psi_{h+1}\leftarrow\psi_{h+1} \cup\{\omega_\mathit{center},\omega_\mathit{left},\omega_\mathit{right}\},\quad \psi_h\leftarrow\psi_h\setminus \omega_{h,i} $ \end{center} \EndIf \State {\bf if} stopping condition is met {\bf then Return} $(h,i)=\arg\max_{h,i}\mathit{val}[\omega_{h,i}]$ \EndFor \color{blue!30!black} \State {\bf for} all intervals $\omega$ with $\mathit{val}[\omega]<(V^+-L)$ {\bf do} $\mathit{val}[\omega]\leftarrow (V^+ - L)$ \color{black} \EndFor \end{algorithmic} \end{algorithm} The LOGO-OP algorithm uses $L$ to determine the degree of the interruption. Because $V^+$ is monotonically increasing along the execution, the value of a policy that is not fully evaluated owing to line 15-6 in early iterations tends to be greater than the value of a policy that is not fully evaluated in the later iterations. The algorithm resolves this problem in line 20 such that it is not biased to divide the interval evaluated in an early iteration. With smaller $L$, the LOGO-OP algorithm can stop the evaluation of a non-optimal policy earlier, at the cost of accuracy in the evaluation of the value function's surface. With larger $L$, the algorithm needs to spend more time on the evaluation of a non-optimal policy, but can obtain a more accurate estimate of the value function's surface. In the regret analysis, we show that a certain choice of $L$ ensures a tighter regret bound when compared to the direct application of the LOGO algorithm. \subsection{A Parallel Version of the LOGO-OP Algorithm}\label{sec3.2} The LOGO-OP algorithm presented in Algorithm~\ref{alg2} has four main procedures: \underline{Select}\ (line 11), \underline{Divide}\ (line 14), \underline{Evaluate}\ (line 15), and \underline{\smash{Group}}\ (line 16). A natural way to parallelize the algorithm is to decouple \underline{Select}\ from the other three procedures. That is, let the algorithm first \underline{Select}\ $z$ hyperrectangles to be divided, and then allocate the $z$ number of \underline{Divide}, \underline{Evaluate}, and \underline{\smash{Group}}\ to $z$ parallel workers. However, this natural parallelization has data dependency from one \underline{Select}\ to another \underline{Select}. In other words, the procedure of the next \underline{Select}\ cannot start before \underline{Divide}, \underline{Evaluate}, and \underline{\smash{Group}}\ for the previous \underline{Select}\ are finalized. As a result, the parallel overhead tends to be non-negligible. In addition, if \underline{Select}\ chooses less hyperrectangles than parallel workers, then the available resources of the parallel workers are wasted. Indeed, the latter problem was tackled by creating multiple initial rectangles in a recent parallelization study of the DIRECT algorithm \cite{he09}. While the use of multiple initial rectangles can certainly mitigate the problem, it still allows the occasional occurrence of the resource wastage, in addition to requiring the user to specify the arrangement of the initial rectangles. To solve these problems, we instead decouple the \underline{Evaluate}\ procedure from the other three procedures and allocate only the \underline{Evaluate}\ task to each parallel worker. We call the parallel version, the pLOGO-OP algorithm. The algorithm uses one master process to conduct \underline{Select}, \underline{Divide}, and \underline{\smash{Group}}\ operations and an arbitrary number of parallel workers to execute \underline{Evaluate}. The main idea is to temporarily use the artificial value assignment to the center point of a hyperrectangle in the master process, which is overwritten by the true value when the parallel worker finishes evaluating the center point. With this strategy, there is no data dependency and all the parallel workers are occupied with tasks almost all the time. In this paper, we use the center value of the original hyperrectangle before division as the temporary artificial value, but the artificial value may be computed using a more advanced method (e.g., methods in surface analysis) in future work. For the center point of the initial hyperrectangle, we simply assign the worst possible value (if we have no knowledge regarding the worst value, we can use $-\infty$). The master process keeps selecting new hyperrectangles unless all the parallel workers are occupied with tasks. This logic ensures that all the parallel workers always have tasks assigned by the master process, but the master process does not select too many hyperrectangles based on the artificial information. Note that this parallelization makes sense only when \underline{Evaluate}\ is the most time consuming procedure, and it is very likely true for policy evaluation. \subsection{Regret Analysis}\label{sec3.3} Under a certain condition, all the finite-loss bounds of the LOGO algorithm are directly translated to the regret bound of the LOGO-OP algorithm. The condition that must be met is that $(V^+ - L)$ is less than the center value of the optimal hyperinterval during the algorithm's execution. We state the regret bound more concretely below. For simplicity, we use the notion of a planning horizon $H$, which is the effective (non-negligible) planning horizon for LOGO in accordance with the discount factor, $\gamma$. Let $H'$ be the effective planning horizon of the LOGO-OP algorithm. Then, the planning horizon for LOGO-OP, $H'$, becomes smaller than that for LOGO, $H$, as the algorithm finds improved function values. This is because the LOGO-OP algorithm terminates each policy evaluation at line 15-6 when the upper bound on the policy value is determined to be lower than $(V^+ - L)$. \begin{corollary}\label{cor3} Let $H'\le H$ be the planning horizon used by the LOGO-OP algorithm at each policy evaluation. Let $V^+$ be the value of the best policy found by the algorithm at any iteration. Assume that the value function of the policy satisfies Assumptions~\ref{ass1} and \ref{assB1}. If $(V^+ - L)$ is maintained to be less than the center value of the optimal hyperinterval, then the algorithm holds the finite-time loss bound of Theorem~\ref{the2} with \[ n\ge\bbfloor{\frac{m}{2H'}}. \] \end{corollary} \begin{proof} As the policy search is just a special case of the optimization problem, it is trivial that the loss bound of Theorem~\ref{the2} holds for the LOGO algorithm when it is applied to policy search. Because every function evaluation takes $H$ steps in the planning horizon, we have $n\ge\floor{m/2H}$ in this case. For the LOGO-OP algorithm, only the effect that new parameter $L$ has in the loss analysis takes place in the proof of Lemma~\ref{lem2}. If $(V^+ - L)$ is maintained to be less than the center value of the optimal interval, then all the statements in the proof hold true for the LOGO-OP algorithm as well. Here, due the effect of $L$, function evaluation may take less than $H$ steps in the planning horizon. Therefore, we have the statement of this corollary. \end{proof} We can tighten the regret bound of the LOGO-OP algorithm by decreasing $L$, since the algorithm can then terminate evaluations of unpromising policies earlier, which means that the value of $H'$ in the bound is reduced. However, using a too small value of $L$ that violates the condition in Corollary~\ref{cor3} leads us to discard the theoretical guarantee. Even in that case, because the too small value of $L$ only results in a more global search, the consistency property, $\lim_{n\to\infty}r_n=0$, is still trivially maintained. On the other hand, if we set $L =\infty$, the LOGO-OP algorithm becomes equivalent to the direct application of the LOGO algorithm to policy search, and thus, we have the regret bound of Corollary~\ref{cor3} with $H' = H$. The pLOGO-OP algorithm also maintains the same regret bound with $n = n_p$ where $n_p$ counts the number of the total divisions that are devoted to the set of $\delta$-optimal hyperinterval $\psi_{kw+l}(l+1)^*$, where $(w - 1)\ge l\ge 0$. While non-parallel versions ensure the devotion to $\psi_{kw+l}(l+1)^*$, the parallelization makes it possible to conduct division on other hyperintervals. Thus, considering the worst case, the pLOGO-OP may not improve the bound in our proof procedure, although the parallelization is likely beneficial in practice. \subsection{Application Study on Nuclear Accident Management}\label{sec3.4} The management of the risk of potentially hazardous complex systems, such as nuclear power plants, is a major challenge in modern society. In this section, we apply the proposed method to accident management of nuclear power plants and demonstrate the potential utility and usage of our method in a real-world application. Our focus is on assessing the efficiency of containment venting as an accident management measure and on obtaining knowledge about its effective operational procedure (i.e., policy $\pi$). This problem requires planning with continuous state space and with a very long planning horizon ($H\ge 86400$), for which dynamic programming (e.g., value iteration), tree-based planning (e.g., $\rm A^*$ search and its variants) would not work well (dynamic programming suffers from the curse of dimensionality for the state space, and the search space of tree-based methods grows exponentially in the planning horizon). Containment venting is an operation that is used to maintain the integrity of the containment vessel and to mitigate accident consequences by releasing gases from the containment vessel to the atmosphere. In the accident at the Fukushima Daiichi nuclear power plant in 2011, the containment venting was activated as an essential accident management measure. As a result, in 2012, the United States Nuclear Regulatory Commission (USNRC) issued an order for $31$ nuclear power plants to install the containment vent system {\DeclareRobustCommand{\us}[2]{#2}\cite{usn13}}. Currently, many countries are considering the improvement of the containment venting system and its operational procedures \cite{oec14}. The difficulty of determining its actual benefit and effective operation comes from the fact that the containment venting also releases fission products (radioactive materials) into the atmosphere. In other words, the effective containment venting must trade off the future risk of containment failure against the immediate release of fission products (radioactive materials). In our experiments, we use the release amount of the main fission product compound, cesium iodide (CsI), as a measure of the effectiveness of the containment venting. In the nuclear accident management literature, an integrated physics simulator is used as the model of world dynamics or the transition function $T$ and the state space $S$. The simulator that we adopt in this paper is THALES2 (\underline{T}hermal \underline{H}ydraulics and radionuclide behavior \underline{A}nalysis of \underline{L}ight water reactor to \underline{E}stimate \underline{S}ource terms under severe accident conditions) \cite{ish02}. Thus, the transition function $T$ and the state space $S$ are fully specified by THALES2. The initial condition $s_0 \in S$ is designed to approximately simulate the accident at the Fukushima Daiichi nuclear power plant. In this experiment, we focus on a single initial condition with the deterministic simulator, the relaxation of which is discussed in the next section. The reward function $R$ is the negative of the amount of CsI being released in the atmosphere as a result of a state-action pair. We use the finite-time horizon $H = 86400$ seconds ($24$ hours), which is a traditional first phase time-window considered in risk analysis with nuclear power plant simulations (owing to the assumption that after $24$ hours, many highly uncertain human operations are expected). We use the following policy structure based on our engineering judgment. \[ \pi_x = \begin{cases} 1 & \text{if }((\mathit{FP}\le x_1) \cap (\mathit{Press}\ge x_2)) \cup (\mathit{Press} > 100490),\\ 0 & \text{otherwise}, \end{cases} \] where $\pi_x = 1$ indicates the implementation of the containment venting, $\mathit{FP}$ (g) represents the amount of CsI in the gas phase of the suppression chamber, and $\mathit{Press}$ (kgf/m$^2$) is the pressure of the suppression chamber. Here, the suppression chamber is the volume in the containment vessel that is connected to the atmosphere via the containment venting system. This policy structure reflects our engineering knowledge that the venting should be done while the fission products exist under a certain amount in the suppression chamber, but should not be operated before the pressure gets larger than a specific value. We consider $x_1 = [0, 3000]$ and $x_2 = [10810, 100490]$. We let $\pi_x = 1$ whenever the pressure exceeds $100490$\,kgf/m$^2$, since the containment failure is considered to probably occur after the pressure exceeds this point. The detail of the experimental setting is outlined in Appendix~A. We first compare the performance of various algorithms in this problem. For all the algorithms, we used the same parameter settings as in the benchmark tests in Section~\ref{sec2.4}. That is, we used $h_\mathit{max}(n) = w\sqrt{n} -w$ and a simple adaptive procedure for the parameter $w$ with $W = \{3, 4, 5, 6, 8, 30\}$. For the LOGO-OP algorithm and the pLOGO-OP algorithm, we blindly set $L = 1000$ (i.e., there is likely a better parameter setting for $L$). We used only eight parallel workers for the pLOGO-OP algorithm. \begin{figure} \centering \hair2pt \labellist \pinlabel \rotatebox{90}{CsI release by the Computed Policy (g)} [r] at 0 100 \pinlabel {Wall time (s)} [t] at 160 0 \endlabellist \includegraphics[width=0.65\textwidth]{fig9} \vskip5mm \caption{Performance of the computed policy (CsI release) vs.\ Wall Time.}\label{fig9} \end{figure} Figure~\ref{fig9} shows the result of the comparison with wall time $\le 12$ hours. The vertical axis is the total amount of CsI released into the atmosphere (g), which we want to minimize. Since we conducted containment venting whenever the pressure exceeded $100490$\,kgf/m$^2$, containment failure was prevented in all the simulation experiments. Thus, the lower the value along the vertical axis gets, the better the algorithm's performances is. As can be seen, the new algorithms performed well compared to the SOO algorithm. It is also clear that the two modified versions of the LOGO algorithm improved the performance of the original. For the LOGO-OP algorithm, the effect of $L$ on the computational efficiency becomes greater as the found best policy improves. Indeed, the LOGO algorithm required $10798$ seconds for ten policy evaluations and $52329$ seconds for $48$ evaluations. The LOGO-OP algorithm required $9297$ seconds for ten policy evaluations, and $44678$ seconds for $48$ evaluations. This data in conjunction with Figure~\ref{fig9} illustrates the property of the LOGO-OP algorithm that the policy evaluation becomes faster as the found best policy improves. For the pLOGO-OP algorithm, the number of function evaluations performed by the algorithm increased by a factor of approximately eight (the number of parallel workers) compared to the non-parallel versions. Notice that the parallel version tends to allocate the extra resources to the global search (as opposed to the local search). We can focus more on the local search by utilizing the previous results of the policy evaluations; however, the parallel version must initiate several policy evaluations without waiting for the previous evaluations, resulting in a tendency for global search. This tendency forced the improvement, in terms of reducing the amount of CsI, to be moderate relative to the number of policy evaluations in this particular experiment. However, such a tendency may have a more positive effect in different problems where increased global search is beneficial. The CPU time per policy evaluation varied significantly for different policies owing to the different phenomenon computed in the simulator. On the average, for the LOGO-OP algorithm, it took approximately $930$ seconds per policy evaluation. Now that we partially confirmed the validity of the pLOGO-OP algorithm, we attempt to use it to provide meaningful information to this application field. Based on the examination of the results in the above comparison, we narrowed the range of the parameter values as $x_1 = [0, 1.2]$ and $x_2 = [10330, 10910]$. After the computation with CPU time of $86400$ (s) and with eight workers for the parallelization, the pLOGO-OP algorithm found the policy with $x_1\approx 0.195$ (g) and $x_2\approx 10880$ (kgf/m$^2$). With the policy determined, containment failure was prevented and the total amount of CsI released into the atmosphere was limited to approximately $0.5$ (g) (approximately $0.002\%$ of the total CsI) in the $24$ hours after the initiation of the accident. This is a major improvement because this scenario with our experimental setting is considered to result in a containment failure or at best, in a large amount of CsI release, more than $2000$ (g) (about $10\%$ of total CsI) in our setting. The computational cost of CPU time of $86400$ (s) is likely acceptable in the application field. In terms of computational cost, we must consider two factors: the offline computation and the variation of scenarios. The computational cost with CPU time of $86400$ (s) for a phenomenon that requires $86400$ (s) is not acceptable for online computation (i.e., determining a satisfactory policy while the accident is progressing). However, such computational cost is likely acceptable if we consider preparing acceptable policies for various scenarios in an offline manner (i.e., determining satisfactory polices before the accident). Such information regarding these polices can be utilized during an accident by first identifying the accident scenario with heuristics or machine learning methods \cite{par10}. For offline preparation, we must determine policies for various major scenarios and thus if each computation takes, for example, one month, it may not be acceptable. \begin{figure} \centering \small\hair2pt \labellist \pinlabel \rotatebox{90}{Venting (-) / CsI (g)} [r] at -10 80 \pinlabel {Time along Accident Progression (s)} [t] at 230 -5 \endlabellist \includegraphics[width=0.8\textwidth]{fig10} \vskip5mm \caption{Action sequence generated by found policy and CsI release}\label{fig10} \end{figure} Note that the policy found by our method is both novel and nontrivial in the literature, and yet worked very well. Accordingly, we explain why the policy performed as well as it did. Figure~\ref{fig10} shows the action sequence generated by the policy found and the amount of CsI (g) released versus accident progression time (s). We analyze the action sequence by dividing it into six phases, as indicated in Figure~\ref{fig10}, with the six numbers inside the parentheses. In the first phase (1), the venting is conducted intermittently in order to keep the pressure around $x_2\approx 10880$ (kgf/m$^2$). In this phase, no fission product has yet been released from the nuclear fuels. Reducing the pressure and the heat should be done preferably without releasing fission products, and the actions in this phase accomplish this. One may wonder why the venting should be done intermittently, instead of continuing to conduct venting to reduce the pressure as much as possible, which can be done without the release of fission products only in this phase. This is because reducing the pressure too much leads to a large difference between the pressures in the suppression chamber and the reactor pressure vessel, which in turn results in a large mass flow and fission product transportation from the reactor pressure vessel to the suppression chamber (see Figure~\ref{fig11} in Appendix~A for information about the mass flow paths). The increase in the amount of fission products in the suppression chamber will likely result in a large release of fission products into the atmosphere when venting is conducted. Therefore, this specific value of $x_2$ that generates the intermittent venting works well in the first phase. In the second phase (2), containment venting is executed all the time since the pressure in the suppression chamber increases rapidly in this phase (due to the operation of depressurizing the reactor pressure vessel via the SRV line), and thus, the criterion ($\mathit{Press}\ge x_2$) in the policy is satisfied all the time from this point. In the beginning of the third phase (3), the amount of CsI in the suppression chamber exceeds $x_1\approx 0.195$ (g) and thereby no venting is conducted. In the fourth phase (4), the pressure reaches $100490$ (kgf/m$^2$) and containment venting is intermittently done in order to keep the pressure under the point to avoid catastrophic containment failure. In the fifth phase (5), the containment vent is kept open because the amount of CsI in the gas phase of the suppression chamber decreases to below $x_1$ (due to the phenomenon illustrated in Figure~\ref{fig12} in Appendix~A). This continuous containment venting decreases the pressure such that no venting is required in terms of the pressure in the final phase (6), where venting is not conducted also because the amount of CsI becomes larger than $x_1$. Thus, it is clear that the policy found by this AI-related method also has a basis in terms of physical phenomenon. In addition, the generated action sequence is likely not simple enough for an engineer to discover with several sensitivity analyses. In particular, not only did our method solve the known tradeoff between the immediate CsI release and the risk of future containment failure, the method also discovered the existence of a new tradeoff between the immediate reduction of the pressure without CsI release and future increase in the mass flow. Although there is no consensus as to how to operate the containment venting system at the moment, the tendency is to use it only when the pressure exceeds a certain point in order to prevent immediate sever damage of containment vessel, which corresponds only to the fourth phase (4) in Figure~\ref{fig10}. In our experiment, such a myopic operation resulted in containment failure, or a significantly large amount of CsI being released into the atmosphere (at least more than $4800$ (g)). In summary, we successfully applied the proposed method to investigate the containment venting policy in nuclear power plant accidents. As a preliminary application study, several topics are left to future work. From a theoretical viewpoint, future work should consider a way to mitigate the simulation bias due to model error and model uncertainty. For the model error, the robotics community is already cognizant that a small error in a simulator can result in poor performance of the derived policy (i.e., simulation bias) \cite{kob13}. We can mitigate this problem by adding a small noise to the model, since the noise works as \emph{regularization} to prevent over-fitting as demonstrated by \citeA{atk98}. For the model uncertainty, recent studies in the field of nuclear accident analysis provide possible directions for the treatment of uncertainty in accident phenomena \cite{zhe15} and accident scenarios \cite{kaw12}. As a result of either or both of these countermeasures, the objective function becomes stochastic, and thereby we may first expand the pLOGO-OP algorithm to stochastic case. On the other hand, from the phenomenological point of view, future work should consider other fission products as well as CsI. Such fission products include, but are not limited to, Xe, Cs, I, Te, Sr, and Ru. In particular, a noble gas element, such as Xe, can be a major concern in an accident (it tends to be released a lot and is easily diffused into the atmosphere), but its property is different from CsI (its half-life is much smaller). Thus, if Xe is identified as a major concern, one may consider a significantly different policy from ours (considering its half-life, one may delay conducting the containment venting). \section{Conclusions}\label{sec4} In this paper, we proposed the LOGO algorithm, the global optimization algorithm that is designed to operate well in practice while maintaining a finite-loss bound with no strong additional assumption. The analysis of the LOGO algorithm generalized previous finite-loss bound analysis. Importantly, the analysis also provided several insights regarding practical usage of this type of algorithm by showing the relationship among the loss bound, the division strategy, and the algorithm's parameters. We applied the LOGO algorithm to an AI planning problem with the policy search framework, and showed that the performance of the algorithm can be improved by leveraging not only the unknown smoothness in policy space, but also the known smoothness in the planning horizon. As our study is motivated to solve real-world engineering applications, we also discussed a parallelization design that utilizes the property of AI planning in order to minimize the overhead. The resulting algorithm, the pLOGO-OP algorithm, was successfully applied to a complex engineering problem, namely, policy derivation for nuclear accident management. Aside from the planning problem that we considered, the LOGO algorithm can be also used, for example, to optimize parameters of other algorithms (i.e., algorithm configuration). In the AI community, the algorithm configuration problem has been addressed by several methods, including a genetic algorithm \cite{ans09}, discrete optimization with convergence guarantee in the limit \cite{hut09}, the racing approach originated from the machine learning community (Hoeffding Races) \cite{bir10}, model-based optimization with convergence guarantee in the limit \cite{hut11}, a simultaneous use of several randomized local optimization methods \cite{gyo11}, and Bayesian optimization \cite{sno12}. Compared to the previous parameter tuning methods, the LOGO algorithm itself is limited to optimizing continuous deterministic functions. To apply it to stochastic functions, a future work would modify the LOGO algorithm as was done for the SOO algorithm in a previous study \cite{val13}. To consider categorical and/or discrete parameters in addition to continuous parameters, a possibility could be to use the LOGO algorithm as a subroutine to deal with the continuous variables in one of the previous methods. The promising results presented in this paper suggest several interesting directions for future research. An important direction is to leverage additional assumptions. Since LOGO is based on a weak set of assumptions, it would be natural to use LOGO as a main subroutine but add other mechanisms to account for additional assumptions. As a example, we illustrated that LOGO would be able to scale up for a higher dimension with additional assumptions in Section \ref{sec2.4}. Another possibility is to add a GP assumption based on the idea presented in a recent paper \cite{kawaguchiNIPS2015}. Future work also would design an autonomous agent by integrating our planning algorithm with a learning/exploration algorithm \cite{kawaguchiAAAI2016}. One remaining challenge of LOGO is to derive a series of methods that adaptively determine the algorithm's free parameter $w$. As illustrated in our experiment, the achievement in this topic not only mitigates the problem of parameter sensitivity, but also would improve the algorithm's performance. \acks{This work was carried out while the first author was at the Japan Atomic Energy Agency. The authors would like to thank Dr.~Hiroto Itoh and Mr.~Jun Ishikawa at JAEA for several discussions on the related topics. The authors would like to thank Mr.~Lawson Wong at MIT for his insightful comments. The authors would like to thank anonymous reviewers for their insightful and constructive comments. }
0b1b62cc2ce7af05de81d6c72cd5c7337bf0bbe9
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The optimal liquidation problem of large trades has been studied extensively in the micro-structure literature. Two major sources of risk faced by large traders are: (i) inventory risk arising from uncertainty in asset value; and (ii) transaction costs arising from market friction. In a frictionless and competitive market, an asset can be traded with any amount at any rate without affecting the market price of the asset. The optimal liquidation problem then becomes an optimal stopping problem. In an incomplete market, the optimal liquidation problem involves delicate market-micro-structural issues. The impacts of transaction costs on optimal liquidation and optimal portfolio selection have been studied via various mechanisms \cite{AC1,RA1,Almgren3, RA2, DN90, SS94}. In this paper, we adopt a simple, but practical market impact model to study some aspects of the optimal liquidation problem. The model is phenomenological and not directly based on the fine details of micro-structure though it may primarily be related to some literature on market micro-structure. Following Almgren and Chriss \cite{AC1}, we decompose the price impact into temporary price impact and permanent price impact. Temporary impact refers to temporary imbalance in supply/demand caused by trading. It disappears immediately when trading activities cease. Permanent impact means changes in the ``equilibrium'' price due to trading, which lasts at least for the whole process of liquidation. We generalize the work of Almgren and Chriss \cite{AC1} to a more realistic and flexible framework in which multiple venues are available for the trader to submit his/her trades. We mainly consider short-term liquidation problem for a large trader who experiences temporary and permanent market impact. Two broad classes of problems are addressed in this paper which we believe are representative. The first one is the case with constant volatility. This assumption considerably simplifies the problem and allows us to exhibit the essential features of liquidation across multiple venues without losing ourselves in complexities. The second one is the case when volatility varies randomly throughout the trading horizon. In this case, we present a ``pure'' stochastic volatility approach, in which the volatility is modeled as an It\^{o} process driven by a Brownian motion that has a component independent of the Brownian motion driving the asset price. In comparison with Almgren's work in \cite{RA2}, we mainly focus on a special class of volatility model, the time-scale volatility model proposed in \cite{Fouque00, Fouque11}: $$ d\nu_t=\epsilon(m-\nu_t)dt+\xi\sqrt{\epsilon} dW_t, $$ and work in the regime $ \epsilon\ll 1$. The separation of time-scales provides an efficient way to identify and track the effect of stochastic volatility, which is desirable from the practical perspective. We also generalize our basic model to include the usage of limit orders. Different from market orders, limit orders are designed to give investors more flexibility over the buying and selling prices of their trades. The most unfavorable feature of limit orders might be the execution risk, i.e., a successful execution is not guaranteed. Cartea and Jaimungal \cite{CJ15} have constructed optimal strategies for this problem under temporary-impact-only assumption. In their model, and also in ours, investors are allowed to trade both via limit sell orders, whose execution are uncertain, and via market sell orders, whose execution are immediate but costly. Limit orders in their model are allowed to be continuously submitted. Market orders, however, are only for one unit, and are executed at a sequence of increasing stopping times. This model works well for small- and mid-size orders. However, when it comes to large orders, execution cannot be guaranteed. Furthermore, the optimal stopping time setting makes the problem difficult to be solved. To ensure execution and track the trace of liquidating strategy, we model the market orders strategy as a continuous control problem. When other market participants' buy market orders arrive, limit sell orders will be used instead to take advantage of the price gap. As mentioned in Rudloff et al. \cite{Rudloff}, a major reason for developing dynamic models instead of static ones is the fact that one can incorporate the flexibility of dynamic decisions to improve our objective function. Time-inconsistent criteria are generally not favorable to introduce in the study, since the associated policies are sub-optimal. The mean-variance criterion is popular for taking both return and risk into account. However, the mean-variance criterion may induce a potential problem of time-inconsistency, i.e., planned and implemented policies are different, and make the problem complicated. Hence, to take both return and risk into account, instead of adopting the mean-variance criterion, we are most interested in the mean-quadratic optimal agency liquidation strategies, as they are proved to be time-consistent in \cite{PA1, PA2}. In the following, we outline the major contributions of this paper. First, we obtain closed-form solutions for optimal liquidation strategy under the constant volatility approach for any level of risk aversion. This leads to an efficient frontier of optimal strategies. Each point on this frontier represents a strategy with the minimal level of cost for its level of risk-aversion. In addition, we also show that in the presence of multiple venues, institutional traders might hide their liquidating purposes and reduce the permanent market impacts through splitting their orders across multiple venues. When more trading venues spring up in financial market, the liquidity of the market enhances. This provides good suppliers of liquidity to the traders, and that the trader with more choices of venues to submit his/her trades may be willing to close his/her position as soon as possible. Second, we present a framework for the liquidation problem that is general enough to include the effect of stochastic volatility. Moreover, it is tractable and the parameters can be estimated efficiently on large datasets that are increasing available. Third, we study the effect of incorporating limit orders so as to detect market information. In general, for a pure liquidation problem, people do not include the usage of limit orders. But in our study of including limit orders, we find that, realizing the information carried by other market participants, traders will slow down their liquidating speed so as to get profit from market momentum. The remainder of this paper is organized as follows. Section 2 devotes to model building for the execution problem in the presence of multiple trading venues. The Profit \& Loss (P\&L) of trading and a mean-quadratic-variation criterion are proposed and discussed in this section. The solutions and numerical results for the constant volatility approximation are presented in Section 3. Extension of the results to the stochastic volatility approximation are then discussed in Section 4. An extension of our model to the incorporation of limit orders is studied in Section 5. The final section summarizes the results. \section{Problem Setup} In this section, we first present our liquidation problem in the presence of multiple trading venues based on the principle of no arbitrage. We then discuss the optimal liquidating strategies. \subsection{The Trader's Liquidation Problem} We consider an institutional trader. Beginning at time $ t=0$, he/she has a liquidation target of $Q$ shares, which must be completed by time $t=T$. The number of shares remaining to liquidate at time $t$ is the trajectory $ X_t$, with $ X_0=Q$ and $X_T=0$. Suppose there are $N$ distinct venues for the trader to submit his/her trades. The rate of liquidating in Venue $n$ ($n=1,2,\ldots, N$) is $ \theta^{(n)}_t$. Thus, $$ \theta^{(1)}_t+\cdots+\theta^{(N)}_t=-\frac{dX_t}{dt}. $$ For a liquidation program, $ Q>0$, we expect $ X_t\ge 0$ is decreasing and $ \theta^{(n)}_t\ge0$ (a buy program may be modeled similarly). Here, $ \theta^{(n)}_t$ ($n = 1, 2, \ldots, N$), are the choice variables or decision variables of the trader at time $t$ based on the observed information available at time $t$. In general, the trajectory $X_t$ depends on the price movements and market conditions that are discovered during trading, so it is a stochastic process. Consider a probability space $ (\Omega, \mathcal F, \mathcal P)$ endowed with a filtration $\{\mathcal F_t\}_{t\ge 0}$ representing the information structure available to the trader. Based on this probability space, we introduce the notion of admissible control. \begin{definition} A stochastic process $ \boldsymbol\theta(\cdot)=\{(\theta_t^{(1)} , \ldots, \theta^{(N)} _t), 0\le t\le T\}$ is called an admissible control process if the following conditions hold: \begin{description} \item[(i)] {\bf(Adaptivity)} For each time $ t\in[0,T]$, $\boldsymbol\theta_t=(\theta_t^{(1)},\cdots,\theta_t^{(N)}) $ is $\mathcal F_t$-adapted; \item[(ii)]{\bf(Non-negativity)} $ \boldsymbol\theta_t\in \mathbb R_+^N$, where $ \mathbb R_+^N$ is the set of nonnegative real-valued $N$-dimensional vectors; \item[(iii)]{\bf(Consistency)} $$ \displaystyle \int_0^T\sum_{n=1}^N\theta^{(n)} _tdt\le Q; $$ \item[(iv)]{\bf(Square-integrability)} $$ \displaystyle \mathbb E\left[\int_0^T[|\theta_t^{(1)} |^2+\cdots+|\theta_t^{(N)} |^2]dt\right]<\infty; $$ \item[(v)]{\bf($L_\infty$-integrability)} $$ \displaystyle \mathbb E\left[\max_{0\le t\le T}[|\theta_t^{(1)} |+\cdots+|\theta_t^{(N)} |]\right]<\infty. $$ \end{description} For convenience, we let $ \Theta_t$ denote the collection of admissible controls with respect to time $ t (<T)$, and let $ \widehat\Theta_t$ denote the collection of controls only satisfying the conditions (i), (iv) and (v). \end{definition} Suppose the stock price evolves over time according to\footnote{When long-term strategies are considered, it is more reasonable to consider geometric rather than arithmetic Brownian motion. In this paper, we mainly focus on short-term liquidating strategies, the total fractional price changes over such a short time are relatively small, and hence the difference between arithmetic and geometric Brownian motions can be negligible.} \begin{equation}\label{basic} d S_t= \sigma_t dW_t \end{equation} where $ \{W_t\}_{t\ge0}$ is a standard Brownian motion with filtration $\{{\cal F}_t\}_{t\ge0}$, and $ \sigma_t>0$ is the absolute volatility of the stock price, which can be (i) constant; (ii) time-dependent; (iii) volatility depending on the current stock price $ S_t$ and time $t$; and (iv) volatility driven by an additional random process. In Eq. (\ref{basic}), the drift term is set to be zero, which means that we expect no obvious trend in its future movement. The total time $T$ is usually one day or less, the drift is generally not significant over such a short trading horizon. \subsection{The Market Impact Model} Generally, risky assets, especially for those with low liquidity, exhibit price impacts due to the feedback effects of trader's liquidating strategies. Price impact refers to how the price moves in response to an order in the market. Small orders usually have insignificant impact, especially for liquid stocks. Large orders, however, may have a significant impact on the price. Investors, especially institutional investors, must keep the price impact in mind when making investment decisions. As discussed in Almgren \cite{RA1, RA2}, the price received on each trade is affected by the rates of buying and selling, both permanently and temporarily. The {\it temporal impact} is related to the liquidity cost faced by the investor while the {\it permanent impact} is linked to information transmitted to the market by the investor's trades. Almgren \cite{RA2} proposed a linear market impact model to describe the dynamics of the asset price (a single venue is considered): \begin{equation}\label{actually} dS^I_t= d S_t -\eta^{per}\theta_t dt. \end{equation} In this model, $\eta^{per}>0$ is the {\it coefficient of permanent impact} and is an absolute coefficient rather than fractional. In this section, in view of Almgren's linear price impact model, we consider the case of multiple venues based on equilibrium and no-arbitrage. Different from the single-venue case in Eq. (\ref{actually}), each venue's price in the multiple-venue case depends not only on its internal transactions but also on its competitor's deals. When one trader submits his orders to a venue, he will lead to a direct price impact in this venue. Other traders, being aware of this impact, will adjust their order scheduling to benefit from favorable prices across the different venues and thus affect prices. Suppose that there are $N$ venues, where the same financial instrument can be traded simultaneously, namely Venue $1$, Venue $2$, $\ldots$, Venue $N$. The following proposition provides a price impact model based on Almgren's linear price impact model. \begin{proposition} Assume that market liquidity remains unchanged over $ [0,T]$, and that there exists no arbitrage opportunity in the financial market. Based on the linear market impact model in Almgren \cite{RA2}, the affected asset price in the presence of $ N$ trading venues is given by $$ dS^I_t=dS_t- \eta^{per}(\beta^{(1)} \cdot \theta^{(1)} _t+\cdots+ \beta^{(N)} \cdot \theta^{(N)} _t)dt $$ where $ \eta^{per}>0$ is the coefficient of permanent impact, $ (\theta_t^{(1)} , \cdots, \theta_t^{(N)} )$ is the trader's liquidating speed, and $$ \begin{array}{l} \displaystyle \sum_{n=1}^N \beta^{(n)} =1, \quad \beta^{(n)} \in(0,1) \end{array} $$ with $\beta^{(n)}$ describing Venue $n$'s market efficiency. \end{proposition} \begin{proof} Without loss of generality, we prove that Proposition 1 holds for $ N=2$. Suppose that our investor's trading scheduling over a small interval $ [t,t+\Delta t)$ is $ (\theta^{(1)} _t,\theta^{(2)}_t)$ and that $ \theta^{(1)} _t<\theta^{(2)}_t$: liquidating $ \theta^{(1)} _t\Delta t$ shares in Book $1$, and $ \theta^{(2)}_t\Delta t$ shares in Book $2$. Here $ \theta^{(1)} _t$ and $ \theta^{(2)}_t$ are interpreted as liquidation rates in Venue $1$ and Venue $2$, respectively. Assume that trades occur immediately after time $t$. With the assumption of linear price impact, stock prices in the two trading venues, immediately after the execution of the orders, are drawn down to $ p_t^{(1)}=S^I_{t}-\eta^{per}\theta^{(1)} _t\Delta t$ and $p_t^{(2)}=S^I_{t}-\eta^{per}\theta^{(2)}_t\Delta t$, respectively. Obviously, we have $ p_t^{(1)}>p_t^{(2)}$. Other investors, being aware of this arbitrage opportunity, will adjust their trading schedules, buying from Venue $2$ and selling to Venue $1$, to benefit from favorable prices across the two venues. Suppose the financial market processes linear convergence, and the convergence speed is very quickly and proportional to market's efficiency. The adjusted stock prices at time $ t+\Delta t$ are then given by $$ \displaystyle p_{t+\Delta t}^{(1)} =S^I_{t}+\int_t^{t+\Delta t} \sigma_udW_u-\eta^{per}\theta^{(1)} _t\Delta t-x(\eta^{per}\theta^{(2)}_t-\eta^{per}\theta^{(1)} _t)\Delta t $$ and $$ \displaystyle p_{t+\Delta t}^{(2)}=S^I_{t}+\int_t^{t+\Delta t} \sigma_udW_u-\eta^{per}\theta^{(2)}_t\Delta t +y(\eta^{per}\theta^{(2)}_t-\eta^{per}\theta^{(1)} _t)\Delta t, $$ respectively. Under the no-arbitrage principle, $p_{t+\Delta t}^{(1)}= p_{t+\Delta t}^{(2)}$ which yields $x+y=1$, and $$ S^I_{t+\Delta t} = p_{t+\Delta t}^{(1)} = p_{t+\Delta t}^{(2)} = S^I_{t}+\int_t^{t+\Delta t}\sigma_udW_u-\eta^{per}(y\theta_t^{(1)} +x\theta_t^{(2)} )\Delta t. $$ Letting $ x=\beta$ and $ \Delta t\to 0$, we obtain $$ dS^I_t=dS_t-\eta^{per}[(1-\beta)\theta^{(1)}_t+\beta \theta^{(2)} _t]dt. $$ \end{proof} Different from the affected stock price (permanent price impact), the actual price received on each trade (temporary price impact) varies with place and time \begin{equation}\label{kkl} \widetilde S^{I,(n)}_t=S^I_t-\eta^{tem,(n)}\theta^{(n)} _t. \end{equation} Here $ \widetilde S_t^{I,(n)}$ is the price actually received in Venue $n$ ($n=1,2,\ldots,N$), and $\eta^{tem, (n)}$ is the coefficient of temporary market impact in that venue. A number of market impact models have been considered in the literature \cite{JG10}, but this simple one is good enough for our concerned problems and discussions. \subsection{Estimation of Model Parameters} Regarding the parameter estimation of $ \sigma_t$ in Eq. (\ref{basic}) and $ (\eta^{per}, \eta^{tem, (n)})$ in Eq. (\ref{actually}) and Eq. (\ref{kkl}), Almgren \cite{RA2} proposed a variety of methods using high-frequency market data: \begin{description} \item[(i)] For volatility $ \sigma_t$, to filter out noise associated with market details and obtain reliable estimates, one could estimate $ \sigma_t$ by using market data from the preceding 5 minutes, which typically would contain hundreds of trades and potentially thousands of quote updates; \item[(ii)] One proxy for price impact parameter $ \eta^{per}$ would be the realized trade volume over the last few minutes: if more people are trading actively in the market, one would be able to liquidate a certain quantity with lower slippage; \item[(iii)] For price impact parameters $ \eta^{tem, (n)}$, one proxy would be the trade volume resting at or near the bid price if one is a seller (at or near the ask price if one is a buyer): a large volume there might indicate the presence of a motivated buyer and a good opportunity for one to go in as a seller with low impact. \end{description} \subsection{The Gain/Loss of Trading } Let $ C_t$ denote the cash flow accumulated by time $t$. Assume that investors withhold the liquidation proceeds, or simply assume that the risk-free interest rate $ r=0$, we have $$ C_t =\int_0^t(\widetilde S^{I,(1)}_s \theta^{(1)}_s+ \cdots+\widetilde S^{I,(N)}_s \theta^{(N)} _s)ds, \quad t< T. $$ Given the state variables $ (S_t^I, C_t, X_t)$ the instant before the end of trading $ t=T-$, we have one final liquidation (if necessary) so that the number of shares owned at $ t=T$ is $ X_T=0$. The liquidation value $ C_T$ after this final trade is defined to be $$ C_T=C_{T-}+ X_{T-}(S^I_{T-}-\mathcal C^o(X_{T-})), $$ where $ \mathcal C^o(q)$, a non-negative increasing function in $q$, represents the market impact costs the trader incurs when liquidating the outstanding position $ X_{T-}$. The gain/loss (G/L) of trading, relative to the arrival price benchmark, is the difference between the total dollar gained/lost by liquidating $Q$ shares and the initial market value: $\displaystyle \mathcal R_T=C_T-QS_0$. After integrating by parts and using $$ X_t=Q-\int_0^t(\theta^{(1)}_s + \cdots+ \theta^{(N)} _s)ds, $$ we have \begin{equation}\label{R} \begin{array}{lll} \displaystyle \mathcal R_T&=&C_T-QS_0\\ &=&\displaystyle \int_0^{T-}\sigma_t X_tdW_t-X_{T-}\mathcal C^o(X_{T-}) \displaystyle -\eta^{per}\int_0^{T-}X_t(\beta^{(1)}\theta^{(1)}_t +\cdots+\beta^{(N)}\theta^{(N)} _t)dt \\ &&\displaystyle - \int_0^{T-}\left[\eta^{tem,(1)}(\theta^{(1)}_t)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _t)^2\right]dt. \end{array} \end{equation} Generally speaking, investors are risk averse and demand a higher return for a more risky investment. The mean-variance criterion is useful when taking both return and risk into account. However, the mean-variance criterion may induce a potential problem of time-inconsistency, i.e., planned and implemented policies are different, and hence complicate the problem. To avoid the time-inconsistent problem incurred by mean-variance criterion, instead of using the variance/standard deviation as the risk measure, one can adopt the quadratic variation, $$ \int_0^{T-}\sigma^2_uX_u^2du. $$ It accumulates the future value of the instantaneous risk, i.e., $ (X_udS_u)^2$, due to holding $ X_u$ units of the risky asset $S$. When properly normalized, the quadratic variation can also be interpreted as the average standard deviation per unit time, see, for instance, Brugiere \cite{PB96}. Let $ (S_t, \sigma_t, X_t)=(s, \sigma, q)$ be the initial state. At any time $ t\in[0, T)$, the optimal policy should solve the following optimization problem: \begin{equation}\label{mean-quadratic} J(t, s,\sigma, q)=\max_{(\theta^{(1)}_u,\cdots, \theta^{(N)} _u)_{ t\le u\le T}\in \Theta_t}\left\{ \mathbb E_{t}[\mathcal R_T-\mathcal R_t] -\lambda\cdot\mathbb E_t\left[\int_t^{T-} \sigma^2_uX_u^2du\right]\right\}, \end{equation} where $ \mathbb E_t[\cdot]$ denotes the conditional expectation with respect to the filtration $\mathcal F_t$. The following proposition discusses the time consistency of the optimal strategies and its proof is given in Appendix A. \begin{proposition} {\bf (Time consistency of the optimal strategies).} Let $ (t_1,s_1,\sigma_1, q_1)$ be some state at time $ t_1$ and $ \boldsymbol\theta^*_{t_1,s_1, \sigma_1, q_1}(\cdot) =\big(\theta^{(1)}_{t_1,s_1,\sigma_1, q_1}(\cdot), \ldots, \theta^{(N)} _{t_1,s_1, \sigma_1, q_1}(\cdot)\big)^T$ be the corresponding optimal strategy. Let $ (t_2, s_2, \sigma_2, q_2)$ be some other state at time $ t_2>t_1$ and $ \boldsymbol\theta^*_{t_2, s_2, \sigma_2, q_2}(\cdot)=\big(\theta^{(1)}_{t_2, s_2, \sigma_2, q_2}(\cdot), \ldots, \theta^{(N)} _{t_2, s_2, \sigma_2, q_2}(\cdot)\big)^T $ be the corresponding optimal strategy. It follows that the optimal controls of Problem (\ref{mean-quadratic}) are {\it time-consistent} in the sense that for the same state $(t^\prime, s^\prime, \sigma^\prime, q^\prime)$ at a later time $t^\prime>t_2$, \begin{equation}\label{proof} \boldsymbol\theta^*_{t_1,s_1, \sigma_1, q_1}(t^\prime, s^\prime, \sigma^\prime, q^\prime)= \boldsymbol\theta^*_{t_2, s_2, \sigma_2, t_2}(t^\prime, s^\prime, \sigma^\prime, q^\prime). \end{equation} \end{proposition} \subsection{Hamilton-Jacobi-Bellman (HJB) Equation} Since the optimal controls satisfy the Bellman's principle of optimality as shown in Appendix A, dynamic programming (DP) approach can be directly applied to this problem. The optimal control $\boldsymbol\theta^*$ can be obtained by solving the following Hamilton-Jacobi-Bellman (HJB) equation derived in Appendix A as follows: $$ \mbox{(HJB-1)}\quad\left\{ \begin{array}{lll} \displaystyle (\partial_t +\mathcal L) J -\lambda \sigma_t^2q^2\\ \displaystyle - \min_{\boldsymbol\theta_t\in\Theta_t } \Big\{ \eta^{per}(\beta^{(1)}\theta^{(1)}_t+\cdots+\beta^{(N)}\theta^{(N)} _t)q+ \eta^{tem,(1)}(\theta^{(1)}_t)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _t)^2\\ \displaystyle +(\theta_t^{(1)} +\cdots+\theta^{(N)} _t)\; \partial_q J\ \Big\} =0\\ J(T-,s, \sigma, q)= -q\mathcal C^o(q), \end{array} \right. $$ where $\mathcal L$ is the generator of the processes $ (S_t, \sigma_t)_{t\ge0}$. Notice that the optimization problem included in HJB-1 is a constrained optimization problem with constraints: (1) $\boldsymbol\theta_t\in\mathbb R_+^N$ and (2) $\displaystyle \int_0^T\sum_{n=1}^N\theta_t^{(n)} dt\le Q$. To solve this constrained optimization problem, we first consider relaxing the constraints, namely, replacing $\Theta_t$ by $\widehat\Theta_t$, and solving the following unconstrained optimization problem: $$ \mbox{(HJB-$1^\prime$)}\quad \left\{ \begin{array}{lll} \displaystyle (\partial_t +\mathcal L) J -\lambda \sigma_t^2q^2\\ \displaystyle - \min_{\boldsymbol\theta_t\in\widehat\Theta_t } \Big\{ \eta^{per}(\beta^{(1)}\theta^{(1)}_t+\cdots+\beta^{(N)}\theta^{(N)} _t)q+ \eta^{tem,(1)}(\theta^{(1)}_t)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _t)^2\\ \displaystyle +(\theta_t^{(1)} +\cdots+\theta^{(N)} _t)\; \partial_q J\ \Big\} =0\\ J(T-,s, \sigma, q)= -q\mathcal C^o(q), \end{array} \right. $$ and then prove that under some assumptions, the two optimization problems, included in HJB-1 and HJB-$1^\prime$, respectively, are equivalent. From the HJB-$1^\prime$ equation, the optimal control without any constraint is \begin{equation}\label{ef} \theta_t^{n,*}=-\frac{1}{2\eta^{tem, (n)}}(\partial_q J+\eta^{per}\beta^{(n)} q), \quad {\rm for} \ n=1,\cdots, N. \end{equation} The corresponding value function $J$ then solves the following partial differential equation (PDE): \begin{equation}\label{PDE} \left\{ \begin{array}{l} \displaystyle (\partial_t +\mathcal L) J -\lambda \sigma_t^2q^2+\sum_{n=1}^N \frac{1}{4\eta^{tem,(n)}}(\partial_q J+\eta^{per}\beta^{(n)} q)^2=0\\ \displaystyle J(T-,s,\sigma, q)=-q \mathcal C^o(q). \end{array} \right. \end{equation} \quad\\ In the following sections, we exhibit solutions to Eq. (\ref{PDE}) with linear penalty: $\mathcal C^o(q)=Kq$ to two special cases: (i) constant volatility and (ii) {\it slow mean-reverting} stochastic volatility. In each case, we work on the ansatz that $$ J(t, s,\sigma, q)=f(t, s,\sigma)+g(t,s,\sigma) q+h(t, s, \sigma) q^2. $$ The first case has explicit liquidating formula. For the {\it slow mean-reverting} stochastic volatility approximation discussed in this paper, there is a multi-scale argument that reduces the optimal liquidation PDE to a formal series expansion that can easily be solved explicitly. We present the argument and the accuracy of this approach in Section 3.3. \section{Constant Volatility} The most illuminating case is that $ \sigma_t$ is constant, i.e., $ \sigma_t\equiv \sigma$. This considerably simplifies the problem and allows us to exhibit the essential features of liquidating across multiple venues without losing ourselves in complexities. The problem becomes essentially the well-known stochastic linear regulator with time dependence. With the constant volatility assumption, $J(t, s, \sigma, q)=J(t, q; \sigma )$ independent of $s$, since the terminal data does not depend on $s$ and HJB-$1^\prime$ introduces no $s$-dependence. We look for a solution quadratic in the inventory variable $q$: \begin{equation}\label{lb1} J(t,q;\sigma)=f(t;\sigma)+g(t;\sigma)q+h(t;\sigma)q^2. \end{equation} With this assumption, the optimal control can be rewritten in the form of \begin{equation}\label{rew} \theta_t^{n,*}=-\frac{1}{2\eta^{tem, (n)}}[(2h(t;\sigma)+\eta^{per}\beta^{(n)})q +g(t;\sigma)].\end{equation} To solve Eq. (\ref{PDE}), $f, g$ and $h$ must satisfy the following ordinary differential equations (ODEs): \begin{equation}\label{ODE} \begin{array}{l} \left\{ \begin{array}{l} \displaystyle \dot{h} =\lambda\sigma^2-\sum_{n=1}^N\frac{1}{4\eta^{tem, (n)}}(2h +\eta^{per}\beta^{(n)})^2\\ \displaystyle h(T-;\sigma)=-K \end{array} \right.\\ \\ \left\{ \begin{array}{l} \displaystyle \dot{g}=-\left(\sum_{n=1}^N\frac{1}{2\eta^{tem, (n)}}(2h+\eta^{per}\beta^{(n)})\right)g\\ \displaystyle g(T-;\sigma)=0 \end{array} \right.\\ \\ \left\{ \begin{array}{l} \displaystyle \dot{f}=-\left(\sum_{n=1}^N\frac{1}{4\eta^{tem, (n)}}\right)g^2\\ \displaystyle f(T-;\sigma)=0. \end{array} \right. \end{array} \end{equation} It is straightforward to show that $ g(t;\sigma)\equiv 0$ and $ f(t;\sigma)\equiv 0$. If we set $$ a_n=\frac{1}{\eta^{tem, (n)}}, \quad b_n=\frac{\eta^{per}\beta^n}{\eta^{tem,(n)}}, \quad c_n= \frac{(\eta^{per}\beta^n)^2}{ 4\eta^{tem,(n)}}, $$ and $$ a=\sum_{n=1}^Na_n,\quad b=\sum_{n=1}^Nb_n,\quad c=\sum_{n=1}^N c_n, $$ then the unknown function $ h(t;\sigma)$ solves the following first-order ODE: \begin{equation}\label{ho} \left\{ \begin{array}{l} \displaystyle \dot{h}=\underbrace{\lambda\sigma^2+\frac{b^2-4ac}{4a}}_{\Delta_N}-a\left(h+\frac{b}{2a}\right)^2\\ \displaystyle h(T-;\sigma)=-K. \end{array} \right. \end{equation} Eq. (\ref{ho}) is a first-order ODE with constant coefficients, and we can find the exact solution via direct integrations: \begin{equation}\label{h} h(t;\sigma)= \left\{ \begin{array}{lll} \displaystyle \sqrt{\frac{\Delta_N}{a}}\cdot \frac{\varsigma e^{-2\sqrt{a\Delta_N}(T-t)}-1}{\varsigma e^{-2\sqrt{a\Delta_N}(T-t)}+1}-\frac{b}{2a},&& \Delta_N\ge0\\ \\ \displaystyle \sqrt{-\frac{\Delta_N}{a}} \tan\Big(\arctan\big(\frac{b-2a K}{2\sqrt{-a\Delta_N}}\big)+ \sqrt{-a\Delta_N} (T-t)\Big) -\frac{b}{2a},&& \Delta_N<0, \end{array} \right. \end{equation} where the constant $\varsigma$ is given by $$ \varsigma= \left(1-\frac{2Ka-b}{2\sqrt{a\Delta_N}}\right)\left(1+\frac{2Ka-b}{2\sqrt{a\Delta_N}}\right)^{-1}. $$ It is worth noting that $$ \theta_t^{n,*}=-\frac{1}{2\eta^{tem, (n)}}(2h(t;\sigma)+\eta^{per}\beta^{(n)})X_t. $$ and $$ \left\{ \begin{array}{l} \displaystyle \dot{X}_t=-\sum_{n=1}^N\theta^{n,*}_t\\ \displaystyle X_0=Q. \end{array} \right. $$ Hence, we have $$ X_t = Q\cdot \exp\left(\displaystyle \int_0^t\sum_{n=1}^N\frac{1}{2\eta^{tem, (n)}}(2h(u;\sigma) +\eta^{per}\beta^{(n)}) du\right). $$ For $n=1,2,\ldots, N$, we have $b_n^2=4a_nc_n$. According to H$\ddot{\rm o}$lder's inequality, we have $b^2-4ac\le 0$, and the equality holds if and only if (1) $ N=1$; or (2) the trading venues have the same market efficiency, namely, $ \beta^{(1)}=\cdots=\beta^{(N)}=1/N$. We then have the following proposition and its proof is given in Appendix B. \begin{proposition} Assume that the model parameters satisfy the condition\footnote{That is, clearing fees associated with the outstanding position $ X_{T-}$ dominate the potential profit arising from arbitrage opportunities incurred by the permanent price impact and the potential position risk involved by price fluctuations.}: \begin{equation}\label{cond1} K>\displaystyle \frac{b}{2a}+\sqrt{\frac{|\Delta_N|}{a}}. \end{equation} If $N=1$ or the trading venues have the same market efficiency, then $ h(t;\sigma)$ is a decreasing function in $t$ and $2h(t;\sigma)+\eta^{per}\beta^{(n)}\le0$ for $n=1,\cdots, N$, which implies that \begin{description} \item[(1)] $\theta^{n,*}_t \ge0$, for any $ n=1,\cdots, N$; and that \item[(2)] $\displaystyle \int_0^T\sum_{n=1}^N \theta^{n,*}_tdt \le Q$. \end{description} Then the control policy in Eq. (\ref{rew}) is optimal. \end{proposition} Since $$ \frac{\partial h(t;\sigma)}{\partial t} =-\frac{\partial h(t;\sigma)}{\partial \tau}<0, $$ where $ \tau=T-t$ is the remaining time to close of trading. $ J(t,q; \sigma)=h(t;\sigma)q^2$ is a strictly decreasing function in $t$ and a strictly increasing function in $T$. Generally speaking, one's ability to bear risk is measured mainly in terms of objective factors, such as time horizon, risk aversion and expected income. Let $ J_T(t,q;\sigma)$ denote the value function of the optimization problem (\ref{mean-quadratic}) with time horizon $T$, for any $ T_1>T_2>t$, we have $J_{T_1}(t,q;\sigma)> J_{T_2}(t,q;\sigma)$. This coincides with the actual situation that an investor's risk affordability directly relates to his/her time horizon. An investor with a two-week time horizon can be considered to have a greater ability to bear risk, other things being equal, than an investor with a two-hour horizon. This difference is because over two weeks there is more scope for losses to be recovered or other adjustments to circumstances to be made than there is over two hours. \subsection{Effect of Multiple Venues} With the development of electronic exchanges, many new trading destinations have appeared to compete the trading capability of the fundamental financial markets such as the NASDAQ's Inet and NYSE in the US, or EURONEXT, the London Stock Exchange and Xetra in Europe. As a result, the same financial instrument can be traded simultaneously in different venues. These trading venues are generally different from each other at any time because of variation in the fees or rebates they demand to trade and the liquidity they offer. Therefore, to liquidate a large order, traders may need to split their order across various venues to reduce market impact. In this subsection, we illustrate theoretically how multiple venues affect an investor's trading strategy. Denote the liquidating strategy corresponding to a single venue case as $ \theta_t^{single,*}$. Suppose that there are $N$ distinct venues for an investor to submit his/her trades and that condition (\ref{cond1}) is satisfied. Without loss of generality, we assume that there is no difference among there trading venues, namely $ \beta^1=\cdots=\beta^N=1/N$ and $ \eta^{tem, (1)}=\cdots=\eta^{tem, (N)}=\eta^{tem}$. The following two conclusions about the effects on optimal liquidating speed and transaction cost can be drawn. \begin{description} \item[(i)] {\bf(Effect on optimal liquidating speed).} First, we observe that $ \theta^{1,*}_t=\cdots=\theta^{N,*}_t:=\theta^*_t(N)$, where $$ \theta^*_t(N)= - \sqrt{\frac{\lambda\sigma^2 }{\eta^{tem}N}}\times \frac{\varsigma(N)e^{-2\sqrt{\lambda\sigma^2N/\eta^{tem}}(T-t)}-1}{\varsigma(N)e^{-2\sqrt{\lambda\sigma^2N/\eta^{tem}}(T-t)}+1}X_t:=\mathfrak{J}(t, N)X_t, $$ with $$ \varsigma(N)=\frac{2\sqrt{\frac{\lambda\sigma^2N}{\eta^{tem}}}-(2K\frac{N}{\eta^{tem}}-b)}{2\sqrt{\frac{\lambda\sigma^2N}{\eta^{tem}}}+(2K\frac{N}{\eta^{tem}}-b)}. $$ Notice that $ X_0=Q$ and that $$ \lim_{N\to\infty} \dot{X}_0=-\lim_{N\to\infty} N\theta^{*}_0(N)=-\lim_{N\to\infty} N\mathfrak{J}(0,N)Q=\infty. $$ For any fixed time $ t\in(0,T)$, we have $$ \lim_{N\to\infty}\mathfrak{J}(t, N)=0 \quad {\rm and} \quad \lim_{N\to\infty}\sqrt{N}\mathfrak{J}(t,N)=\sqrt{\frac{\lambda\sigma^2}{\eta^{tem}}}. $$ Hence $$ \begin{array}{lll} \displaystyle \lim_{N\to\infty} X_t&=& \displaystyle\lim_{N\to\infty} Q e^{-\int_0^tN\mathfrak{J}(u,N)du} = \displaystyle \lim_{N\to\infty} Q e^{-\sqrt{N}\times \int_0^t\sqrt{N} \mathfrak{J}(u,N)du}= \displaystyle \lim_{N\to\infty} Q e^{-a_1 \sqrt{N}} =0, \end{array} $$ where\footnote{According to the structure of $\mathfrak{J}(t, N)$, there exists a finite number $M>0$ such that $$ |\sqrt{N}\mathfrak{J}(t, N)|\le M,\quad \mbox{for all $(t, N)\in[0,T)\times \mathbb Z_+$,} $$ where $\mathbb Z_+$ is the set of all nonnegative integers. Directly applying the Dominated convergence theorem to $\{\sqrt{N}\mathfrak{J}(t, N)\}$, we obtain the result in Eq. (\ref{pk}). } \begin{equation}\label{pk} 0<a_1=\lim_{N\to\infty} \int_0^t\sqrt{N}\mathfrak{J}(u, N)du=\sqrt{\frac{\lambda\sigma^2}{\eta^{tem}}} t<\infty. \end{equation} We also have $$ \begin{array}{lll} \displaystyle \lim_{N\to\infty} -\dot{X}_t &=& \displaystyle \lim_{N\to\infty}N\theta^{*}_t(N)\\ &=& \displaystyle \lim_{N\to\infty}QN\mathfrak{J}(t,N)\times e^{-\int_0^tN\mathfrak{J}(u,N) du }\\ &=& \displaystyle \lim_{N\to\infty} Q \frac{\sqrt{N}\times\sqrt{N}\mathfrak{J}(t,N)}{e^{\sqrt{N}\times\int_0^t\sqrt{N}\mathfrak{J}(u, N)du}}\\ &=& \displaystyle \lim_{N\to\infty} Q\sqrt{\frac{\lambda\sigma^2}{\eta^{tem}}}\times \frac{\sqrt{N}}{e^{a_1\sqrt{N}}}\\ &=& \displaystyle \lim_{N\to\infty} Q\sqrt{\frac{\lambda\sigma^2}{\eta^{tem}}}\times \frac{1}{a_1e^{a_1\sqrt{N}}}\\ &=&0. \end{array} $$ That is to say, as $N$ approaches infinity, the investor would immediately close his/her position at the beginning of the trading horizon. \item[(ii)]{\bf(Effect on transaction cost).} The two strategies: 1) $ \{\theta_t^{single,*}\}_{t\in[0,T]}$, liquidating in a single venue; and 2) $$ \Big\{\theta^{(1)}_t=\theta^{(2)}_t=\cdots=\theta^{(n)} _t= \frac{1}{N}\theta_t^{single,*}\Big\}_{t\in[0,T]}, $$ equally splitting the original target among $N$ venues, transmit the same information to the market (i.e., have the same permanent impact), but involve different transaction costs (i.e., have different temporary price impacts) $$ \begin{array}{lll} \displaystyle\underbrace{ \int_0^{T-}\sum_{n=1}^N\left(S_t^I-\widetilde S_t^{I,(n)}\right)\theta^{(n)} _tdt}_{\mbox{costs in $N$-venue}} &=& \displaystyle\int_0^{T-} \left(\sum_{n=1}^N\frac{\eta^{tem}}{N^2}\right)(\theta_t^{single,*})^2dt\\ &=& \displaystyle\int_0^{T-} \frac{\eta^{tem}(\theta_t^{single,*})^2}{N}dt\\ &\le& \displaystyle \int_0^{T-}\eta^{tem}(\theta_t^{single,*})^2dt=\underbrace{ \int_0^{T-}\left(S_t^I-\widetilde S_t^I\right)\theta_t^{single,*}dt.}_{\mbox{costs in a single venue}}\\ \end{array} $$ That is, liquidating schedule across multiple venues can indeed help to reduce transaction costs arising from market liquidity. \end{description} \subsection{Numerical Results} In this section, we provide some numerical results to illustrate the effects of different market factors on investor's liquidation strategy. We assume that there are $N$ distinct trading venues for an investor to submit his/her trades, and that there is no difference among these venues. As far as our simulation is concerned, we choose the following hypothetical values for the model parameters: $$ \sigma=\exp(1), \ \ T=1, \ \ K=0.1, \ \ \eta^{per}=0.005, \ \ \eta^{tem}=0.01. $$ The risk-aversion parameter $\lambda$ ranges across all nonnegative values, as the actual choice of trajectory will be determined by the trader's risk preference. \subsubsection{Effect of Multiple Venues} When multiple trading venues are available, dividing a target quantity across these venues may help an investor to hide his liquidation purpose and hence reduce the permanent market impact. To facilitate our analysis, we assume that these trading venues are identical for the investor. Some numerical results are presented in Table 1. \begin{table}[H] \caption{1, 000 Simulations with $ Q=100, s=15, \lambda=0.1$.} \begin{center} \begin{tabular}{cccccc} \hline \cline{1-6} & &Std&&Std& \\ \#\{Venues\} &G/L & (G/L$)^*$&$ q_T$&$ (q_T)^{**}$&Value function\\ \hline 1&-461.80&63.38&1.43&2.22& -902.90\\ 2&-324.86&54.19&0.03&0.04&-638.30\\ 3&-265.08&46.12&0 &0 &-522.83\\ 4& -230.51& 43.03& 0 & 0 & -455.06\\ 10& -147.22& 33.51&0&0 &-293.05\\ \vdots&\\ 50&-72.92&18.57&0&0&-144.59\\ \hline \cline{1-6} {\scriptsize*: in $10^{-2}$}\\ {\scriptsize**: in $10^{-16}$}\\ \end{tabular} \end{center} \label{table1} \end{table} Table 1 provides a comparison of optimal liquidation strategy among investors facing multiple venues. The first column shows the number of venues available for an investor to submit his/her trades. The second and third columns show, respectively, the mean and standard derivation of G/L (Eq. (\ref{GL})). The negative G/L implies that the investor is trading at a loss. The absolute value represents the {\it cost of trading}. As defined so far, the {\it cost of trading}, relative to the arrival price benchmark, is the difference between the total dollars received to liquidate Q shares and the initial market value: \begin{equation}\label{GL} { G/L}= \int_0^T(\widetilde S_t^{I,(1)}\theta^{(1)}_t +\cdots+\widetilde S_t^{I,(N)}\theta^{(N)} _t)dt+X_{T-}(S_{T-}^I-\mathcal C^o(X_{T-}))-QS_0\quad(<0). \end{equation} It is an indicator of market liquidity. The greater the loss, the worse the situation. The last column shows the corresponding value taken by the value function: $$ { G/L}-\lambda \mathbb E\left[\int_0^T \sigma_t^sX_t^2dt \right] \quad(<0), $$ an indicator of the investor's satisfaction. Investors are risk-averse ($\lambda>0$), so usually they do not hold outstanding stock uncleared at the end of trading $ t=T-$. The fourth and fifth columns show, respectively, the mean and standard deviation of the outstanding position at the end of trading. The effect of multiple venues on investors' liquidation strategies is fairly straightforward. With the increase of trading venue, investors' loss of trading decreases. Meanwhile, as the number of trading venue increases, the risks involved decreases, which in turn enhances investors' satisfaction (measured by the value function). \subsubsection{Trading Curve} The average number of shares at each point of time, say the trading curves, with respect to different number of multiple venues are depicted in Figure \ref{fig2}. \begin{figure}[H] \begin{center} \includegraphics[width=5.in, height=4.5in]{shares.pdf} \caption{ Trading curves with $ Q=100$. } \label{fig2} \end{center} \end{figure} It is clear that as the number of alternative choices increases, the average liquidation speed increases. Indeed, a trader needs to trade fast to reduce position risk, when the quantity to liquidate is large within the remaining time. However, an immediate execution is often not possible or at a very high cost due to insufficient liquidity. The trading cost reflects the difference between the amount at which the trader expects to sell and the sales proceeds the trader actually receives. It is an indicator of the illiquidity of a market. We can see from the results that, when more trading venues spring up in the financial market, the liquidity of the asset enhances. This provides good supplies of liquidity to the trader, and that the trader with more choices of venues to submit his/her trades may be willing to close his/her position earlier. These observations are consistent with financial intuition. \subsubsection{Efficient Frontier} The efficient frontier consists of all optimal trading strategies. Here the ``optimal'' refers to the situation where no strategy has a smaller variance for the same or higher level of expected transaction profits. For each level of risk-aversion, $\lambda$, there is an optimal liquidating strategy. By running 1,000 simulations with initial inventory $ Q=100$, we obtain an efficient frontier (see Figure \ref{fig3}). \begin{figure}[H] \begin{center} \includegraphics[width=5.in, height=4.5in]{ef.pdf} \caption{The efficient frontier. } \label{fig3} \end{center} \end{figure} Each point on the frontier represents a distinct strategy for an investor with certain level of risk-aversion. It shows the tradeoff between the expected revenues and the standard deviation. As we can see from this figure, the frontier increases along an approximate smooth concave curve. The slope of the ``tangent line'' indicates the trader's risk aversion level $\lambda$. The point on the most right is obtained for a risk-neutral trader with $\lambda=0$. We define this point as $ (Std_0, R_0)$. For any other point on the left, we have $$ R_t-R_0\approx\frac{1}{2}(Std_t-Std_0)^2\frac{d^2R}{d Std^2} \Big|_{Std=Std_0}. $$ A crucial insight is that for a risk neutral trader, a first-order decrease in the expected revenue can approximately incur a second-order decrease in the standard deviation. The efficient frontier depicted in Figure 2 is consistent with the one for the mean-variance portfolio selection. \subsubsection{Effect of Illiquidity} The cost of trading has three main components, brokerage commission, bid-ask spread (we will discuss this factor later) and price impact, of which liquidity affects the latter two. Brokerage commission is usually negotiable and does not constitute a large fraction of the total cost of trading except in small-size trades, which is a topic beyond the scope of this paper. Stocks with low liquidity can have wide bid-ask spreads. The bid-ask spread, which is the difference between the buying price and the selling price, is incurred as a cost of trading a security. The larger the bid-ask spread, the higher is the cost of trading. Liquidity also has implications for the price impact of trade. The extent of the price impact depends on the liquidity of the stock. A stock that trades millions of shares a day may be less affected than a stock that trades only a few hundred thousand shares a day. The effect of temporary market impact $ \eta^{tem}$, a key indicator of the illiquidity of a market, is shown in Figure 3. It is clear that traders in a market with larger temporary impact will involve higher transaction costs; and that the cost distribution behaves like a normal distribution under the assumption of constant volatility. We refer interested readers to the analysis of the effects of market impact, both permanently and temporarily, to the work by Almgren \cite{Almgren3}. \begin{figure}[H] \begin{center} \includegraphics[width=5.in, height=3.4in]{tem2.pdf} \label{fig4} \end{center} \end{figure} \begin{figure}[H] \begin{center} \includegraphics[width=5.in, height=3.4 in]{qq2.pdf} \caption{ {\bf (Left-hand Side):} $ \eta^{tem}=0.005$; \quad \quad {\bf(Right-hand Side):} $ \eta^{tem}=0.01$. } \label{fig40} \end{center} \end{figure} \section{Stochastic Volatility Model} The constant volatility assumption might be reasonable for large-cap US stocks. However, this assumption becomes defective when we move to the analysis of small and medium-capitalization stocks, whose volatilities vary randomly through the day. In this section, we relax the constant volatility assumption. Thus the value function $J=J(t, s, \sigma, q)$ is a function satisfies HJB-1. Similarly, we consider relaxing the constraints associated with HJB-1 and solve the unconstrained optimization problem in HJB-$1^\prime$. We then verify that the obtained optimal control does satisfy all the constraints in HJB-1. We present a framework that is general enough to take into account the effect of stochastic volatility in the following. \subsection{Slow Mean-reverting Volatility Model} We analyze models in which stock prices are conditionally normal\footnote{As we have discussed before, it is reasonable when considering short-term liquidating problem.}, and the volatility process is a positive and increasing function of a mean-reverting Ornstein-Uhlenbeck (OU) process. That is, \begin{equation}\label{stochastic} \left\{ \begin{array}{lll} \displaystyle dS_t=\phi(\nu_t)dW_t\\ \displaystyle d\nu_t=\epsilon (m -\nu_t)dt + \xi \sqrt{ \epsilon}\; dB_t\\ B_t=\rho W_t+\sqrt{1-\rho^2}Z_t\\ \end{array} \right. \end{equation} where $ \{W_t\}$ and $ \{Z_t\}$ are two independent one-dimensional Brownian motions, and $\rho$ is the correlation between price and volatility shocks, with $ |\rho|<1$. The parameter $m$ is the equilibrium or long-term mean, $\epsilon (>0)$ is the intrinsic time-scale of the process, and $\xi$, combining with the scalar $\sqrt{\epsilon}$, represents the degree of volatility around the equilibrium $m$ caused by shocks. Here $ \{\nu_t\}$ is a simple building-block for a large class of stochastic volatility models described by choice of $\phi(\cdot)$. According to It\^{o}'s lemma, $$ d\phi(\nu_t)=\epsilon\left[\phi^\prime(\nu_t)(m-\nu_t)+\xi^2\phi^{\prime\prime}(\nu_t)\right]dt+\xi \phi^\prime(\nu_t)\sqrt{\epsilon}dB_t. $$ We call these models mean-reverting because the volatility is a monotonic function of $ \{\nu_t\}$ whose drift pulls it towards the mean value $m$. The volatility is correspondingly pulled towards {\it approximately} $ \phi(m)$.\\ Plenty of analysis on specific It\^{o} models by numerical and analytical methods can be found in the literature \cite{RA2, Fouque00}. Our goal here is to identify and capture the relevant features of liquidating strategies for small- and medium-capitalization stocks. Our framework in Eq. (\ref{stochastic}) is adequate and efficient enough to capture this. Here we will work in the regime $\epsilon\ll 1$ (slow mean-reverting). For simplicity, we consider the case of $ N=1$. The value function $J$ under this setting solves the following PDE: \begin{equation}\label{PDE2} \left\{ \begin{array}{lll} \displaystyle\left( \partial_t +\mathcal L \right)J -\lambda \phi^2(\nu)q^2 +\frac{1}{4\eta^{tem}}(\partial_q J+\eta^{per} q)^2=0\\ J(T-,s, \nu, q)=-Kq^2. \end{array} \right. \end{equation} Furthermore, we have $ J(t, s, \nu,q)=J(t, \nu,q)$ independent of $s$, since the terminal data does not depend on $s$ and PDE (\ref{PDE2}) introduces no $s$-dependence. Hence, the operator $\mathcal L$ takes the form of $$ \mathcal L=\epsilon\left((m-\nu)\partial_\nu+\frac{1}{2}\xi^2 \partial_{\nu\nu}\right). $$ \noindent Similarly as before, we look for a solution quadratic in the inventory variable $q$ \begin{equation}\label{rt} J(t,\nu,q)=f(t,\nu)+g(t,\nu) q+h(t,\nu ) q^2. \end{equation} The optimal liquidating schedule can then be calculated through the relation \begin{equation}\label{rep1} \theta^{*}(t,\nu,q)=-\frac{1}{2\eta^{tem}}\big[\big(2h(t,\nu)+ \eta^{per}\big)q+g(t,\nu)\big]. \end{equation} To solve Eq. (\ref{rt}), $f, g$ and $h$ must satisfy the following PDEs: \begin{equation}\label{poly} \begin{array}{lll} \left\{ \begin{array}{l} \displaystyle (\partial_t+\mathcal L) h -\lambda\phi^2(\nu)+ \frac{1}{4\eta^{tem}}(2h+\eta^{per})^2=0\\ \displaystyle h(T-, \nu)=-K \end{array} \right.\\ \\ \left\{ \begin{array}{l} \displaystyle (\partial_t+\mathcal L) g+ \frac{1}{2\eta^{tem}}(2h+\eta^{per})g=0\\ \displaystyle g(T-,\nu)=0\\ \end{array} \right.\\ \\ \left\{ \begin{array}{l} \displaystyle (\partial_t+\mathcal L) f+ \frac{1}{4\eta^{tem}}g^2=0\\ \displaystyle f(T-, \nu)=0. \end{array} \right. \end{array} \end{equation} \quad\\ It is straightforward to verify that (Feynman-Kac formula) $$ g(t,\nu)\equiv 0\quad \mbox{and} \quad f(t,\nu)\equiv 0. $$ If we set $$ Y_t=\frac{1}{2\eta^{tem}}\int_0^t(2h(u, \nu_u)+\eta^{per})du\quad{\rm and}\quad \mathfrak{h}(t,\nu_t, Y_t)= e^{Y_t} (2h(t,\nu_t)+\eta^{per}). $$ According to It\^{o}'s lemma, $$ \begin{array}{lll} d\mathfrak{h}(t,\nu_t, Y_t) &=& \displaystyle e^{Y_t} \Big\{\big[ 2(\partial_t+\mathcal L)h+\frac{1}{2\eta^{tem}}(2h+\eta^{per})^2\big]dt +2\xi\sqrt{\epsilon}\partial_\nu h\; dB_t\Big\}\\ &=& \displaystyle e^{Y_t} \left\{2\lambda \phi^2(\nu_t) dt +2\xi\sqrt{\epsilon}\partial_\nu h\; dB_t\right\}. \end{array} $$ Applying It\^{o}'s formula to $\mathfrak{h}(u,\nu_u, Y_u)$ between $t$ and $T-$, with $h$ satisfying PDE (\ref{poly}), we get: $$ \mathfrak{h}(t,\nu_t, Y_t)= e^{Y_{T-}}(-2K+\eta^{per})-2\lambda\int_t^Te^{Y_u}\phi^2(\nu_u)du-2\int_t^T\xi\sqrt{\epsilon} \partial_\nu h(u,\nu_u)dB_u. $$ Hence, \begin{equation}\label{sol} \mathfrak{h}(t,\nu, y)=\mathbb E\left[ e^{Y_{T-}}(-2K+\eta^{per})-2\lambda\int_t^Te^{Y_u}\phi^2(\nu_u)du\Big|\nu_t=\nu, Y_t=y\right]. \end{equation} If condition (\ref{cond1}) is satisfied, then $ -2K+\eta^{per}<0$, and hence $ \mathfrak{h}(t,\nu,y)\le 0$ for any $ (t,\nu,y)\in[0,T)\times \mathbb R\times \mathbb R$. Therefore, $ 2h(t,\nu)+\eta^{per}\le 0$ for any $ (t,\nu)\in[0,T)\times \mathbb R$. Then the control policy in Eq. (\ref{rep1}) is optimal. \\ Eq. (\ref{sol}) just provides an implicit analytical solution to $\mathfrak{h}$. However, for the general coefficients $(\phi, m, \epsilon, \xi)$, we do not have an explicit solution. The small-$\epsilon$ regime gives rise to a regular perturbation expansion in the powers of $\epsilon$. As in \cite{Fouque00}, we seek an asymptotic approximation for $h$: \begin{equation}\label{ghj} h(t,\nu)=h^{(0)}(t,\nu)+\epsilon h^{(1)}(t,\nu)+\epsilon^2 h^{(2)}(t,\nu)+\cdots+. \end{equation} This is a series expansion, for which we find $ h^{(i)}$ for $ i=0,1$ explicitly, and study the accuracy when using the truncated series in Section 4.3.2. In the next sections, we present our asymptotic approach for $h$ and the accuracy of our approximate optimal liquidation policy. \subsection{Asymptotics} We first construct a regular perturbation expansion in the powers of $\epsilon$ by writing $$ \mathcal L=\epsilon \mathcal L_0, \quad h(t,\nu)=\sum_{i\ge 0}\epsilon^i h^{(i)}(t,\nu) $$ where $$ \mathcal L_0=(m-\nu)\partial_\nu+\frac{1}{2}\xi^2 \partial_{\nu\nu}. $$ Substituting these expansions into Eq. (\ref{poly}) and grouping the terms of the powers of $\epsilon$, we find that the lowest order equations of the regular perturbation expansion are \begin{equation}\label{like} \begin{array}{lll} \mathcal O(1):&& \displaystyle \quad\quad\quad\quad0= \partial_t h^{(0)}-\lambda\phi^2(\nu)+\frac{1}{4\eta^{tem}}(2h^{(0)}+\eta^{per})^2 \\ \mathcal O(\epsilon):&& \displaystyle \quad\quad\quad\quad0= \partial_t h^{(1)}+ \mathcal L_0 h^{(0)}+\frac{1}{\eta^{tem}} (2h^{(0)}+\eta^{per})h^{(1)} \end{array} \end{equation} with terminal data $$ h^{(0)}(T-, \nu)=-K\quad \mbox{and}\quad h^{(1)}(T-, \nu)=0. $$ The solutions of Eq. (\ref{like}) can be obtained easily. Indeed, we have \begin{eqnarray}\label{lm} \left\{ \begin{array}{lll} \displaystyle h^{(0)}(t,\nu)= \sqrt{ \eta^{tem}\lambda \phi^2(\nu)} \cdot \frac{\varsigma(\nu)e^{-2\sqrt{\lambda \phi^2(\nu)/\eta^{tem}}(T-t)}-1}{\varsigma(\nu) e^{-2\sqrt{\lambda \phi^2(\nu)/\eta^{tem}}(T-t)}+1}-\frac{1}{2}\eta^{per} \\ \displaystyle h^{(1)}(t,\nu)=\int_t^T D(r;t) \mathcal L_0h^{(0)}(r,\nu) dr \\ \end{array} \right. \end{eqnarray} where $$ \varsigma(\nu)=\frac{\sqrt{\eta^{tem}\lambda\phi^2(\nu)}-(K-\eta^{per}/2)}{ \sqrt{\eta^{tem}\lambda\phi^2(\nu)}+(K-\eta^{per}/2)} \quad {\rm and} \quad D(r;t)=\exp\left(\frac{1}{\eta^{tem}}\int_t^r(2h^{(0)}+\eta^{per})d\tau\right). $$ \subsection{Optimal Strategy} We now analyze and interpret how the principle expansion terms for the value function can be used in the expression for the optimal liquidation strategy $\theta^*$, which leads to an approximate feedback policy of the form: $$ \theta^*(t,\nu,q)=\theta^{(0),*}(t,\nu,q)+\epsilon \;\theta^{(1),*}(t,\nu,q)+\cdots+. $$ \subsubsection{Moving-Constant-Volatility Strategy} First, we introduce the zeroth order terms $ h^{(0)}$ in the expansion for $h$. This gives the zeroth order optimal liquidation strategy $$ \theta^{(0),*}(t,\nu,q):=-\frac{1}{2\eta^{tem}}\left(2h^{(0)}(t,\nu) +\eta^{per}\right)q. $$ Recall that, in the case of constant volatility approach (Section 3.1), \begin{equation}\label{cons} \theta^{Cons}(t,q;\sigma)=-\frac{1}{2\eta^{tem}}\left(2h^{(0)}(t,\nu^{Cons})) +\eta^{per}\right)q, \end{equation} where $ \nu^{Cons}=\phi^{-1}(\sigma)$ is a constant over $ [0,T)$. This approximation might be reasonable for most large-cap US stocks, for which, over short-term time frames like one day or less, the volatility can be regard as a constant. However, this might not be a reasonable approximation for assets that are less heavily traded than large-cap US stocks. The naive constant-approach liquidation strategy for small- and medium-capitalization stocks would adopt the strategy for large-cap US stocks, $ \theta^{Cons}$, but with the coefficient $\sigma$ driven by a varying factor $\nu$, i.e., \begin{equation}\label{moving} \theta^{Cons}(t, q; \phi(\nu))=\theta^{(0),*}(t,\nu,q). \end{equation} We call this strategy the {\it moving constant-approach liquidation strategy}. This strategy is a successful one if the parameters remains constant when they are estimated from updated segments of historical data. However, this is not always the case as the risk environment is dynamic. \subsubsection{First-order Correlation to Optimal Policy} The approximation to the optimal strategy can be more accurate by going into the higher order terms. Substituting the expansion of $h$ up to terms in $\epsilon$ gives $$ \theta^*(t,\nu,q)=\widetilde\theta^*(t,\nu,q)+\mbox{higher order terms} $$ where \begin{equation}\label{po23} \widetilde\theta^*(t,\nu,q)=-\frac{1}{2\eta^{tem}}\left(2h^{(0)}(t,\nu) +\eta^{per}\right)q- \frac{\epsilon}{\eta^{tem}}h^{(1)}(t,\nu)q, \end{equation} and $ h^{(0)}$ and $ h^{(1)}$ are given by Eq (\ref{lm}). In Eq.(\ref{po}), the term with $\epsilon$ corresponds to traders' response to the risk arising from stochastic volatility, which can be regarded as the principle term hedging the risk factor. In the previous sections, we presented a ``pure'' stochastic volatility approach to the liquidating problem, in which volatility $ \sigma_t$ is modeled as an It\^{o} process driven by a Brownian motion that has a component independent of the Brownian motion driving the asset price. In comparison with Almgren's work in \cite{RA2}, we mainly focus on a special class of volatility model, time-scale volatility model, and worked in the regime $\epsilon\ll 1$. The separation of time-scales with respect to this kind of models provided an efficient way to identify and track the effect of randomly varying volatility. The theorem below justifies the accuracy of this approximation. \begin{theorem} For fixed $ t<T$, $\nu$ and $ q$, the optimal liquidating strategy $ \theta^*(t,\nu,q)$ defined in Eq. (\ref{rep1}), and our asymptotic approximation $ \widetilde \theta^*(t,\nu,q)$ defined in Eq. (\ref{po23}) satisfy $$ | \theta^*(t,\nu,q) -\widetilde \theta^*(t,\nu,q) |=\mathcal O(\epsilon^2). $$ \end{theorem} \begin{proof} We note that, at any time $ t<T$, according to Eqs. (\ref{rep1}), (\ref{ghj}) and (\ref{po23}) $$ |\theta^*(t,\nu,q)-\widetilde\theta^*(t,\nu,q)|=\frac{|h-(h^{(0)}+\epsilon\;h^{(1)})|}{\eta^{tem}}q. $$ Defining the residual $ R^\epsilon= h -( h^{(0)}+\epsilon\; h^{(1)})$. It solves the following PDE: $$ \left\{ \begin{array}{lll} \displaystyle (\partial_t+\mathcal L)R^\epsilon+ \frac{1}{\eta^{tem}}R^\epsilon[h+ h^{(0)}+\epsilon h^{(1)}+\eta^{per}]+\epsilon^2 \left\{\mathcal L_0h^{(1)}+\frac{1}{\eta^{tem}}(h^{(1)})^2\right\}=0\\ R^\epsilon(T-,\nu)=0. \end{array} \right. $$ Directly applying the Feynman-Kac formula to this equation yields \begin{equation}\label{hj} \displaystyle R^\epsilon(t,\nu)=\epsilon^2\cdot \mathbb E_t \int_t^TD_2(r; t)\{ \mathcal L_0h^{(1)}+\frac{1}{\eta^{tem}}(h^{(1)})^2\} dr \end{equation} where $$ D_2(r; t)= \exp\left(\frac{1}{\eta^{tem}}\int_t^r[h+ h^{(0)}+\epsilon h^{(1)}+\eta^{per}]d\tau\right). $$ One can see by direct computation that the integrand in Eq. (\ref{hj}) is bounded in $ [0,T]$, i.e., there exists a constant $ C$, such that $ |R^\epsilon(t,\nu)|\le C\epsilon^2$ which completes our proof. \end{proof} \quad\\ We remark that, given an initial value $\nu_0=\nu$, $$ \nu_t\big|\nu \sim \mathcal N\Big(m+(\nu-m)e^{-\epsilon t}, \;\frac{\xi^2}{2 }(1-e^{-2\epsilon t})\Big), $$ where $ \mathcal N(\mu,\sigma^2)$ is the Gaussian distribution with mean $\mu$ and standard variation $\sigma$. Assume that $$ \sigma_t=\phi(\nu_t) \quad { or} \quad \nu_t=\phi^{-1}(\sigma_t) $$ is observable in real time with some reasonable degree of confidence. Indeed, we cannot observe directly the volatility, what we may observe is the Volatility Index (VIX) which may be used as a proxy for $ \sigma_t$. There are different techniques to estimate the parameters $ (\epsilon, m,\xi)$. These estimators rely on the persistence of market properties (volatility and liquidity), so that information about the past provides reasonable forecasts for the future. Such persistence, at least across short horizons, is well documented (see, for instance, Bouchaud et al. \cite{JP} for more details). \subsection{Simulation Results} In this section, we assume $ \phi(\nu)=e^{\nu}$ to test the performance of three different strategies on small- and medium-capitalization stocks: (i) constant volatility approach in Eq. (\ref{cons}); (ii) moving constant volatility approach in Eq. (\ref{moving}); and (iii) first-order correction in Eq. (\ref{po}). We refer to these strategies as ``(i) Constant-Vol'', `` (ii) Moving-Constant-Vol" and ``(ii) Vol-adjust'', respectively. As far as our simulation is concerned, we use the following hypothetical values of the model parameters: $ \nu=0.5, \ m=1, \ \epsilon=0.01, \ \xi=2, \ \rho=-0.4$\footnote{A negative value for $\rho$ is used here to capture the leverage effect.}. The rest of the parameters are assumed to be the same as those used in Section 3.2. The simulation is obtained through the following procedure: at time t, the trader's trading rate $ \theta_t$ is computed, given the state variables. At time $ t+dt$, the mid-price is updated by a random increment $ \pm e^\nu\sqrt{dt}$. The volatility is updated accordingly by a random increment: $$ \epsilon(m-\nu)dt+\xi\sqrt{\epsilon}(\rho\sqrt{dt}\pm \sqrt{1-\rho^2}\sqrt{dt} ) \;\; \mbox{or} \;\; \epsilon(m-\nu)dt+\xi\sqrt{\epsilon} (-\rho\sqrt{dt}\pm \sqrt{1-\rho^2}\sqrt{dt} ). $$ Figure \ref{fig5} illustrates the dynamics of the volatility for one simulation path. From the simulation path depicted in Figure \ref{fig5}, we can see the leverage effect between the price and the volatility. \begin{figure}[H] \begin{center} \includegraphics[width=5in, height=4.5in]{dyn.pdf} \caption{ Dynamics of $ (S_t, \exp(\nu_t))$ with $\rho=-0.4$. } \label{fig5} \end{center} \end{figure} We then run $1,000$ simulations to compare the performance of these strategies, primarily focusing on the shape of the profit and loss (P\&L) profile. The table below shows the final results. \begin{table}[H] \caption{1,000 simulations with $\lambda=0.1$ w.r.t. $ \mathcal R_T$.} \begin{center} \begin{tabular}{lccc} \hline \cline{1-4} Statistics &Constant-Vol & Moving-Constant-Vol &Vol-adjust \\ \hline \cline{1-4} Mean &-300.70&-294.50&-288.46\\ Std&54.76&27.67&27.45\\ Skewness&1.03&0.23&0.24\\ Kurtosis&5.26&3.03&3.05\\ \hline \cline{1-4} Objective function &-577.58&-565.39&-560.36\\ \hline \cline{1-4} \end{tabular} \end{center} \label{t2} \end{table}% \begin{figure}[H] \begin{center} \includegraphics[width=5in, height=3.8in]{Costpdf.pdf} \caption{ The performance of the ``Constant-Vol" strategy on small- and medium-capitalization assets with their volatilities varying randomly through the day. } \label{fig6} \end{center} \end{figure} As we can see from Table \ref{t2}, the differences among these strategies are significant. Profiting from hedging the risk arising from stochastic volatility is possible. Compared with the ``Constant-Vol'' strategy, the results in the second column, other strategies adapting to the varying volatility, either ``Moving-Constant-Vol'' (the results in the third column) or ``Vol-adjust'' (the results in the fourth column), obtain a lower cost and a smaller risk. The more accurate the adaption, the lower the cost of trading and the smaller the risk of liquidating. The ``Constant-Vol" strategy is a successful one if the underlying assets are large-cap US stocks. However, it is not applicable for assets that are less heavily traded than large-cap US stocks, where volatilities are believed to vary randomly through the day. Figure \ref{fig6} displays the performance of the ``Constant-Vol" strategy on small- and medium-capitalization assets. As we can see from this figure, the distribution of its P \& L profile is significantly different from a normal distribution (which is the case for large-cap US stocks) with quite high levels of skewness and kurtosis due to the presence of stochastic volatility effect. Thus at least some adjustment has to be made to identify and capture the relevant characteristics for these assets.\\ \section{An Extension: The Incorporation of Limit Orders } Consider a trader, he/she may submit sell limit orders specifying the quantity and the indicated price per share, which will be executed only when incoming buy market orders are matching with the limit orders. Otherwise, he/she can post sell market orders for an immediate execution, but in this case he/she is going to obtain a less favorable execution price. Suppose that sell limit orders are submitted at the current best ask, and that the ask-bid spread $\Delta$ keeps unchanged throughout the trading horizon. Moreover, assume that other participants' buy market orders arrive at Poisson rate $ \lambda^M=A\exp(-\kappa\Delta)$, a decreasing function of $\Delta$ as in the literature. Let $ M_t$ count the numbers of buy market orders arriving at the system by time $t$. The number of shares held by the trader, $ X_t$, can then be expressed by ($N=1$) \begin{equation}\label{shares} X_t=X_0-M_t-\int_0^t\theta_udu,\quad X_0=Q. \end{equation} \begin{remark} Actually, controls belonging to $ \Theta_t$ need to satisfy one more condition: \begin{equation}\label{condition} \theta_t\Big|_{dM_t=1}=0, \end{equation} i.e., market order can only be used at times when sell limit orders are not executed. Since for any admissible control process, the integrands in the following two integrals $$ \int_0^t\mathbb I_{\{dM_u=0\}}\theta_udu\quad\mbox{and}\quad \int_0^t\theta_udu $$ differ at only finitely many times, and when we integrate with respect to the Lebesgue measure $ du$, these differences do not matter. Technically speaking, the set of all jumps of the integrand is a countable set, and hence, it has Lebesgue measure zero. Therefore, no mandatory requirement is made to satisfy condition (\ref{condition}) in a continuous-time liquidation setting. \end{remark} The advantage of market orders over limit orders is that they are less affected by averse price impacts. Suppose a company revises its growth estimates upwards. Those investors who observe piece of this news may react by submitting buy market orders, executed against sell limit orders in the limit order book that have not yet been withdrawn. All these executed sell limit orders may incur opportunity costs due to the adverse selection risk. Market orders, however, will involve unfavorable transaction fees--market impact. \\ \subsection{Adverse Selection} It is possible to incorporate the effect of adverse selection in the trading strategy by assuming that when other participants' market orders arrive, they may induce a jump in the stock price in the direction of the trade. Assume that the bid-price\footnote{Since sell market orders are executed against buy limit orders, $S_t^I$ is actually the best bid at time $t$.} satisfies the stochastic differential equation \begin{equation}\label{Mid} d S^I_t=\displaystyle dS_t+\eta^{u}dM_t-\eta^dd\hat M_t -\eta^{per}\theta_td t, \end{equation} where $ \{\hat M_t\}_{t\ge0}$ is a jump process satisfying \begin{equation}\label{po} \left\{ \begin{array}{lll} \mathbb P( d\hat M_t=1|dM_t=0)=\frac{\lambda^M\eta^u}{\eta^d} dt\\ \mathbb P( d\hat M_t=1|dM_t=1)=0, \end{array} \right. \end{equation} and $ \eta^u, \eta^d>0$ represent the jump in bid-price just after a market order event. Let us give a brief comment on Eqs. (\ref{Mid}) and (\ref{po}). If bid-side liquidity is found for the order at time $ t\in[0,T)$, i.e., $dM_t=1$, the expected execution price for the order at time $t$ moves up (by $ \eta^u$). If no liquidity is found on that side, i.e., $ dM_t=0$, the expected price move is supposed to be in the opposite direction. By Eq. (\ref{po}), \begin{equation}\label{po1} \begin{array}{lll} \displaystyle \mathbb P(d\hat M_t=1)&=& \displaystyle \mathbb P(d\hat M_t=1|d M_t=0)\times \mathbb P(dM_t=0)+\mathbb P(d\hat M_t=1|d M_t=1)\times \mathbb P(dM_t=1)\\ &=& \frac{\lambda^M\eta^u}{\eta^d} dt \times \mathbb P(dM_t=0)\\ &=& \frac{\lambda^M\eta^u}{\eta^d} dt+o(dt), \end{array} \end{equation} and $$ \mathbb P(d\hat M_t=0)= 1-\frac{\lambda^M\eta^u}{\eta^d} dt+o(dt). $$ Therefore, $$ \begin{array}{lll} \mathbb E_t[\Delta S_t^I]&=& \mathbb E_t[ S_t^I-S_{t-}^I]\\ &=& [\lambda^M dt+o(dt)]\times \eta^u -[1-\lambda^Mdt+o(dt)]\times [\frac{\lambda^M\eta^u}{\eta^d} dt+o(dt)]\times \eta^d \\ &=& o(dt). \end{array} $$ To evaluate the effect of the adverse selection on trader's trading strategy, we consider the simplest case: risk-neutral investors. Similar conclusions may also hold for risk-averse investors. \begin{itemize} \item For a market-order only trader who does not have any limit orders in the market and simply uses market orders to implement his/her liquidation mandate, we have $ \eta^u=\eta^d=0$. That is, the trader does not use limit orders to ``monitor" other participants' trading behavior, and believes that that stock price evolves over time according to $$ dS^I_t=dS_t-\eta^{per}\theta_tdt. $$ It is straightforward to verify that the solution to the associated HJB equation is $$ J(t, q)=\left(\frac{1}{\frac{2}{\eta^{per}-2K}-\frac{1}{\eta^{tem}}(T-t)}-\frac{\eta^{per}}{2}\right)q^2. $$ The optimal liquidating strategy can then be derived through the relations, \begin{equation}\label{ac} \begin{array}{ll} \displaystyle \theta^M_t=-\frac{1}{2\eta^{tem}}\left[\partial_q J+\eta^{per}q\right],\\ \displaystyle X_T=\frac{\eta^{tem}Q}{\eta^{tem}+T\left(2K-\eta^{per}\right)}.\\ \end{array} \end{equation} It is straightforward to verify that if $K$ satisfies Condition (\ref{cond1}), then $\theta^M_t\ge0, X_T\in(0,Q)$, and as $ K\to\infty$, $ X_T\to0$. \end{itemize} It is the simplest liquidating strategy in limit order books for impatient traders (recall the urgency for liquidation is the primary consideration). Now let us take limit orders into consideration. In addition to earning ask-bid spreads, limit orders can also be used to incorporate the effect of adverse selection on the trading strategy. \begin{itemize} \item For a risk-neutral trader who seeks to maximize expected terminal wealth with limit and market orders. Under some smoothness assumptions for a classical solution or the situation when a unique viscosity solution is considered, the value function $J$ without any constraints solves \begin{equation}\label{case2} \left\{ \begin{array}{lll} \displaystyle \partial_t J +\lambda^M \left[\eta^u+\Delta+J(t,q-1)-J(t,q)\right] + \frac{1}{4\eta^{tem}}\left[\partial_q J+\eta^{per}q\right]^2=0\\ \displaystyle J(T-, q)=-K q^2. \end{array} \right. \end{equation} We conjecture that the solution has the following form: $$ J(t,q)= f(t)+g(t)q+h(t)q^2. $$ Some simple calculations lead to: \begin{equation}\label{relation} \left\{ \begin{array}{lll} \displaystyle h(t)=\frac{1}{\frac{2}{\eta^{per}-2K}-\frac{1}{\eta^{tem}}(T-t)}-\frac{\eta^{per}}{2}\\ \displaystyle g(t)=-2\lambda^M\int_t^Te^{\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau}h(r)dr\\ \displaystyle f(t)=\int_t^T\Big[\lambda^M\big(\eta^u+\Delta+h(r)-g(r)\big) +\frac{g^2(r)}{4\eta^{tem}} \Big]dr, \end{array} \right. \end{equation} and \begin{equation}\label{appo} \displaystyle \theta^{M\&L}_t=\displaystyle -\frac{1}{2\eta^{tem}}\left[g(t)+\left(2h(t)+\eta^{per}\right)q\right]. \end{equation} Here $ \theta^{M\&L}_t$ denotes the optimal liquidating strategy with limit and market orders. \end{itemize} \quad\\ We remark that, \begin{description} \item[(1)] For any $ t\in[0,T)$, $ h(t)<0$, $ g(t)>0$. Therefore, $$ \theta^{M\&L}_t <\theta_t^M. $$ That is, realizing the information carried by other marker participants (as we assumed in our paper, it is monitored by the execution of limit orders), traders will slow down their liquidating speed so as to get profit from market momentum. \item[(2)] If condition (\ref{cond1}) is satisfied, $$ \begin{array}{lll} \displaystyle g(t)&=& \displaystyle -2\lambda^M\int_t^Te^{\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau}h(r)dr\\ &=& \displaystyle -2\eta^{tem} \lambda^M\bigg\{ \int_t^Te^{\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau}d\left(\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau\right)\\ && \displaystyle \quad\quad\quad \quad \quad-\frac{\eta^{per}}{2\eta^{tem}} \int_t^Te^{\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau}dr\bigg\} \\ &=& \displaystyle -2\eta^{tem}\lambda^M\left\{ e^{\frac{1}{2\eta^{tem}}\int_t^T(2h(\tau)+\eta^{per})d\tau} -1-\frac{\eta^{per}}{2\eta^{tem}} \int_t^Te^{\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau}dr\right\}. \end{array} $$ Combining the results in Eq. (\ref{relation}), we have $$ e^{\frac{1}{2\eta^{tem}}\int_t^r(2h(\tau)+\eta^{per})d\tau}=\frac{\frac{1}{2K-\eta^{per}}+\frac{1}{2\eta^{tem}}(T-r)}{\frac{1}{2K-\eta^{per}}+\frac{1}{2\eta^{tem}}(T-t)}. $$ If the time horizon $T$ satisfies $$ T\le \frac{2\eta^{tem}}{2K-\eta^{per}}\left[\sqrt{1+\frac{2(2K-\eta^{per})}{\eta^{per}}}-1\right], $$ then $ g(t)\le 2\eta^{tem}\lambda^M$ holds for any time $ t\in[0,T)$. That is, the reduced liquidation speed $\frac{g(t)}{2\eta^{tem}}$ is always smaller than the rate at which limit sell orders are executed. The overall expected liquidation speed has increased. \item[(3)] $$ \left\{ \begin{array}{lll} \displaystyle\mathbb E[\dot{X}_t]= -\lambda^M-\mathbb E[\theta^{M\&L}_t] = -\lambda^M+\frac{1}{2\eta^{tem}}\left[g(t)+\left(2h(t)+\eta^{per}\right)\mathbb E[X_t] \right]\\ \displaystyle X_0=Q, \end{array} \right. $$ which yields $$ \mathbb E[X_t] =Qe^{\frac{1}{2\eta^{tem}}\int_0^t(2h(\tau)+\eta^{per})d\tau}-\int_0^t \left[\lambda^M-\frac{1}{2\eta^{tem}}g(u)\right]e^{\frac{1}{2\eta^{tem}}\int_u^t(2h(\tau)+\eta^{per})d\tau}du. $$ The liquidation target $$ Q\ge \max_{t\in[0,T)} e^{-\frac{1}{2\eta^{tem}}\int_0^t(2h(\tau)+\eta^{per})d\tau} \left[\int_0^t\left[\lambda^M-\frac{1}{2\eta^{tem}}g(u)\right]e^{\frac{1}{2\eta^{tem}}\int_u^t(2h(\tau)+\eta^{per})d\tau}du-\frac{g(t)}{2h(t)+\eta^{per}}\right], $$ and we have $ \mathbb E[\theta_t^{M\&L}]\ge0$ and $ X_t\in[0,Q]$ for any time $ t\in[0,T)$. The obtained optimal strategy in Eq. (\ref{appo}) is actually the one we are looking for. \end{description} \quad\\ The estimation procedure for the parameter $ \lambda_M$ in Eq. (\ref{relation}) basically matches the intuition that one must count the number of executions at ask/bid and normalized this quantity by the length of the time horizon. For high-frequency data over $ [0,T]$, we have a consistent estimator of $ \lambda_M$, which is given by $$ \widehat{\lambda}_M=\frac{\mbox{\#\{ executions at ask/bid\}}}{T}. $$ \subsection{Numerical Results } Consider the situation in which stocks are traded in a single exchange, i.e., $ N=1$. As far as our simulation is concerned, we adopt the hypothetical values of the model parameters $ \lambda_M=100$, $\Delta=0.3$ and $ \eta^u=\eta^d=\eta^I=0.02$. Other parameters not listed here are assumed to be the same as those used in Section 3.2. The simulation is obtained through the following procedure: \begin{table}[H] \caption{ The procedure } \begin{center} \begin{tabular}{ll} \hline \cline{1-2} Step 0.&Set initial values at time $ t=0$;\\ Step 1.& Compute the trader's liquidating rate $ \theta_t$, given the state variables; \\ Step 2. & If a buy market order arrives at time $t$, then update the quantities by:\\ &\quad\quad$$ $ \left\{ \begin{array}{l} \displaystyle X_{t+dt}=X_t-1\\ \displaystyle C_{t+dt}=C_t+(S^I_t+\Delta)\\ \displaystyle S^I_{t+}=S^I_t+\eta^I \end{array} \right. $ $$\quad execute a trade using sell LOs;\\ & otherwise, \\ & \quad\quad$$ $ \left\{ \begin{array}{l} \displaystyle X_{t+dt}=X_t-\theta_tdt\\ \displaystyle C_{t+dt}=C_t+( S^I_t-\eta^{tem}\theta_t)\cdot\theta_tdt \end{array} \right. $ $$\quad execute a trade using sell MOs,\\ &and \\ &\quad $ S^I_{t+}=\left\{ \begin{array}{lll} \displaystyle S^I_t-\eta^I-\eta^{per}\theta_tdt, &&\mbox{with probability $ \lambda^M dt$,}\\ \displaystyle S^I_t-\eta^{per}\theta_tdt, &&\mbox{with probability $ 1-\lambda^M dt$.}\\ \end{array} \right. $\\ Step 3.& Update the affected price by a random increment $ \pm \sigma\sqrt{dt}$:\\ &\quad\quad\quad\;$ S^I_{t+dt}=\left\{ \begin{array}{lll} S^I_{t+}+\sigma \sqrt{dt},&&\mbox{with probability 0.5,}\\ S^I_{t+}-\sigma \sqrt{dt},&&\mbox{with probability 0.5;}\\ \end{array} \right. $\\ Step 4.& Let $ t=t+dt$. If $ t< T$, return to Step 1; otherwise, stop and exit.\\ \hline \cline{1-2} \end{tabular} \end{center} \label{default} \end{table}% \noindent Figures \ref{fig7} illustrates the dynamics of inventory and affected price for one simulation of a stock path. \begin{figure}[H] \begin{center} \includegraphics[width=5in, height=6.5in]{share1.pdf} \caption{ The dynamics of inventory and affected price: {(\bf Black}) Market orders only; {(\bf Blue}) Market \& limit. } \label{fig7} \end{center} \end{figure} We observe that investors without making use of limit orders would liquidate on a linear trajectory $ \theta=\frac{q}{a+b(T-t)}$ and receive a relatively lower execution price from a certain point in time during the trading horizon.\\ We then run 1000 simulations to investigate the effect of adverse selection on the trading strategy. The performances of strategies aware of this effect and strategies not aware of this effect are depicted in Figure \ref{fig9}. \begin{figure}[H] \begin{center} \includegraphics[width=5 in, height=3.4 in]{d1.pdf} \caption{ Profit \& Loss profile distributions. } \label{fig9} \end{center} \end{figure} The numerical simulations show that the strategy including the use of limit orders performs better than the market only strategy, it receives a significantly higher expected Profit \& Loss profile (from risk-neutral traders' point of view), say $294.9$ with both MOs and LOs and $-371.3$ with MOs only. \section{Conclusions} In this paper, we utilize a quantitative model to discuss the optimal liquidating problem in an illiquid market. Three different aspects of the optimal liquidation problem are discussed in our paper: (i) optimal liquidating strategy across multiple venues; (ii) optimal liquidating strategy with stochastic volatility (a special case, slow mean-reverting stochastic volatility, was discussed); and (iii) the incorporation of limit orders in optimal liquidation problem. We formulate the optimal liquidation problem with both temporary and permanent market impacts. Under the no arbitrage assumption, we propose a linear model to determine the equilibrium price in a competitive market with multiple trading venues. Multi-scale analysis method is used to discuss the case of stochastic volatility, where the stochastic volatility is driven by a slow-varying factor $ \nu_t$. As an extension of our model, we include the use of limit orders. These orders are submitted at the current best ask/bid with buy limit order not being executed. In addition to earn ask-bid spreads, these orders can be used to incorporate the effect of adverse selection on the trading strategy. In our model, the quantity of limit sell orders is assumed to be {\it one} share per trade. Whereas, in reality, it can be multiple shares per trade. This will be an interesting direction for future research. \section*{Appendix} \subsection*{Appendix A.} \begin{proof} In this appendix, we give the steps for deriving HJB-1. Using Eq. (\ref{R}), the objective at any time $t$ becomes $$ \begin{array}{lll} \displaystyle \mathbb E_t\bigg[\int_t^{T-}\sigma_uX_udW_u-\eta^{per} \int_t^{T-}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ \displaystyle\quad\quad -\int_t^{T-}[\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots +\eta^{tem,(N)}(\theta^{(N)} _u)^2]du -\lambda\int_t^{T-}\sigma_u^2X_u^2du -X_{T-}\mathcal C^o(X_{T-})\bigg]. \end{array} $$ Thus, given initial value $ X_t=q$, $$ \begin{array}{lll} J(t,s, \sigma, q)&=&\displaystyle \max_{\boldsymbol\theta(\cdot)\in \Theta_t}\; \mathbb E^{t,s,\sigma, q}_{\boldsymbol\theta(\cdot)}\bigg[\int_t^{T-}\sigma_uX_udW_u-\eta^{per}\int_t^{T-}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ \displaystyle &&-\int_t^{T-}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]- \lambda \sigma_u^2X_u^2\right)du-X_{T-}\mathcal C^o(X_{T-})\bigg]\\ &=&\displaystyle \max_{\boldsymbol\theta(\cdot)\in\Theta_t}\; \mathbb E^{t,s,\sigma, q}_{\boldsymbol\theta(\cdot)}\Bigg[\int_t^{t+\Delta t}\sigma_uX_udW_u-\eta^{per}\int_t^{t+\Delta t}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ && \displaystyle -\int_t^{t+\Delta t}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du\\ && \displaystyle +\mathbb E^{t+\Delta t,s+\Delta s, \sigma+\Delta\sigma, q+\Delta q}_{\boldsymbol\theta(\cdot)} \bigg[\int_{t+\Delta t}^{T-}\sigma_uX_udW_u-\eta^{per}\int_{t+\Delta t}^{T-}X_u(\beta^1_u\theta^{(1)}_u+\cdots+\beta^N_u\theta^{(N)} _u)du\\ && \displaystyle -\int_{t+\Delta t}^{T-}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du -X_{T-}\mathcal C^o(X_{T-})\bigg]\Bigg],\\ \end{array} $$ where $\mathbb E^{t,s,\sigma, q}_{\boldsymbol\theta(\cdot)}[\cdot]$ is the conditional expectation conditioned on the control process $ \boldsymbol\theta(\cdot)$ and the initial state $ (S_t, \sigma_t, X_t)=(s,\sigma,q)$. Notice that for any control process $ \boldsymbol\theta(\cdot)\in\Theta_t$, we have $$ \begin{array}{lll} && \displaystyle \mathbb E^{t+\Delta t, s+\Delta s, \sigma+\Delta \sigma, q+\Delta q}_{\boldsymbol\theta(\cdot)} \bigg[\int_{t+\Delta t}^{T-}\sigma_uX_udW_u-\eta^{per}\int_{t+\Delta t}^{T-}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ && \displaystyle -\int_{t+\Delta t}^{T-}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du-X_{T-}\mathcal C^o(X_{T-})\bigg]\\ &=& \displaystyle U(t+\Delta t, s+\Delta s, \sigma+\Delta \sigma, q+\Delta q, \boldsymbol\theta(\cdot))\\ &\le & \displaystyle \max_{\boldsymbol\theta(\cdot)\in\Theta_{t+\Delta t}} U(t+\Delta t, s+\Delta s, \sigma+\Delta \sigma, q+\Delta q, \boldsymbol\theta(\cdot))\\ &=& \displaystyle J(t+\Delta t, s+\Delta s, \sigma+\Delta \sigma, q+\Delta q). \end{array} $$ Thus, we have \begin{equation}\label{le} \begin{array}{lll} J(t,s,\sigma, q) &\le& \displaystyle \max_{\boldsymbol\theta(\cdot)\in\Theta_t}\; \mathbb E^{t,s,\sigma, q}_{\boldsymbol\theta(\cdot)}\Big[J(t+\Delta t,s+\Delta s, \sigma+\Delta\sigma, q+\Delta q)+\int_t^{t+\Delta t}\sigma_uX_udW_u\\ && \displaystyle -\eta^{per}\int_t^{t+\Delta t}X_u(\beta^{(1)}_u\theta^{(1)}_u +\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ && \displaystyle-\int_t^{t+\Delta t}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du\Big].\\ \end{array} \end{equation} Let $ \tilde{\boldsymbol\theta}^*(\cdot)$ be the optimal control over $ [t+\Delta t, T)$, modifying the optimal control between t and $ t+\Delta t$ by an arbitrary control $ \boldsymbol\theta(\cdot)$, i.e. $ \boldsymbol\theta^\prime_u=\boldsymbol\theta\mathbb I_{\{u\le t+\Delta t\}}+\tilde{\boldsymbol\theta}^*_{u}\mathbb I_{\{u> t+\Delta t\}}$. Then, we obtain $$ \begin{array}{lll} J(t,s,\sigma, q) &\ge& J_{ \boldsymbol\theta^\prime(\cdot)}(t,s,\sigma, q)\\ &=& \displaystyle \mathbb E^{t,s, \sigma, q}_{\boldsymbol\theta(\cdot)}\Big[J(t+\Delta t, s+\Delta s, \sigma+\Delta \sigma, q+\Delta q)+\int_t^{t+\Delta t}\sigma_uX_udW_u\\ && \displaystyle -\eta^{per}\int_t^{t+\Delta t}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ && \displaystyle -\int_t^{t+\Delta t}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du\Big].\\ \end{array} $$ Hence, we obtain \begin{equation}\label{ge} \begin{array}{lll} J(t,s,\sigma, q) &\ge& \displaystyle \max_{\boldsymbol\theta(\cdot)\in\Theta_t}\; \mathbb E^{t,s,\sigma, q}_{\boldsymbol\theta(\cdot)}\Big[J(t+\Delta t,s+\Delta s, \sigma+\Delta\sigma, q+\Delta q)+\int_t^{t+\Delta t}\sigma_uX_udW_u\\ && \displaystyle -\eta^{per}\int_t^{t+\Delta t}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ && \displaystyle -\int_t^{t+\Delta t}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du\Big].\\ \end{array} \end{equation} Putting both inequalities, Eq.(\ref{le}) and Eq.(\ref{ge}), together, we arrive at the Dynamic Programming Principle (DPP): $$ \begin{array}{lll} J(t,s,\sigma, q)&=& \displaystyle \max_{\boldsymbol\theta(\cdot)\in\Theta_t}\; \mathbb E^{t,s,\sigma, q}_{\boldsymbol\theta(\cdot)}\Big[J(t+\Delta t, s+\Delta s, \sigma+\Delta\sigma, q+\Delta q)+\int_t^{t+\Delta t}\sigma_uX_udW_u\\ && \displaystyle -\eta^{per}\int_t^{t+\Delta t}X_u(\beta^{(1)}_u\theta^{(1)}_u+\cdots+\beta^{(N)}_u\theta^{(N)} _u)du\\ && \displaystyle -\int_t^{t+\Delta t}\left([\eta^{tem,(1)}(\theta^{(1)}_u)^2+\cdots+\eta^{tem,(N)}(\theta^{(N)} _u)^2]-\lambda \sigma_u^2X_u^2\right)du\Big]. \end{array} $$ Let $ \boldsymbol\theta^*(\cdot)$ be the optimal control over $ [t, T)$, then we have $$ \boldsymbol\theta^*_{t,s,\sigma, q}(t^\prime,s^\prime, \sigma^\prime, q^\prime)=\tilde{\boldsymbol\theta}_{t+\Delta t,s+\Delta s, \sigma+\Delta \sigma, q+\Delta q}^*(t^\prime, s^\prime, \sigma^\prime, q^\prime), \quad\mbox{for any \; $ t^\prime>t+\Delta t$}. $$ The dynamic programming equation, HJB equation, is the infinitesimal version of this principle. \end{proof} \subsection*{Appendix B.} \begin{proof} In the case of (1) $N=1$; or (2) the trading venues have the same market efficiency, $\Delta_N=\lambda \sigma^2>0$. Hence, $$ h(t;\sigma)=\sqrt{\frac{\Delta_N}{a}}\cdot \frac{\varsigma e^{-2\sqrt{a\Delta_N}(T-t)}-1}{\varsigma e^{-2\sqrt{a\Delta_N}(T-t)}+1}-\frac{\eta^{per}}{2N}. $$ Under the assumption that $ K>\displaystyle \frac{b}{2a}+\sqrt{\frac{|\Delta_N|}{a}}$, $ -1<\varsigma<0$. Therefore, $h$ is a decreasing function in $t$, and for any $ n=1,\cdots, N$ \begin{equation}\label{proof1} \begin{array}{lll} 2h(t;\sigma)+\eta^{per}\beta^{(n)}&=& \displaystyle \sqrt{\frac{\Delta_N}{a}}\cdot \frac{\varsigma e^{-2\sqrt{a\Delta_N}(T-t)}-1}{\varsigma e^{-2\sqrt{a\Delta_N}(T-t)}+1} \le0. \end{array} \end{equation} Recall that $$ \left\{ \begin{array}{l} \displaystyle \dot{X}_t=-\sum_{n=1}^N \theta^{n,*}_t\\ \displaystyle \theta^{n,*}_t =-\frac{1}{2\eta^{tem}} \left[2h(t;\sigma)+ \eta^{per}\beta^{(n)}\right]X_t. \end{array} \right. $$ Hence, $X_t$ satisfies the following first-order ODE: $$ \left\{ \begin{array}{l} \displaystyle \dot{X}_t=\left( \sum_{n=1}^N\frac{1}{2\eta^{tem}} \left[2h(t;\sigma) + \eta^{per}\beta^{(n)}\right]\right)X_t\\ \displaystyle X_0=Q, \end{array} \right. $$ which yields \begin{equation}\label{proof2} X_t=Q\cdot \exp\left(\int_0^t\sum_{n=1}^N\frac{1}{2\eta^{tem}} \left[2h(u;\sigma)+ \eta^{per}\beta^{(n)}\right]du \right) \in (0, Q],\quad \mbox{for} \ t \in [0,T). \end{equation} Combining the results in Eq. (\ref{proof1}) and Eq. (\ref{proof2}), we conclude that, for any $n=1,\ldots, N$, $$ \theta^{n,*}_t=-\frac{1}{2\eta^{tem}} \left[2h(t;\sigma)+ \eta^{per}\beta^{(n)}\right]X_t\ge 0, $$ and $$ \int_0^T\sum_{n=1}^N\theta^{n,*}_tdt=X_0-X_T =Q\left[ 1-\exp\left(\int_0^t\sum_{n=1}^N\frac{1}{2\eta^{tem}} \left[2h(u;\sigma)+ \eta^{per}\beta^{(n)}\right]du\right) \right]\le Q. $$ \end{proof} \section*{Acknowledgements} This research work was supported by Research Grants Council of Hong Kong under Grant Number 17301214, HKU Strategic Theme on Computation and Information and National Natural Science Foundation of China Under Grant number 11671158.
cb3281b5135975a1b90a8c3247e88eeceb6505a0
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \begin{quotation} \textit{``One Picture is Worth a Thousand Words''}\\ \hspace*{\fill} [An advertisement for the \textit{San Antonio Light} (1918)] \end{quotation} Teaching undergraduate \textit{Analysis of Algorithms} has been a rewarding, although a bit taxing, experience. I was often surprised to learn that many basic problems that clearly belong to its core syllabus had been left unanswered or partially answered. Also, it seemed a bit odd to me that many otherwise decent texts offered unnecessarily imprecise computations\footnote{A notable exception in this category is \cite{sedfla:ana}.} of several rather fundamental results. \medskip In this article, I pursue a seemingly marginal topic, the best-case behavior of a well-known sorting algorithm $ {\tt MergeSort} $, which pursuit, however, yields some interesting findings that could hardly be characterized as \textit{``marginal.''} It turns out that - contrary to what a casual student of this subject might believe - computing the exact formula for the number of comparisons of keys that $ {\tt MergeSort} $ performs on any $ n $-element array in the best case is not a routine exercise and leads to a problem that gained some notoriety for being a hard nut to crack analytically: the \textit{sum of digits} problem. Even more unexpectedly, a relatively straightforward\footnote{Although not quite \textit{closed-form}.} formula for the said number of comparisons yields an improvement of a well-known answer to this instance of the \textit{sum of digits} problem: \begin{quote} How many $ 1 $s appear in binary representations of all integers between (but not including) $ 0 $ and $ n $? \end{quote} \section{$ {\tt MergeSort} $ and its best-case behavior} \label{sec:mergsor} A call to $ {\tt MergeSort} $ inherits an $ n $-element array $ {\tt A} $ of integers and sorts it non-decreasingly, following the steps described below. \begin{algMergeSort} \label{alg:mersor} To sort an $ n $-element array $ {\tt A} $ do: \begin{enumerate} \item \label{alg:mersor:item1} If $ n \leq 1 $ then return $ {\tt A} $ to the caller, \item \label{alg:mersor:item2} If $ n \geq 2 $ then \begin{enumerate} \item \label{alg:mersor:item2:it1} pass the first $ \lfloor \frac{n}{2} \rfloor $ elements of $ {\tt A} $ to a recursive call to $ {\tt MergeSort} $, \item pass the last $ \lceil \frac{n}{2} \rceil $ elements of $ {\tt A} $ to another recursive call to $ {\tt MergeSort} $, \item \label{alg:mersor:item2:it2} linearly merge, by means of a call to $ {\tt Merge} $, the non-decreasingly sorted arrays that were returned from those calls onto one non-decreasingly sorted array $ {\tt A}^{\prime} $, \item \label{alg:mersor:item2:it3} return $ {\tt A}^{\prime} $ to the caller. \end{enumerate} \end{enumerate} \end{algMergeSort} A Java code of $ {\tt Merge} $ is shown on the Figure~\ref{fig:Merge}. \begin{figure}[h] \centering \includegraphics[width=0.7\linewidth]{./Merge1} \includegraphics[width=0.7\linewidth]{./Merge2} \caption{A Java code of $ {\tt Merge} $, based on a pseudo-code from \cite{baa:ana}. Calls to $ {\tt Boolean} $ method $ {\tt Bcnt.incr()} $ count the number of comps.} \label{fig:Merge} \end{figure} \medskip A typical measure of the running time of $ {\tt MergeSort} $ is the number of \textit{comparisons of keys}, which for brevity I call \textit{comps}, that it performs while sorting array $ {\tt A} $. Since no comps are performed outside $ {\tt Merge} $, the running time of $ {\tt MergeSort} $ can be computed as the sum of numbers of comps performed by all calls to $ {\tt Merge} $ during the execution of $ {\tt MergeSort} $. Since the minimum number of comps performed by $ {\tt Merge} $ on two list is equal to the length of the shorter list, and any increasingly sorted array on any size $ N \geq 2 $ produces only best-case scenarios for all subsequent calls to $ {\tt Merge} $, a rudimentary analysis of the recursion tree for $ {\tt MergeSort} $ easily yields the exact formula for the minimum number of comps for the entire $ {\tt MergeSort} $. The problem arises when one tries to reduce the said formula, which naturally involves long summations, to one that can be evaluated in a logarithmic time. \subsection{Recursion tree} \label{subsecsec:rectre} The obvious recursion tree for $ {\tt MergeSort} $ and sufficiently large $ n $ is shown on Figure~\ref{fig:rectre}. \begin{figure} [h] \includegraphics[scale=.15]{RecTreeMergeSort4.JPG} \caption{\label{fig:rectre} A sketch of the recursion 2-tree $ T $ for $ {\tt MergeSort} $ for a sufficiently large $ n $, with levels shown on the left and the numbers of nodes shown on the right. The nodes correspond to calls to $ {\tt MergeSort} $ and show sizes of (sub)arrays passed to those calls. The last level is $ h $; it only contains nodes with value 1. The root corresponds to the original call to $ {\tt MergeSort} $. If a call that is represented by a node $ p $ executes further recursive call to $ {\tt MergeSort} $ then these calls are represented by the children of $ p $; otherwise $ p $ is a leaf.} \end{figure} A recursive application of the equality\footnote{It can be verified separately for odd and even values of $ n $.} \begin{equation} \label{eq:n/2=n+1/2} \lceil \frac{n}{2}\rceil = \lfloor \frac{n+1}{2}\rfloor \end{equation} allows for rewriting of that tree onto one whose first four levels are shown on Figure~\ref{fig:rectre1}. \begin{figure} [h] $ \,\,\,\,\,\,\,\, \,\,\,\,\,\,\,\, \,\,\,\,\,\,\,\, $\includegraphics[width=.75\linewidth]{RecursionTreeSimpleNarrowerColored} \caption{\label{fig:rectre1} The first four levels of the recursion 2-tree $ T $ from Figure~\ref{fig:rectre}, with the equality (\ref{eq:n/2=n+1/2}) applied, recursively. The number of comparisons of keys performed in the best case by $ {\tt Merge} $ invoked in step~\ref{alg:mersor:item2:it2} of Algorithm~\ref{alg:mersor} as a result to a call to $ {\tt MergeSort} $ corresponding to a node of $ T $ is equal to the number that is shown in its left child, highlighted yellow. All the right children at level $ k \geq 1 $ of $ T $ show numbers of the form $ \lfloor \frac{n+i}{2^{k+1}} \rfloor $, where $ i \geq 2^k $. Thus all the left children (highlighted yellow) at level $ k \geq 1$ show numbers of the form $ \lfloor \frac{n+i}{2^{k+1}} \rfloor $, where $ i < 2^k $.} \end{figure} \subsection{Best-case and its characterization $ B(n) $} \label{subsec:bestchar} The best-case arrays of sizes $ \lfloor \frac{n}{2} \rfloor $ and $ \lceil \frac{n}{2} \rceil $ for $ {\tt Merge} $, where $ n \geq 2 $, are those in which every element of the first array is less than all elements of the second one. In such a case, $ {\tt MergeSort} $ performs $ \lfloor \frac{n}{2} \rfloor $ of comps. \medskip \medskip Thus the following recurrence relation for the least number $ B(n) $ of comparisons of keys that $ {\tt MergeSort} $ performs on any $ n $-element array is straightforward to derive from its description given by Algorithm~\ref{alg:mersor}. \begin{equation} \label{eq:rec_rel1} B(1) = 0, \end{equation} and, for $ n \geq 2 $, \begin{equation} \label{eq:rec_rel2} B(n) = \lfloor \frac{n}{2}\rfloor + B( \lfloor \frac{n}{2}\rfloor) + B( \lceil \frac{n}{2}\rceil). \end{equation} Using the equality (\ref{eq:n/2=n+1/2}), the recurrence relation (\ref{eq:rec_rel2}) is equivalent to: \begin{equation} \label{eq:rec_rel3} B(n) = \lfloor \frac{n}{2}\rfloor + B( \lfloor \frac{n}{2}\rfloor) + B( \lfloor \frac{n+1}{2}\rfloor) . \end{equation} A graph of $ B(n) $ is shown on Figure~\ref{fig:Best-case_rec_solution}. \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Best-case_MergeSort_old_rec_solution} \caption{Graph of the solution $ B(n) $ of the recurrence \eqref{eq:rec_rel1} and \eqref{eq:rec_rel2}.} \label{fig:Best-case_rec_solution} \end{figure} \medskip Unfolding the recurrence (\ref{eq:rec_rel3}) allows for noticing that the minimum number $ B(n) $ of comps performed by all calls to $ {\tt Merge} $ is equal to the sum of all values shown at nodes highlighted yellow in the recursion tree $ T $ of Figure~\ref{fig:rectre1}. They may be summed-up level-by-level. One can notice from Figure~\ref{fig:rectre1} that the number of comps performed at any level $ k $ with the maximal number $ 2^k $ of nodes is given by this formula: \begin{equation} \label{eq:formula_level_i} \sum _{i=0} ^{2^k-1} \lfloor \frac{n+i}{2^{k+1}} \rfloor . \end{equation} What is not clear is whether all levels of the recursion tree $ T $ are maximal. Fortunately, the answer to this question does not depend on whether given instance of $ {\tt MergeSort} $ is running on a best-case array or on any other case of array. It has been known form a classic analysis of the worst-case running time of $ {\tt MergeSort} $ that every level of its recursion tree $ T $ that contains at least one non-leaf, or - in other words - a node that shows value $ p \geq 2 $, is maximal. \ref{sec:mergesort} page~\pageref{sec:mergesort} contains a detailed derivation of that fact. Thus all levels 0 through $ h-1 $ of $ T $ are maximal. Therefore, the formula (\ref{eq:formula_level_i}) gives the number of comps for every level $ 0 \leq k \leq h-1 $. \medskip The last level $ h $ of $ T $ may be not maximal because the level $ h-1 $ may contain leaves, or - in other words - nodes that show value $ p = 1 $, where $ p = \lfloor \frac{n+i}{2^{h-1}} \rfloor $ for some $ 0 \leq i \leq 2^{h-1} - 1 $, and as such do not have any children in level $ h $. However, for each such node the value of $ \lfloor \frac{p}{2} \rfloor = \lfloor \frac{n+i}{2^{h}} \rfloor $ is 0, so it can be included in summation (\ref{eq:formula_level_i}) without affecting its value even though the said value does not correspond to any node in level $ h $. Therefore, the formula (\ref{eq:formula_level_i}) gives the number of comps for level $ k = h $. \medskip Also, the depth of $ T $ is $ \lfloor \lg n \rfloor $, as the Theorem~\ref{thm:depthRectree} page \pageref{thm:depthRectree} in \ref{sec:excerpts} states. Thus the minimum number of comps performed by $ {\tt MergeSort} $ is given by this formula: \begin{equation} \label{eq:formula1} \sum _{k=0} ^{\lfloor \lg n \rfloor} \sum _{i=0} ^{2^k-1} \lfloor \frac{n+i}{2^{k+1}} \rfloor . \end{equation} Unfortunately, the summation~(\ref{eq:formula1}) contains $ n - 1 $ non-zero terms, so it cannot be evaluated quickly in its present form. Fortunately, its inner summation (\ref{eq:formula_level_i}) can be reduced to a closed-form formula. \subsection{\textit{Zigzag} function} \label{subsec:zig} In order to reduce (\ref{eq:formula_level_i}) to a closed form, I am going to use function \textit{Zigzag} defined by: \begin{equation} \label{eq:def_zigzag} \mbox{\textit{Zigzag}}\,(n) = \min (x - \lfloor x \rfloor, \lceil x \rceil - x) . \end{equation} The following fact is instrumental for that purpose. \medskip \begin{thm} \label{thm:formula_level_Zigzag} For every natural number n and every positive natural number m, \begin{equation} \label{eq:formula_level_Zigzag} \sum _{i=m} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor - \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = 2m \times \mbox{\textit{Zigzag}}\,(\frac{n}{2m}), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{thm} \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Takagi_function_zigzag_basic_half} \caption{Graph of function $ \mbox{\textit{Zigzag}} (x) $ $ = $ $ \min (x - \lfloor x \rfloor, \lceil x \rceil - x) $.} \label{fig:zigzag} \end{figure} \begin{proof} The equality (\ref{eq:formula_level_Zigzag}) can be verified experimentally, for instance, with a help of software for symbolic computation\footnote{\label{foot:Mat} I used Wolfram Mathematica for that purpose.}. The analytic proof is deferred to Section~\ref{sec:proof_formula_level_Zigzag}. \end{proof} \medskip \begin{corollary} \label{cor:formula_level_Zigzag_1} For every natural number n and every positive natural number m, \begin{equation} \label{eq:formula_level_Zigzag_1} \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = \frac{n}{2} - m \times \mbox{\textit{Zigzag}}\,(\frac{n}{2m}), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{corollary} \begin{proof} First, let's note\footnote{Analytic proof of that fact is a straightforward exercise; see Appendix~\ref{sec:proof1} page \pageref{sec:proof1}.} that \begin{equation} \label{eq:100} \sum _{i=0} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor = n. \end{equation} From (\ref{eq:100}) I conclude \begin{equation} \label{eq:110} \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor + \sum _{i=m} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor = n. \end{equation} Solving equations (\ref{eq:formula_level_Zigzag}) and (\ref{eq:110}) for $ \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor $ yields (\ref{eq:formula_level_Zigzag_1}). \end{proof} Here is the closed-form of the summation (\ref{eq:formula_level_i}). \begin{corollary} \label{cor:formula_level_Zigzag_2} For every natural number n and every natural number k, \begin{equation} \label{eq:formula_level_Zigzag_2} \sum _{i=0} ^{2^k-1} \lfloor \frac{n+i}{2^{k+1}} \rfloor = \frac{n}{2} - 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{corollary} \begin{proof} Substitute $ m = 2^k $ in (\ref{eq:formula_level_Zigzag_1}). \end{proof} The following theorem yields the formula (\ref{eq:formula_Zigzag*}) for the minimum number $ B(n) $ of comps performed by $ {\tt MergeSort} $. \begin{thm} \label{thm:formula_Zigzag} For every natural number n, \begin{equation} \label{eq:formula_Zigzag*} \sum _{k=0} ^{\lfloor \lg n \rfloor} \sum _{i=0} ^{2^k-1} \lfloor \frac{n+i}{2^{k+1}} \rfloor = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{thm} \begin{proof} \[\sum _{k=0} ^{\lfloor \lg n \rfloor} \sum _{i=0} ^{2^k-1} \lfloor \frac{n+i}{2^{k+1}} \rfloor = \sum _{k=0} ^{\lfloor \lg n \rfloor}( \frac{n}{2} - 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}})) = \] \[ = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}) .\] \end{proof} \medskip \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Best-case_MergeSort_both_formulas} \caption{Graphs of functions $ \sum _{k=0} ^{\lfloor \lg n \rfloor} \sum _{i=0} ^{2^k-1} \lfloor \frac{n+i}{2^{k+1}} \rfloor $ (bottom line) and $ \frac{n}{2}(\lfloor \lg n \rfloor + 1) - \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}) $ (top line) of equality \eqref{eq:formula_Zigzag*}. They coincide with each other for all natural numbers $ n $.} \label{fig:Best-case_MergeSort_worksheet_best-case} \end{figure} Formula (\ref{eq:formula_Zigzag*}), although not quite closed-form, comprises of summation with only $ \lfloor \lg n \rfloor +1 $ closed-form terms, so it may be evaluated in $ O(\log ^{c}) $ time, where $ c $ is a constant. I will show in Section~\ref{sec:frac} that (\ref{eq:formula_Zigzag*}) does not have a closed form. Graphs of both sides of equality (\ref{eq:formula_Zigzag*}) are shown on Figure~\ref{fig:Best-case_MergeSort_worksheet_best-case}. Once can see that for natural numbers $ n $ they coincide with the solution $ B(n) $ of recurrences \eqref{eq:rec_rel1} and \eqref{eq:rec_rel2} visualized on Figure~\ref{fig:Best-case_rec_solution}. \begin{cor} \label{thm:formula_Zigzag_B} For every natural number n, the minimum number $ B(n) $ of comps that $ {\tt MergeSort} $ performs while sorting an $ n $-element array is: \begin{equation} \label{eq:formula_Zigzag_B} B(n) = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{cor} \section{A fractal in $ B(n) $} \label{sec:frac} A deceitfully simple expression \begin{equation} \label{eq:sum_zigzag} \sum _{k=0} ^{\lfloor \lg x \rfloor} 2^{k+1} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k+1}}), \end{equation} half of which occurs in formula (\ref{eq:formula_Zigzag_B}) of Corollary~\ref{thm:formula_Zigzag_B}, is a formidable adversary for those who may try to turn it into a closed form, although the time required for its evaluation for any given $ n $ is $ O (\log ^ c) $ \footnote{So, to all practical purposes, (\ref{eq:formula_Zigzag_B}) is a closed-form formula.}. That does not come as a surprise, taking into account that its graph, shown on Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag}, bears a resemblance of fractal. This can be easily seen as soon as a sawtooth function $ 2^{\lfloor \lg x \rfloor + 1} - x $ is subtracted from it, yielding the function $ F(x) $ given by \begin{equation} \label{eq:def_F} F(x) = \sum _{k=0} ^{\lfloor \lg x \rfloor} 2^{k+1} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k+1}}) - 2^{\lfloor \lg x \rfloor + 1} + x. \end{equation} \medskip \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Best-case_MergeSort_worksheet_sum_zig_with_chainsaw} \caption{A graph of function $\sum _{k=0} ^{\lfloor \lg x \rfloor} 2^{k+1} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k+1}}) $ plotted against a sawtooth function $ 2^{\lfloor \lg x \rfloor + 1} - x $.} \label{fig:Best-case_MergeSort_worksheet_sum_zigzag} \end{figure} \noindent Since $ \frac{1}{2} \leq \frac{x}{2^{\lfloor \lg x \rfloor + 1}} < 1 $, equality \eqref{eq:def_zigzag} implies \begin{equation} \nonumber \mbox{\textit{Zigzag}}\,(\frac{x}{2^{\lfloor \lg x \rfloor+1}}) = 1 - \frac{x}{2^{\lfloor \lg x \rfloor+1}}, \end{equation} or \begin{equation} \label{eq:zigprop1} 2^{\lfloor \lg x \rfloor+1} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{\lfloor \lg x \rfloor+1}}) = 2^{\lfloor \lg x \rfloor+1} - x. \end{equation} The equality \eqref{eq:zigprop1} simplifies definition \eqref{eq:def_F} of function $ F $ to \begin{equation} \label{eq:fractal} F(x) = \sum _{k=1} ^{\lfloor \lg x \rfloor} 2^{k} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k}}), \end{equation} visualized on Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw}. \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Best-case_MergeSort_zigzags_and_sum_and_top_lines} \caption{A graph of function $ F(x) = \sum _{k=1} ^{\lfloor \lg x \rfloor} 2^{k} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k}})$ plotted below its tight linear upper bound $ y = \frac{x-1}{2} $ (if can be shown that $ F(x) = \frac{x-1}{2} $ whenever $ x $ $ = $ $ \frac{1}{3}(2^{k+1} + (-1)^k) $ for some integer $ k \geq 0 $); also shown below $ F(x) $ are the terms $ 2^{k} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k}}) $ of the summation and their tight linear upper bound $ y = \frac{x}{3} $.} \label{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw} \end{figure} \medskip The function $ F $ is a fractal with quasi similarity that repeats at intervals of exponentially growing length. It is a union \begin{equation} \label{eq:F_union_def} F = \bigcup _{k=0} ^{\infty} f_k \end{equation} of functions $ f_k $, each having an interval $ [ 2^k, 2^{k+1}) $ as its domain. In other words, for every integer $ k \geq 0 $, \begin{equation} \label{eq:def_f_k} f_k = F \restriction [ 2^k, 2^{k+1}), \end{equation} which, of course, yields \eqref{eq:F_union_def}. \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Takagi_function_progression_5_with_upper_and_terms_of_sum} \caption{A graph of the first six (the first one is $ 0 $) normalized parts of function of Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw} plotted against the line $ y = \sum _{i=0} ^{\infty} \frac{1}{2^{2i+1}} = \frac{2}{3} $. Also shown (in blue) are the first five terms $ \frac{1}{2^i} \mbox{\textit{Zigzag}} \, (2^i x) $, $ i = 0,...,4 $, of sums that occur in the formula \eqref{eq:f_k_simplified} for $ \tilde{f}_k(x) $; for each integer $ n $ and all $ x \in [n, n+1) $, their parts above the $ X $-axis restricted to $ [n, n+1) $ visualize a fragment of an infinite binary search \textit{trie} $ T $ defined as the set of shortest binary expansions of $ x - \lfloor x \rfloor $ with the last digit $ 1 $ (if the said binary expansion is finite) being interpreted as the sequence terminator; in particular, the root of $ T $ is $ .1 $, and if $ {\bf a} $ is a finite binary sequence then the children of binary expansion $ .{\bf a}1 $ are $ .{\bf a}01 $ and $ .{\bf a}11 $ .} \label{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw_progression} \end{figure} Let $ \hat{f}_k $ be the normalized $ f_k $ on interval $ [0,1) $, defined by: \begin{equation} \label{eq:f_k_normalized} \hat{f}_k (x) = \frac{1}{2^k} f_k (2^k (x + 1)) , \end{equation} and $ \tilde{f}_k $ be the periodized $ \hat{f}_k $ by composing it with a sawtooth function $ x- \lfloor x \rfloor $,\footnote{The fractional part of $ x $.} defined by: \begin{equation} \label{eq:f_k_periodized} \tilde{f}_k(x) = \hat{f}_k(x- \lfloor x \rfloor). \end{equation} Contracting definitions \eqref{eq:def_f_k}, \eqref{eq:f_k_normalized}, and \eqref{eq:f_k_periodized}, yields \begin{equation} \label{eq:f_k_simplified_a} \tilde{f}_k(x) = \frac{1}{2^k} F (2^k (x- \lfloor x \rfloor + 1)). \end{equation} One can compute\footnote{An elementary geometric argument based on the graph visualized on Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw_progression} will do.} from \eqref{eq:f_k_simplified_a} the following alternative formula for $ \tilde{f}_k(x) $: \begin{equation} \label{eq:f_k_simplified} \tilde{f}_k(x) = \sum _{i=0} ^{k-1} \frac{1}{2^i} \mbox{\textit{Zigzag}} \, (2^i x). \end{equation} Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw_progression} shows functions $ \tilde{f}_0,...,\tilde{f}_6 $ drawn on the same graph. \medskip Since each function $ f_k $, and - therefore - each function $ \hat{f}_k $, and - therefore - each function $ \tilde{f}_k $, are a result of smaller and smaller triangles piled, originating in function \textit{Zigzag} of definition (\ref{eq:fractal}) of function $ F $, on one another as shown on Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw_progression}, for any integers $ 0 \leq i < j $, $ \tilde{f}_i $ linearly interpolates $ \tilde{f}_j $. Because of that, each $ \tilde{f}_i $ linearly interpolates the limit $ \tilde{F} $ of all $ \tilde{f}_k $s defined by: \begin{equation} \label{eq:def_lim} \tilde{F}(x) = \lim _{k \rightarrow \infty} \tilde{f}_k (x), \end{equation} as Figure~\ref{fig:Takagi_function_progression_5} illustrates. An application of \eqref{eq:f_k_simplified} to \eqref{eq:def_lim} yields: \begin{equation} \label{eq:def_lim2} \tilde{F}(x) = \sum _{i=0} ^{\infty} \frac{1}{2^i} \mbox{\textit{Zigzag}} \, (2^i x). \end{equation} \begin{figure}[h] \centering \includegraphics[width=0.7\linewidth]{./Takagi_function_progression_5_with_upper} \caption{Functions functions $ \tilde{f}_0 (x),\tilde{f}_1 (x), ... $ and their limit (the topmost curve) $ \tilde{F} (x) $ given by (\ref{eq:def_lim2}). Collapsing the \textit{Zigzag}$ (x) $ would yield the same, albeit scaled-down (by the factor of 2) pattern $ \frac{1}{2}\tilde{F}(2x) $, as (\ref{eq:def_lim2}) does imply.} \label{fig:Takagi_function_progression_5} \end{figure} Since for every integer $ n $ and $ i \geq k $, $ 2^i \frac{n}{2^k} $ is integer, $ \mbox{\textit{Zigzag}} \, (2^i \frac{n}{2^k}) = 0 $. Therefore, by virtue of \eqref{eq:f_k_simplified} and \eqref{eq:def_lim2}, for every non-negative integer $ k $ and $ n $, \begin{equation} \label{eq:Ftilde=ftilde_k} \tilde{F}(\frac{n}{2^k}) = \tilde{f}_k(\frac{n}{2^k}). \end{equation} This and \eqref{eq:f_k_simplified} eliminate the need for infinite summation\footnote{As it appears in \eqref{eq:def_lim2}} while computing $ \tilde{F}(\frac{n}{2^k}) $. \medskip It can be shown that although a continuous function, $ \tilde{F} $ is nowhere-differentiable. As such, it does not have a closed-form formula as any closed-form formula on a real interval must define a function have a derivative at every point of that interval, except for a non-dense set of its points. Since $ \tilde{F} $ can be expressed in function, described by a closed-form formula, of the right-hand side of formula \eqref{eq:formula_Zigzag*}, the latter does not have a closed-form formula, either. \begin{thm} \label{thm:no_closed_form} There is no closed-form formula $ \varphi(n) $ the values of which coincide with $ \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}), $ for all positive integers $ n $, that is, for every closed-form formula $ \varphi(n) $ on function \textit{Zigzag} there is a positive $ n $ such that \begin{equation} \label{eq:no_closed_form} \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}) \neq \varphi(n), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{thm} \begin{proof} Follows from the above discussion. A more detailed proof is deferred to Section~\ref{sec:proof_no_closed_form}. \end{proof} \medskip This way I arrived at the following conclusion. \medskip \begin{cor} \label{cor:no_closed_form} There is no closed-form formula for $ B(n) $. \end{cor} \begin{proof} A closed-form formula for $ B(n) $ would, by virtue of \eqref{eq:formula_Zigzag_B} page~\pageref{eq:formula_Zigzag_B}, yield a closed-form formula for $ \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}) $, which by Theorem~\ref{thm:no_closed_form} does not exist. \end{proof} \medskip \textbf{\textit{Note}}. One can apply the reverse transformations to those used in Section~\ref{sec:frac} on function $ \tilde{F} $ and construct a fractal function $ \breve{F} $, shown on Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zig_minus_chainsaw_w_upper}, given by the equation \begin{figure}[h] \centering \includegraphics[width=0.7\linewidth]{./Best-case_MergeSort_worksheet_sum_zig_minus_chainsaw_w_upper} \caption{A graph of function $ \breve{F} (x) = 2^{\lfloor \lg x \rfloor } \tilde{F}(\frac{x}{2^{\lfloor \lg x \rfloor }}) $ plotted above a graph of the function $ F(x) = \sum _{k=1} ^{\lfloor \lg x \rfloor} 2^{k} \mbox{\textit{Zigzag}}\,(\frac{x}{2^{k}}) $. } \label{fig:Best-case_MergeSort_worksheet_sum_zig_minus_chainsaw_w_upper} \end{figure} \begin{equation} \label{eq:def_F_u} \breve{F} (x) = 2^{\lfloor \lg x \rfloor } \tilde{F}(\frac{x}{2^{\lfloor \lg x \rfloor}}), \end{equation} that for every positive integer $ n $ satisfies \begin{equation} \label{eq:F_u=F} \breve{F} (n) = F(n), \end{equation} where $ F $ is given by \eqref{eq:fractal}. \section{Computing $ \tilde{F} (x) $ and $ B(n) $ from one another } \label{sec:compfromfrac} Computing values of function $ \tilde{F} (x) $ does not have to be as complex as (or more complex than) the definition \eqref{eq:def_lim2} implies. Of course, for every integer $ n $, $ \tilde{F} (n) $ $ = $ $ 0 $. One can apply some elementary arguments based on a structure visualized on Figure~\ref{fig:Takagi_function_progression_5} to conclude that \begin{equation} \tilde{F} (\frac{2}{3}) = \tilde{F} (\frac{1}{3}) = \frac{2}{3} , \end{equation} (the latter being the maximum of $ \tilde{F} (x) $) or that for every positive integer $ k $, \begin{equation} \tilde{F} (\frac{1}{2^k}) = \frac{ k }{2^k} . \end{equation} It takes a bit more work to compute \begin{equation} \tilde{F} (\frac{3}{2^k}) = \frac{3 k - 4 }{2^k} . \end{equation} \medskip It turns out that computing values of function $ \tilde{F} (x) $ for every $ x $ that has a finite binary expansion can be done easily if an oracle for computing the values of the function $ B(n) $ defined by \eqref{eq:rec_rel1} and \eqref{eq:rec_rel2} is given\footnote{Which is not that surprising after a glance at Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zig_minus_chainsaw_w_upper}.}. Once that is accomplished, since $ \tilde{F} (x) $ is a continuous function and the set of numbers with finite binary expansions is dense in the set $ \mathfrak{R} $ of reals, it allows for fast approximations of $ \tilde{F} (x) $ for every real $ x $. \footnote{It helps to remember that $ \tilde{F} $ is a periodic function with $ \tilde{F} (x) = \tilde{F} (x - \lfloor x \rfloor) $.} \medskip \begin{thm} \label{thm:mainB} For every positive integer $ n $ \footnote{Of course, one if free to assume that $ n $ is odd here.} and integer $ k $ with $ n \leq 2^k $, \begin{equation} \label{eq:mainB} \tilde{F} (\frac{n}{2^k}) = \frac{n \times k - 2 B(n)}{2^k} . \end{equation} \end{thm} \begin{proof} The equality \eqref{eq:mainB} can be verified experimentally, for instance, with a help of software for symbolic computation$ ^{\ref{foot:Mat}} $. The analytic proof is deferred to Section~\ref{sec:comment}. \end{proof} Theorem~\ref{thm:mainB} allows for easy computing of $ B(n) $ if $ \tilde{F} (\frac{n}{2^k}) $ is given for some $ k \geq \lg n $ using this form of \eqref{eq:mainB}: \begin{cor} \label{cor:mainB2} For every positive integer $ n $ and integer $ k $ with $ n \leq 2^k $, \begin{equation} \label{eq:mainB2} B(n) = \frac{n \times k }{2} - 2^{k-1} \tilde{F} (\frac{n}{2^k}). \end{equation} \end{cor} \begin{proof} An obvious conclusion from \eqref{eq:mainB}. \end{proof} For instance, putting $ k = \lfloor \lg n \rfloor + 1 $ in \eqref{eq:mainB2} easily yields \eqref{eq:formula_Zigzag_B}. For $ k = \lceil \lg n \rceil $ we obtain \[ B(n) = \frac{n \lceil \lg n \rceil }{2} - 2^{\lceil \lg n \rceil-1} \tilde{F} (\frac{n}{2^{\lceil \lg n \rceil}}) = \] [by \eqref{eq:def_lim2}] \[ = \frac{n \lceil \lg n \rceil }{2} - 2^{\lceil \lg n \rceil-1} \sum _{i=0} ^{\infty} \frac{1}{2^i} \mbox{\textit{Zigzag}} \, (2^i \frac{n}{2^{\lceil \lg n \rceil}}) = \] [since for $ i \geq \lceil \lg n \rceil $, $ 2^i \frac{n}{2^{\lceil \lg n \rceil}} $ is integer and $ \mbox{\textit{Zigzag}} \, (2^i \frac{n}{2^{\lceil \lg n \rceil}}) = 0 $] \[ = \frac{n \lceil \lg n \rceil }{2} - 2^{\lceil \lg n \rceil-1} \sum _{i=0} ^{\lceil \lg n \rceil - 1} \frac{1}{2^i} \mbox{\textit{Zigzag}} \, (2^i \frac{n}{2^{\lceil \lg n \rceil}}) = \] \[ = \frac{n \lceil \lg n \rceil }{2} - \frac{1}{2} \sum _{i=0} ^{\lceil \lg n \rceil - 1} 2^{\lceil \lg n \rceil - i} \mbox{\textit{Zigzag}} \, ( \frac{n}{2^{\lceil \lg n \rceil - i}}) .\] Substituting $ k $ for $ \lceil \lg n \rceil - i $ we conclude \begin{equation} \label{eq:mainAlt} B(n) = \frac{n \lceil \lg n \rceil }{2} - \frac{1}{2} \sum _{k=1} ^{\lceil \lg n \rceil} 2^{k} \mbox{\textit{Zigzag}} \, ( \frac{n}{2^{k}}), \end{equation} a similar to \eqref{eq:formula_Zigzag_B} characterization of $ B(n) $. \section{Relationship between the best case and the worst case} A casual student of $ {\tt MergeSort} $ tends to believe that its worst-case behavior is about twice as bad as its best-case behavior. This, of course, is only approximately true. In this Section, I will derive the exact difference between $ 2 B(n) $ and $ W(n) $ using function $ F $ defined by \eqref{eq:def_F} page \pageref{eq:def_F}. \medskip An exact formula for the number $ W(n) $ of comparisons of keys performed by $ {\tt MergeSort} $ in the worst case is known\footnote{See \eqref{eq:recmergesort900} in the \ref{sec:excerpts}.} and is given for any positive integer $ n $ by the following equality: \begin{equation} \label{eq_worst} W(n) = \sum _{i = 1} ^{n} \lceil \lg i \rceil . \end{equation} From \eqref{eq:formula_Zigzag_B} and \eqref{eq:def_F}, one can derive \[ 2B(n) = n \lfloor \lg n \rfloor - 2^{\lfloor \lg n \rfloor + 1} + 2n - F(n) = \] [by $ \sum _{i = 1} ^{n} \lceil \lg i \rceil = n \lfloor \lg n \rfloor - 2^{\lfloor \lg n \rfloor + 1} + n + 1 $ from \cite{knu:art}] \[ = \sum _{i = 1} ^{n} \lceil \lg i \rceil - 1 + n - F(n) = \] [by \eqref{eq_worst}] \[ = W(n) - 1 + n - F(n) . \] The above yield the following characterization. \begin{thm} \label{thm:dif2BW} For every positive integer $ n $, the difference between twice the number $ B(n) $ of comparison of keys performed in the best case and the number $ W(n) $ of comparison of keys performed in the worst case by $ {\tt MergeSort} $ while sorting an $ n $-element array is: \begin{equation} \label{eq:dif2BW} 2B(n) - W(n) = n - 1 - F(n) , \end{equation} where $ F(n) $, visualized on Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw}, is given by \eqref{eq:fractal}. \end{thm} \begin{proof} Follows from the above discussion. \end{proof} In particular, since for every positive integer $ n $, \begin{equation} \label{eq:F_ub} 0 \leq F(n) \leq \frac{n-1}{2} \end{equation} (see Figure~\ref{fig:Best-case_MergeSort_worksheet_sum_zigzag_minus_chainsaw} for explanation), I conclude with the following tight linear bounds on $ 2B(n) - W(n) $. \begin{cor} For every positive integer $ n $, the difference between twice the minimum number $ B(n) $ and the maximum number $ W(n) $ of comparison of keys performed in the worst case by $ {\tt MergeSort} $ while sorting an $ n $-element array satisfies this inequality: \begin{equation} \frac{n-1}{2} \leq 2B(n) - W(n) \leq n-1 . \end{equation} \end{cor} \begin{proof} Follows from \eqref{eq:dif2BW} and \eqref{eq:F_ub}. \end{proof} \begin{figure}[h] \centering \includegraphics[width=0.7\linewidth]{Best-case_MergeSort_bounds_2B-W} \caption{A graph of $ 2B(n) - W(n) $ shown between graphs of its tight linear bounds $ n - 1 $ and $ \frac{n - 1}{2} $.} \label{fig:2B-W_with_bounds} \end{figure} Obviously, $ 2B(n) - W(n) = n - 1$ whenever $ F(n) = 0 $, that is, whenever $ n = 2^{\lfloor \lg n \rfloor} $. It can be shown that $ 2B(n) - W(n) = \frac{n-1}{2} $ whenever $ n $ $ = $ $ \frac{1}{3}(2^{k+1} + (-1)^k) $ for some integer $ k \geq 0 $. \medskip A graph of $ 2B(n) - W(n) $ and its tight bounds are shown on Figure~\ref{fig:2B-W_with_bounds}. \section{The sum of digits problem} \label{sec:sumdig} A known explicit formula, published in \cite{tro:digsum}, for the total number of bits in all integers between 0 and $ n $ (not including 0 and $ n $) is expressed in terms of function \textit{Zigzag} (referred to as $ 2 g $ in \cite{tro:digsum}) and is given by:\footnote{The following are screen shots and an excerpt from \cite{tro:digsum}.} {\centering \includegraphics[width=1\linewidth]{Quote_from_Trollope_A}} \noindent $ \,\, $ Let $ g(x) $ be periodic of period 1 and defined on $ [0,1] $ by {\centering \includegraphics[width=1\linewidth]{Quote_from_Trollope}} \medskip It has been shown in \cite{mci:numones} that the recurrence relation for $ A(n,2) $ is the same as the recurrence relation for $ B(n) $ given by (\ref{eq:rec_rel1}) and (\ref{eq:rec_rel2}). Therefore, the formula (\ref{eq:formula_Zigzag*}) derived in this paper is equivalent to $ A(n,2) $ given above by the considerably more complicated definition. Interestingly, the above definition can be simplified to \eqref{eq:formula_Zigzag*} along the lines of the elementary derivation of the alternative formula \eqref{eq:mainAlt} for $ B(n) $ on page~\pageref{eq:mainAlt}\footnote{Even more interestingly, if someone did bother to simplify Trollope's formula of \cite{tro:digsum} then I am not aware of it.}. \section{Proof of Theorem~\ref{thm:formula_level_Zigzag} page~\pageref{thm:formula_level_Zigzag}, Subsection~\ref{subsec:zig}} \label{sec:proof_formula_level_Zigzag} In this Section, I provide an analytic proof of the experimentally-derived Theorem~\ref{thm:formula_level_Zigzag} page~\pageref{thm:formula_level_Zigzag}, Subsection~\ref{subsec:zig} that was instrumental for the derivation of a logarithmic-length formula\footnote{$ B(n) = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}) $, where $ \mbox{\textit{Zigzag}}\,(x) = \min (x - \lfloor x \rfloor, \lceil x \rceil - x) $.} for $ B(n) $. The result and its proof have a flavor of Concrete Mathematics. Although they are interesting in their own right, they cannot be found in \cite{knu:concrete}. \begin{thm} \label{thm:formula_level_Zigzag_App} {\rm (Same as Theorem~\ref{thm:formula_level_Zigzag}.)} For every natural number n and every positive natural number m, \begin{equation} \label{eq:formula_level_Zigzag_App} \sum _{i=m} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor - \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = 2m \times \mbox{\textit{Zigzag}}\,(\frac{n}{2m}), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag} page~\pageref{fig:zigzag}. \end{thm} \begin{proof} First, let's note that \[ \sum _{i=m} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor - \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = \sum _{i=0} ^{m-1} \lfloor \frac{n+i+m}{2m} \rfloor - \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = \sum _{i=0} ^{m-1} (\lfloor \frac{n+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{n+i}{2m} \rfloor) , \] that is, \begin{equation} \label{eq:formula_level_Zigzag_App_var} \sum _{i=m} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor - \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = \sum _{i=0} ^{m-1} (\lfloor \frac{n+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{n+i}{2m} \rfloor). \end{equation} Let \begin{equation} \label{eq:def_r} n = k \times 2 m + r , \end{equation}where $ 0 \leq r < 2m $, and let $ 0 \leq i < m $. We have \[ \lfloor \frac{n+i}{2m} + \frac{1}{2} \rfloor = \lfloor \frac{k \times 2 m + r+i}{2m} + \frac{1}{2} \rfloor = k + \lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor \] and \[ \lfloor \frac{n+i}{2m} \rfloor = \lfloor \frac{k \times 2 m + r+i}{2m} \rfloor = k + \lfloor \frac{r+i}{2m} \rfloor .\] Thus, by virtue of \eqref{eq:formula_level_Zigzag_App_var}, \begin{equation} \label{eq:formula_level_Zigzag_App_var2} \sum _{i=m} ^{2m-1} \lfloor \frac{n+i}{2m} \rfloor - \sum _{i=0} ^{m-1} \lfloor \frac{n+i}{2m} \rfloor = \sum _{i=0} ^{m-1} (\lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{r+i}{2m} \rfloor) . \end{equation} We have \begin{equation} \label{eq:dif=0_or_1} \lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{r+i}{2m} \rfloor = \left\{ \begin{array}{ll} 1 \mbox{ if } \; \frac{1}{2} \leq \frac{r+i}{2m} < 1 \\ \\ 0 \mbox{ otherwise,} \end{array} \right. \end{equation} because $ \frac{r+i}{2m} + \frac{1}{2} < \frac{3m}{2m} + \frac{1}{2} = 2 $ so that $ \lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor \leq 1 $ and, therefore, $ \lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{r+i}{2m} \rfloor \leq 1 $. \medskip Let $ I $ be defined as \begin{equation} \label{eq:defI} I = \{ i \in \mathbb{N} \mid \frac{1}{2} \leq \frac{r+i}{2m} < 1 \} = \{ i \in \mathbb{N} \mid m-r \leq i < 2m-r \} . \end{equation} By virtue of \eqref{eq:dif=0_or_1}, we have \begin{equation} \label{eq:formula_level_Zigzag_App_var3} \sum _{i=0} ^{m-1} (\lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{r+i}{2m} \rfloor) = \sum _{i \in I} (\lfloor \frac{r+i}{2m} + \frac{1}{2} \rfloor - \lfloor \frac{r+i}{2m} \rfloor) = \sum _{i \in I} 1 = \#{I}, \end{equation} where $ \# (I) $ denotes the cardinality of $ I $. \medskip If $ r \leq m $ then, by \eqref{eq:defI}, $ \# I = m - (m-r) = r $. If $ r > m $ then, by \eqref{eq:defI}, $ \# I = 2 m - r $. In any case, \begin{equation} \nonumber \# I = \min (r, 2m - r) = 2m \min (\frac{r}{2m}, 1- \frac{r}{2m}) = \end{equation} [since $ 0 \leq \frac{r}{2m} < 1 $ so that $ \lfloor \frac{r+i}{2m} \rfloor = 0 $ and $ \lceil \frac{r+i}{2m} \rceil = 1 $] \[ = 2m \min (\frac{r}{2m} - \lfloor \frac{r}{2m} \rfloor, \lceil \frac{r}{2m} \rceil- \frac{r}{2m}) = \] [by the definition \eqref{eq:def_zigzag} of function {\em Zigzag}] \[ = 2m \times \mbox{\em Zigzag} \, ( \frac{r}{2m}) =\] [since {\em Zigzag} is a periodic function with period 1] \[= 2m \times \mbox{\em Zigzag} \, ( k+ \frac{r}{2m}) = 2m \times \mbox{\em Zigzag} \, ( \frac{k \times 2m + r}{2m}) =\] [by \eqref{eq:def_r}] \[= 2m \times \mbox{\em Zigzag} \, ( \frac{n}{2m}) .\] Thus \begin{equation} \label{eq:cardI=2mZ} \# I = 2m \times \mbox{\em Zigzag} \, ( \frac{n}{2m}) . \end{equation} From \eqref{eq:formula_level_Zigzag_App_var}, \eqref{eq:formula_level_Zigzag_App_var2}, \eqref{eq:formula_level_Zigzag_App_var3}, and \eqref{eq:cardI=2mZ}, I conclude~\eqref{eq:formula_level_Zigzag_App}. \end{proof} \section{Proof of Theorem~\ref{thm:no_closed_form} page~\pageref{thm:no_closed_form}, Section~\ref{sec:frac} \label{sec:proof_no_closed_form}} In this Section, I present a brief discussion/motivation of what can be generally considered a \textit{closed-form formula} for a function from the set of real numbers into a set of real numbers. I provide an analytic proof of Theorem~\ref{thm:no_closed_form} page~\pageref{thm:no_closed_form}, Section~\ref{sec:frac} that implies the non-existence of closed-form formula for the minimum number $ B(n) $ of comparisons of keys by $ {\tt MergeSort} $ while sorting an $ n $-element array. I am going to use the acronym $ cf\!f $ as an abbreviation for \textit{closed-form formula}. \medskip \medskip For reader's convenience, the Theorem~\ref{thm:no_closed_form} is quoted below as Theorem~\ref{thm:no_closed_form_add}. \begin{thm} \label{thm:no_closed_form_add} {\rm (Same as Theorem~\ref{thm:no_closed_form}.)} There is no $ cf\!f $ $ \varphi(n) $ the values of which coincide with $ \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}), $ for all positive integers $ n $, that is, for every $ cf\!f $ $ \varphi(n) $ on function \textit{Zigzag} there is a positive $ n $ such that \begin{equation} \label{eq:no_closed_form_add} \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^k \mbox{\textit{Zigzag}}\,(\frac{n}{2^{k+1}}) \neq \varphi(n), \end{equation} where \textit{Zigzag} is a function defined by \eqref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag}. \end{thm} The rest of this Section constitutes the proof of Theorem~\ref{thm:no_closed_form_add}. \medskip First, let me use an example of function $ 2^x : \mathbb{R} \longrightarrow \mathbb{R} $ as an insight of what may be accepted as a $cf\!f$ for a \textit{continuous} function - like, say, $ \tilde{F}(x) $ - on the set $ \mathbb{R} $ of reals or on an interval thereof. One picks a dense\footnote{In the metric topology of $ \mathbb{R} $.} subset $ \mathbb{Q} $ of $ \mathbb{R} $, with a collection of mappings $ \rho_x(i) : \mathbb{N} \longrightarrow \mathbb{Q} $, where $ x \in \mathbb{R} $, given by $ \rho_x(i) = \frac{\lfloor i \times x \rfloor}{i} $ so that $ \lim _{i \rightarrow \infty}\rho_x(i) = x $. Since for any $ x \in \mathbb{R} \setminus \mathbb{Q} $, $ 2^x $ has been defined as \begin{equation} 2^x = \lim _{i \rightarrow \infty} 2^{\rho_x (i)} = \lim _{i \rightarrow \infty} \sqrt[i]{2^{\lfloor i \times x \rfloor}} , \end{equation} $ \lim _{i \rightarrow \infty} \sqrt[i]{2^{\lfloor i \times x \rfloor}} $ is considered a $cf\!f$ $ \alpha : \mathbb{R} \longrightarrow \mathbb{R} $ for $ 2^x $. \medskip \medskip \begin{lem} \label{lem:cff_F_tilde} For every positive integer $ n $, \begin{equation} \label{eq:proof5} \tilde{F} (\frac{n}{2^{\lfloor \lg n \rfloor}}) = \frac{n ( \lfloor \lg n \rfloor + 2) - 2 B(n)}{2^{\lfloor \lg n \rfloor}} -2, \end{equation} where the function $F$ has been defined by the equality~\eqref{eq:def_F} page~\pageref{eq:def_F}, the function $\tilde{F} $, visualized on Figure~\ref{fig:Tak} \footnote{Also, together with its partial sums, on Figure~\ref{fig:Takagi_function_progression_5}, page~\pageref{fig:Takagi_function_progression_5}.}, has been defined by the equality~\eqref{eq:def_lim2} page~\pageref{eq:def_lim2}, and the function $ B(n) $ has been defined by the equations \eqref{eq:rec_rel1} and \eqref{eq:rec_rel3} page~\pageref{eq:rec_rel1}. \end{lem} \begin{proof} \begin{figure} \centering \includegraphics[width=0.7\linewidth]{Takagi_function} \caption{A graph of the Blancmange function $ \tilde{F}(x) = \sum _{i=0} ^{\infty} \frac{1}{2^i} Zigzag (2^i x) $. \label{fig:Tak}} \end{figure} From \eqref{eq:def_F} page~\pageref{eq:def_F}, I compute \[ \sum _{i=0} ^{\lfloor \lg n \rfloor} 2^{i+1} \mbox{\textit{Zigzag}}\,(\frac{n}{2^{i+1}}) = F(n) + 2^{\lfloor \lg n \rfloor + 1} - n ,\] that is, \begin{equation} \label{eq:proof1} \sum _{i=0} ^{\lfloor \lg n \rfloor} 2^{i} \mbox{\textit{Zigzag}}\,(\frac{n}{2^{i+1}}) = \frac{1}{2} F(n) + 2^{\lfloor \lg n \rfloor} - \frac{n}{2} . \end{equation} Applying \eqref{eq:proof1} to the equality \eqref{eq:formula_Zigzag_B} page~\pageref{eq:formula_Zigzag_B}, I conclude \[ B(n) = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - (\frac{1}{2} F(n) + 2^{\lfloor \lg n \rfloor} - \frac{n}{2}) , \] or \[ B(n) = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - \frac{1}{2} F(n) - 2^{\lfloor \lg n \rfloor} + \frac{n}{2} , \] that is, \[ \frac{1}{2} F(n) = \frac{n}{2}(\lfloor \lg n \rfloor + 1) - B(n) - 2^{\lfloor \lg n \rfloor} + \frac{n}{2} , \] or \begin{equation} \label{eq:proof2} F(n) = n (\lfloor \lg n \rfloor + 2) - 2 B(n) - 2^{\lfloor \lg n \rfloor+1} . \end{equation} On the other hand, by virtue of \eqref{eq:fractal} page~\pageref{eq:fractal}, \[ F(n) = \sum _{i=1} ^{\lfloor \lg n \rfloor} 2^{i} \mbox{\textit{Zigzag}}\,(\frac{n}{2^{i}}) = \] [putting $ j = \lfloor \lg n \rfloor - i $] \[ = \sum _{j=0} ^{\lfloor \lg n \rfloor-1} 2^{\lfloor \lg n \rfloor-j} \mbox{\textit{Zigzag}}\,(\frac{n}{2^{\lfloor \lg n \rfloor-j}}) = 2^{\lfloor \lg n \rfloor} \sum _{j=0} ^{\lfloor \lg n \rfloor-1} \frac{1}{2^j} \mbox{\textit{Zigzag}}\,(\frac{2^j n}{2^{\lfloor \lg n \rfloor}}) =\] [since for $ j \geq \lfloor \lg n \rfloor $, $ \frac{2^j n}{2^{\lfloor \lg n \rfloor}} \in \mathbb{N} $ so that $ \mbox{\textit{Zigzag}}\,(\frac{2^j n}{2^{\lfloor \lg n \rfloor}}) = 0 $] \[ = 2^{\lfloor \lg n \rfloor} \sum _{j=0} ^{\infty} \frac{1}{2^j} \mbox{\textit{Zigzag}}\,(\frac{2^j n}{2^{\lfloor \lg n \rfloor}}) = \] [by \eqref{eq:def_lim2} page~\pageref{eq:def_lim2}] \[ 2^{\lfloor \lg n \rfloor} \tilde{F} (\frac{n}{2^{\lfloor \lg n \rfloor}}) . \] Thus, \begin{equation} \label{eq:proof3} F(n) = 2^{\lfloor \lg n \rfloor} \tilde{F} (\frac{n}{2^{\lfloor \lg n \rfloor}}) \end{equation} or \begin{equation} \label{eq:proof4} \tilde{F} (\frac{n}{2^{\lfloor \lg n \rfloor}}) = \frac{1}{2^{\lfloor \lg n \rfloor}} F(n) . \end{equation} Combining equalities \eqref{eq:proof2} and \eqref{eq:proof4} yields \[ \tilde{F} (\frac{n}{2^{\lfloor \lg n \rfloor}}) = \frac{1}{2^{\lfloor \lg n \rfloor}} ( n (\lfloor \lg n \rfloor + 2) - 2 B(n) - 2^{\lfloor \lg n \rfloor+1}) , \] or \eqref{eq:proof5}. \end{proof} \begin{lem} \label{lem:main_no_cff} If the function $ B(n) $ defined by the equations \eqref{eq:rec_rel1} and \eqref{eq:rec_rel3} has a $ cf\!f $ $ \beta : \mathbb{N} \longrightarrow \mathbb{N} $ then the function $ \tilde{F}(x) $ defined by the equation \eqref{eq:def_lim2} page~\pageref{eq:def_lim2} has a $ cf\!f $ $ \varphi : [1,2) \longrightarrow [0,\frac{2}{3}] $. \end{lem} \begin{proof} Let \begin{equation} \label{eq:def_D} D = \{ \frac{n}{2^{\lfloor \lg n \rfloor}} \mid n \in \mathbb{N} \} \end{equation} be the set of rationals in the interval $ [1,2) $ with finite binary expansions\footnote{It is a trivial exercise to show that every real number with finite binary expansion in the interval $ [1,2) $ is of the form $ \frac{n}{2^{\lfloor \lg n \rfloor}} $ for some $ n \in \mathbb{N} $, and it is obvious that every real number of that form has a finite binary expansion and falls into that interval.}, enumerated by $ \nu (n) : \mathbb{N} \longrightarrow D $ given by $ \nu (n) = \frac{n}{2^{\lfloor \lg n \rfloor}} $ and visualized on Figure~\ref{fig:mapping}. \begin{figure} [h] \centering \includegraphics[width=0.7\linewidth]{Verifications_for_best-case_MergeSort_iterator_of_D} \caption{A graph of enumeration $ \nu (n) = \frac{n}{2^{\lfloor \lg n \rfloor}} $ of the set $ D $.} \label{fig:mapping} \end{figure} \medskip $ D $ is a dense subset of the interval $ [1,2) $ of reals. Indeed, if $ x \in [1,2) $ then for every $ n \in \mathbb{N} $, $ \frac{\lfloor 2^n x \rfloor}{2^n} \in D $ and \begin{equation} \label{eq:convD} \lim _{n \rightarrow \infty} \frac{\lfloor 2^n x \rfloor}{2^n} = x. \end{equation} Hence, for any $ x \in [1,2) $, putting \begin{equation} \label{eq:n=floor} n = \lfloor 2^i x \rfloor , \end{equation} so that \[ {\lfloor \lg n \rfloor} = {\lfloor \lg \lfloor 2^i x \rfloor \rfloor} = {\lfloor \lg 2^i x \rfloor} = { \lfloor i+ \lg x \rfloor} ={i+ \lfloor \lg x \rfloor} = i \] [the last equality holds because $ 1 \leq x < 2 $ so that $ 0 \leq \lg x < 1 $ and $ \lfloor \lg x \rfloor = 0 $], or \begin{equation} \label{eq:lg=i} {\lfloor \lg n \rfloor} = i , \end{equation} we conclude, by virtue of \eqref{eq:convD}, \begin{equation} \nonumber \tilde{F}(x) = \tilde{F}(\lim _{i \rightarrow \infty} \frac{\lfloor 2^i x \rfloor}{2^i}) = \end{equation} [by the continuity of $ \tilde{F}(x) $] \[ = \lim _{i \rightarrow \infty} \tilde{F}( \frac{\lfloor 2^i x \rfloor}{2^i}) = \] [by the equality \eqref{eq:proof5} of Lemma~\ref{lem:cff_F_tilde}] \[ = \lim _{i \rightarrow \infty} \frac{ \lfloor 2^i x \rfloor (i + 2) - 2 B(\lfloor 2^i x \rfloor)}{2^{i}} - 2 = \lim _{i \rightarrow \infty} \frac{i \lfloor 2^i x \rfloor - 2 B(\lfloor 2^i x \rfloor)}{2^{i}} + \lim _{i \rightarrow \infty} 2\frac{ \lfloor 2^i x \rfloor }{2^{i}} - 2 =\] [by the equality \eqref{eq:convD}] \[ \lim _{i \rightarrow \infty} \frac{i \lfloor 2^i x \rfloor - 2 B(\lfloor 2^i x \rfloor)}{2^{i}} + 2x - 2. \] Thus, for any $ x \in [1,2) $, \begin{equation} \label{eq:proof6} \tilde{F}(x) = \lim _{i \rightarrow \infty} \frac{i \lfloor 2^i x \rfloor - 2 B(\lfloor 2^i x \rfloor)}{2^{i}} + 2x - 2 . \end{equation} The equality \eqref{eq:proof6} shows that if there is a $ cf\!f $ $ \beta : \mathbb{N} \longrightarrow \mathbb{N} $ for function $ B $ defined by the equations \eqref{eq:rec_rel1} and \eqref{eq:rec_rel3} page~\pageref{eq:rec_rel1} then there is a $ cf\!f $ $ \varphi : [1,2) \longrightarrow [0,\frac{2}{3}] $ given by \[ \lim _{i \rightarrow \infty} \frac{i \lfloor 2^i x \rfloor - 2 \beta(\lfloor 2^i x \rfloor)}{2^{i}} + 2x - 2 \] for $ \tilde{F}(x) $. This completes the proof of Lemma~\ref{lem:main_no_cff}. \end{proof} \medskip Should the nowhere-differentiable function $ \tilde{F} $ have a $ cf\!f $, it would be differentiable everywhere except, perhaps, on a non-dense subset of $ \mathbb{R} $. The following inductive argument demonstrates that. All atomic $ cf\!f $s are differentiable except, perhaps, on a non-dense subset of $ \mathbb{R} $. If a finite number of $ cf\!f $s are differentiable except, perhaps, on non-dense subsets of $ \mathbb{R} $ then their composition is differentiable except, perhaps, on non-dense subsets of $ \mathbb{R} $. \footnote{For instance, function $ \sqrt{x}Zigzag(\frac{1}{x}) $ is differentiable on $ [0,1] $, except for the non-dense set $ \{0\} \cup \{ \frac{1}{n} \mid n \in \mathbb{N} \} $.} Thus $ \tilde{F} $ has no $ cf\!f $. \medskip The above observation, together with Lemma~\ref{lem:main_no_cff}, complete the proof of Theorem~\ref{thm:no_closed_form_add}. \section{Proof of Theorem~\ref{thm:mainB} page~\pageref{thm:mainB}, Section~\ref{sec:compfromfrac} \label{sec:comment}} In this Section, I provide an analytic proof of experimentally-derived Theorem~\ref{thm:mainB} page~\pageref{thm:mainB}, Section~\ref{sec:compfromfrac}. This result, re-stated by Theorem~\ref{thm:mainB_add} below, allows for practically efficient computations of values of the continuous Blackmange function for reals with finite binary floating-point representations. I also provide some properties (Lammas~\ref{lem:100}, \ref{lem:200}, and \ref{lem:300}) of the \textit{Zigzag} function, given by the equality \eqref{eq:def_zigzag} page~\pageref{eq:def_zigzag} and visualized on Figure~\ref{fig:zigzag} page~\pageref{fig:zigzag}, that are useful for a neat derivation of a formula for the Blancmange function as (the limit of) a finite sum of some values of the \textit{Zigzag} function. \medskip Let the function\footnote{Known as the Blancmange function.} $\tilde{F} $, visualized on Figure~\ref{fig:Tak} page~\pageref{fig:Tak}, be defined by \eqref{eq:def_lim2} page~\pageref{eq:def_lim2}, and $ B(n) $, given by \eqref{eq:rec_rel1} and \eqref{eq:rec_rel3} page~\pageref{eq:rec_rel1}, be the least number of comparisons of keys that $ {\tt MergeSort} $ performs while sorting an $ n $-element array. \begin{thm} \label{thm:mainB_add} {\rm (Same as Theorem~\ref{thm:mainB}.)} For every positive integer $ n $ \footnote{Of course, one if free to assume that $ n $ is odd here.} and integer $ k $ with $ n \leq 2^k $, \begin{equation} \label{eq:mainB_add} \tilde{F} (\frac{n}{2^k}) = \frac{n \times k - 2 B(n)}{2^k} . \end{equation} \end{thm} \medskip The reminder of this Section constitutes a proof of Theorem~\ref{thm:mainB_add}. \medskip \textit{Note}. Function \textit{Zigzag}, visualized on Figure~\ref{fig:zigzag} page~\pageref{fig:zigzag}, has been defined by \eqref{eq:def_zigzag} page~\pageref{eq:def_zigzag}. \begin{lem} \label{lem:100} For every $ k \geq \lfloor \lg n \rfloor + 2 $, \begin{equation} \label{eq:100x} 2^k Zigzag (\frac{n}{2^k}) = n. \end{equation} \end{lem} \begin{proof} Let $ k \geq \lfloor \lg n \rfloor + 2 $, or $ 2^k \geq 2 \times 2^{\lfloor \lg n \rfloor+1} > 2n$, that is \begin{equation} \label{eq:110x} 2^k > 2n. \end{equation} We have: \[0 \leq \lfloor \frac{n}{2^k} \rfloor \leq \] [by \eqref{eq:110x}] \[\leq \lfloor \frac{n}{2n} \rfloor = \lfloor \frac{1}{2} \rfloor =0 \] or \begin{equation} \label{eq:120} \lfloor \frac{n}{2^k} \rfloor = 0. \end{equation} Also, \[1 \leq \lceil \frac{n}{2^k} \rceil \leq \] [by \eqref{eq:110x}] \[\leq \lceil \frac{n}{2n} \rceil = \lceil \frac{1}{2} \rceil =1 \] or \begin{equation} \label{eq:130} \lceil \frac{n}{2^k} \rceil = 1. \end{equation} Now, \[2^k Zigzag (\frac{n}{2^k}) =\] [by \eqref{eq:def_zigzag} page~\pageref{eq:def_zigzag}] \[ = 2^k \min \{ \frac{n}{2^k} - \lfloor \frac{n}{2^k} \rfloor,\lceil \frac{n}{2^k} \rceil - \frac{n}{2^k} \} = \min \{ n - 2^k \lfloor \frac{n}{2^k} \rfloor,2^k \lceil \frac{n}{2^k} \rceil - n \} = \] [by \eqref{eq:120} and \eqref{eq:130}] \[ = \min \{ n ,2^k - n \} = \] [since by \eqref{eq:110x}, $ 2^k - n \geq n $] \[=n.\] Hence, \eqref{eq:100x} holds. \end{proof} \begin{lem} \label{lem:200} For every $ k \geq \lfloor \lg n \rfloor + 1 $, \begin{equation} \label{eq:200} \sum_{i=\lfloor \lg n \rfloor + 2}^{k} 2^i Zigzag (\frac{n}{2^i}) = n \times k - n(\lfloor \lg n \rfloor + 1). \end{equation} \end{lem} \begin{proof} By induction on $ k $. \medskip \textit{Basis step}: $ k = \lfloor \lg n \rfloor + 1 $. \medskip \[ L = \sum_{i=\lfloor \lg n \rfloor + 2}^{\lfloor \lg n \rfloor + 1} 2^i Zigzag (\frac{n}{2^i}) = 0. \] \[R = n (\lfloor \lg n \rfloor + 1) - n (\lfloor \lg n \rfloor + 1) = 0. \] Hence, $ L=R $. This completes the \textit{Basis step}. \bigskip \textit{Inductive step}: $ k \geq \lfloor \lg n \rfloor + 2 $. \medskip \textit{Inductive hypothesis}: \eqref{eq:200}. \medskip \[ \sum_{i=\lfloor \lg n \rfloor + 2}^{k+1} 2^i Zigzag (\frac{n}{2^i}) = \sum_{i=\lfloor \lg n \rfloor + 2}^{k} 2^i Zigzag (\frac{n}{2^i}) + 2^k Zigzag (\frac{n}{2^k}) = \] [by the \textit{Inductive hypothesis} and by the equality~\eqref{eq:100x} of Lemma~\ref{lem:100}] \[= n \times k - n(\lfloor \lg n \rfloor + 1) + n = n \times (k+1) - n(\lfloor \lg n \rfloor + 1) .\] Thus \[\sum_{i=\lfloor \lg n \rfloor + 2}^{k+1} 2^i Zigzag (\frac{n}{2^i}) = n \times (k+1) - n(\lfloor \lg n \rfloor + 1) .\] This completes the \textit{Inductive step}. \end{proof} \begin{lem} \label{lem:300} For every $ k \geq \lfloor \lg n \rfloor + 1 $, \begin{equation} \label{eq:300} 2^k \tilde{F} (\frac{n}{2^k}) = \sum_{i=1}^{k} 2^i Zigzag (\frac{n}{2^i}). \end{equation} \end{lem} \begin{proof} By the definition \eqref{eq:def_lim2} page~\pageref{eq:def_lim2} of function $\tilde{F} $ , we get: \[ 2^k \tilde{F} (\frac{n}{2^k}) = 2^k \sum _{i=0} ^{\infty} \frac{1}{2^i} Zigzag (2^i \frac{n}{2^k}) = \] [since for every integer $ x $, $ Zigzag (x) = 0 $, so that for $ i \geq k $, $ Zigzag (2^i \frac{n}{2^k}) = 0 $] \[ = 2^k \sum _{i=0} ^{k-1} \frac{1}{2^i} Zigzag (2^i \frac{n}{2^k}) = \sum _{i=0} ^{k-1} 2^{k-i} Zigzag ( \frac{n}{2^{k-i}}) = \] [putting $ j = k - i $] \[ = \sum _{j=1} ^{k} 2^{j} Zigzag ( \frac{n}{2^{j}}),\] which completes the proof of \eqref{eq:300}. \end{proof} At this point, we are ready to conclude the proof of Theorem~\ref{thm:mainB}. \medskip By virtue of \eqref{eq:formula_Zigzag_B} page~\pageref{eq:formula_Zigzag_B}, we have: \[ 2 B(n) = n(\lfloor \lg n \rfloor + 1) - \sum _{k=0} ^{\lfloor \lg n \rfloor} 2^{k+1} Zigzag (\frac{n}{2^{k+1}}) = \] \[ = n(\lfloor \lg n \rfloor + 1) - \sum _{i=1} ^{\lfloor \lg n \rfloor+1} 2^{i} Zigzag (\frac{n}{2^{i}}) = \] \[ = n(\lfloor \lg n \rfloor + 1) - \sum _{i=1} ^{k} 2^{i} Zigzag (\frac{n}{2^{i}}) + \sum _{i=\lfloor \lg n \rfloor+2} ^{k} 2^{i} Zigzag (\frac{n}{2^{i}}) = \] [by Lemmas \ref{lem:200} and \ref{lem:300}] \[ = n(\lfloor \lg n \rfloor + 1) - 2^k \tilde{F} (\frac{n}{2^k}) + n \times k - n(\lfloor \lg n \rfloor + 1) = n \times k - 2^k \tilde{F} (\frac{n}{2^k}) ,\] that is, \[ 2 B(n) = n \times k - 2^k \tilde{F} (\frac{n}{2^k}) ,\] from which \eqref{eq:100x} follows. \medskip This completes the proof of Theorem~\ref{thm:mainB}. \bigskip \bigskip \textit{Note}. A glance at the proof of Lemma~\ref{lem:200} suffices to notice that it fails if $ n > 2^k $, and so does Theorem~\ref{thm:mainB}. In particular, for $ k = \lfloor \lg n \rfloor $, Lemma~\ref{lem:cff_F_tilde} page~\pageref{lem:cff_F_tilde} yields \begin{equation} \tilde{F} (\frac{n}{2^k}) = \frac{n \times k - 2 B(n)}{2^k} + \frac{2n}{2^k} -2 > \frac{n \times k - 2 B(n)}{2^k} \end{equation} since for $ n > 2^{\lfloor \lg n \rfloor} $, $ \frac{2n}{2^{\lfloor \lg n \rfloor}} -2 >0. $ \bibliographystyle{alpha}
4a8e621317426190a3f357f9d4025576dda881d1
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} We start with a brief recall of $t$-designs. Let ${\mathcal{P}}$ be a set of $v \ge 1$ elements, and let ${\mathcal{B}}$ be a set of $k$-subsets of ${\mathcal{P}}$, where $k$ is a positive integer with $1 \leq k \leq v$. Let $t$ be a positive integer with $t \leq k$. The pair ${\mathbb{D}} = ({\mathcal{P}}, {\mathcal{B}})$ is called a $t$-$(v, k, \lambda)$ {\em design\index{design}}, or simply {\em $t$-design\index{$t$-design}}, if every $t$-subset of ${\mathcal{P}}$ is contained in exactly $\lambda$ elements of ${\mathcal{B}}$. The elements of ${\mathcal{P}}$ are called points, and those of ${\mathcal{B}}$ are referred to as blocks. We usually use $b$ to denote the number of blocks in ${\mathcal{B}}$. A $t$-design is called {\em simple\index{simple}} if ${\mathcal{B}}$ does not contain repeated blocks. In this paper, we consider only simple $t$-designs. A $t$-design is called {\em symmetric\index{symmetric design}} if $v = b$. It is clear that $t$-designs with $k = t$ or $k = v$ always exist. Such $t$-designs are {\em trivial}. In this paper, we consider only $t$-designs with $v > k > t$. A $t$-$(v,k,\lambda)$ design is referred to as a {\em Steiner system\index{Steiner system}} if $t \geq 2$ and $\lambda=1$, and is denoted by $S(t,k, v)$. A necessary condition for the existence of a $t$-$(v, k, \lambda)$ design is that \begin{eqnarray}\label{eqn-tdesignnecessty} \binom{k-i}{t-i} \mbox{ divides } \lambda \binom{v-i}{t-i} \end{eqnarray} for all integer $i$ with $0 \leq i \leq t$. The interplay between codes and $t$-designs goes in two directions. In one direction, the incidence matrix of any $t$-design generates a linear code over any finite field ${\mathrm{GF}}(q)$. A lot of progress in this direction has been made and documented in the literature (see, for examples, \cite{AK92}, \cite{DingBook}, \cite{Tonchev,Tonchevhb}). In the other direction, the codewords of a fixed Hamming weight in a linear or nonlinear code may hold a $t$-design. Some linear and nonlinear codes were employed to construct $t$-designs \cite{AK92,HKM04,HMT05,KM00,MT04,Tonchev,Tonchevhb}. Binary and ternary Golay codes of certain parameters give $4$-designs and $5$-designs. However, the largest $t$ for which an infinite family of $t$-designs is derived directly from codes is $t=3$. According to the references \cite{AK92}, \cite{Tonchev,Tonchevhb} and \cite{KLhb}, not much progress on the construction of $t$-designs from codes has been made so far, while many other constructions of $t$-designs are documented in the literature (\cite{BJL,CMhb,KLhb,RR10}). The objective of this paper is to construct infinite families of $2$-designs and $3$-designs from a type of binary linear codes with five weights. The obtained $t$-designs depend only on the weight distribution of the underlying binary codes. The total number of $2$-designs and $3$-designs presented in this paper are exponential in $m$, where $m \geq 5$ is an odd integer. In addition, the block size of the designs can vary in a huge range. \section{The classical construction of $t$-designs from codes} Let ${\mathcal{C}}$ be a $[v, \kappa, d]$ linear code over ${\mathrm{GF}}(q)$. Let $A_i:=A_i({\mathcal{C}})$, which denotes the number of codewords with Hamming weight $i$ in ${\mathcal{C}}$, where $0 \leq i \leq v$. The sequence $(A_0, A_1, \cdots, A_{v})$ is called the \textit{weight distribution} of ${\mathcal{C}}$, and $\sum_{i=0}^v A_iz^i$ is referred to as the \textit{weight enumerator} of ${\mathcal{C}}$. For each $k$ with $A_k \neq 0$, let ${\mathcal{B}}_k$ denote the set of the supports of all codewords with Hamming weight $k$ in ${\mathcal{C}}$, where the coordinates of a codeword are indexed by $(0,1,2, \cdots, v-1)$. Let ${\mathcal{P}}=\{0, 1, 2, \cdots, v-1\}$. The pair $({\mathcal{P}}, {\mathcal{B}}_k)$ may be a $t$-$(v, k, \lambda)$ design for some positive integer $\lambda$. The following theorems, developed by Assumus and Mattson, show that the pair $({\mathcal{P}}, {\mathcal{B}}_k)$ defined by a linear code is a $t$-design under certain conditions. \begin{theorem}\label{thm-AM1}[Assmus-Mattson Theorem \cite{AM74}, \cite[p. 303]{HP03}] Let ${\mathcal{C}}$ be a binary $[v, \kappa, d]$ code. Suppose ${\mathcal{C}}^\perp$ has minimum weight $d^\perp$. Suppose that $A_i=A_i({\mathcal{C}})$ and $A_i^\perp=A_i({\mathcal{C}}^\perp)$, for $0 \leq i \leq v$, are the weight distributions of ${\mathcal{C}}$ and ${\mathcal{C}}^\perp$, respectively. Fix a positive integer $t$ with $t < d$, and let $s$ be the number of $i$ with $A_i^\perp \ne 0$ for $0 < i \leq v-t$. Suppose that $s \leq d -t$. Then \begin{itemize} \item the codewords of weight $i$ in ${\mathcal{C}}$ hold a $t$-design provided that $A_i \ne 0$ and $d \leq i \leq v$, and \item the codewords of weight $i$ in ${\mathcal{C}}^\perp$ hold a $t$-design provided that $A_i^\perp \ne 0$ and $d^\perp \leq i \leq v$. \end{itemize} \end{theorem} To construct $t$-designs via Theorem \ref{thm-AM1}, we will need the following lemma in subsequent sections, which is a variant of the MacWilliam Identity \cite[p. 41]{vanLint}. \begin{theorem} \label{thm-MI} Let ${\mathcal{C}}$ be a $[v, \kappa, d]$ code over ${\mathrm{GF}}(q)$ with weight enumerator $A(z)=\sum_{i=0}^v A_iz^i$ and let $A^\perp(z)$ be the weight enumerator of ${\mathcal{C}}^\perp$. Then $$A^\perp(z)=q^{-\kappa}\Big(1+(q-1)z\Big)^vA\Big(\frac {1-z} {1+(q-1)z}\Big).$$ \end{theorem} Later in this paper, we will need also the following theorem. \begin{theorem}\label{thm-allcodes} Let ${\mathcal{C}}$ be an $[n, k, d]$ binary linear code, and let ${\mathcal{C}}^\perp$ denote the dual of ${\mathcal{C}}$. Denote by $\overline{{\mathcal{C}}^\perp}$ the extended code of ${\mathcal{C}}^\perp$, and let $\overline{{\mathcal{C}}^\perp}^\perp$ denote the dual of $\overline{{\mathcal{C}}^\perp}$. Then we have the following. \begin{enumerate} \item ${\mathcal{C}}^\perp$ has parameters $[n, n-k, d^\perp]$, where $d^\perp$ denotes the minimum distance of ${\mathcal{C}}^\perp$. \item $\overline{{\mathcal{C}}^\perp}$ has parameters $[n+1, n-k, \overline{d^\perp}]$, where $\overline{d^\perp}$ denotes the minimum distance of $\overline{{\mathcal{C}}^\perp}$, and is given by \begin{eqnarray*} \overline{d^\perp} = \left\{ \begin{array}{ll} d^\perp & \mbox{ if $d^\perp$ is even,} \\ d^\perp + 1 & \mbox{ if $d^\perp$ is odd.} \end{array} \right. \end{eqnarray*} \item $\overline{{\mathcal{C}}^\perp}^\perp$ has parameters $[n+1, k+1, \overline{d^\perp}^\perp]$, where $\overline{d^\perp}^\perp$ denotes the minimum distance of $\overline{{\mathcal{C}}^\perp}^\perp$. Furthermore, $\overline{{\mathcal{C}}^\perp}^\perp$ has only even-weight codewords, and all the nonzero weights in $\overline{{\mathcal{C}}^\perp}^\perp$ are the following: \begin{eqnarray*} w_1, \, w_2,\, \cdots,\, w_t;\, n+1-w_1, \, n+1-w_2, \, \cdots, \, n+1-w_t; \, n+1, \end{eqnarray*} where $w_1,\, w_2,\, \cdots,\, w_t$ denote all the nonzero weights of ${\mathcal{C}}$. \end{enumerate} \end{theorem} \begin{proof} The conclusions of the first two parts are straightforward. We prove only the conclusions of the third part below. Since $\overline{{\mathcal{C}}^\perp}$ has length $n+1$ and dimension $n-k$, the dimension of $\overline{{\mathcal{C}}^\perp}^\perp$ is $k+1$. By assumption, all codes under consideration are binary. By definition, $\overline{{\mathcal{C}}^\perp}$ has only even-weight codewords. Recall that $\overline{{\mathcal{C}}^\perp}$ is the extended code of ${\mathcal{C}}^\perp$. It is known that the generator matrix of $\overline{{\mathcal{C}}^\perp}^\perp$ is given by (\cite[p. 15]{HP03}) \begin{eqnarray*} \left[ \begin{array}{cc} {\mathbf{\bar{1}}} & 1 \\ G & {\mathbf{\bar{0}}} \end{array} \right]. \end{eqnarray*} where ${\mathbf{\bar{1}}}=(1 1 1 \cdots 1)$ is the all-one vector of length $n$, ${\mathbf{\bar{0}}}=(000 \cdots 0)^T$, which is a column vector of length $n$, and $G$ is the generator matrix of ${\mathcal{C}}$. Notice again that $\overline{{\mathcal{C}}^\perp}^\perp$ is binary, the desired conclusions on the weights in $\overline{{\mathcal{C}}^\perp}^\perp$ follow from the relation between the two generator matrices of the two codes $\overline{{\mathcal{C}}^\perp}^\perp$ and ${\mathcal{C}}$. \end{proof} \section{A type of binary linear codes with five-weights and related codes}\label{sec-maincodes} In this section, we first introduce a type of binary linear codes ${\mathcal{C}}_m$ of length $n=2^m-1$, which has the weight distribution of Table \ref{tab-zhou3}, and then analyze their dual codes ${\mathcal{C}}_m^\perp$, the extended codes $\overline{{\mathcal{C}}_m^\perp}$, and the duals $\overline{{\mathcal{C}}_m^\perp}^\perp$. Such codes will be employed to construct $t$-designs in Sections \ref{sec-2designs} and \ref{sec-3designs}. Examples of such codes will be given in Section \ref{sec-examplecodes}. \begin{table}[ht] \caption{The weight distribution of ${\mathcal{C}}_{m}$ for odd $m$.}\label{tab-zhou3} \centering \begin{tabular}{ll} \hline Weight $w$ & No. of codewords $A_w$ \\ \hline $0$ & $1$ \\ $2^{m-1}-2^{(m+1)/2}$ & $(2^m-1)\cdot 2^{(m-5)/2}\cdot (2^{(m-3)/2}+1)\cdot (2^{m-1}-1)/3$ \\ $2^{m-1}-2^{(m-1)/2}$ & $(2^m-1)\cdot 2^{(m-3)/2}\cdot (2^{(m-1)/2}+1)\cdot (5\cdot 2^{m-1}+4)/3$ \\ $2^{m-1}$ & ${(2^m-1)}\cdot (9\cdot 2^{2m-4}+3\cdot 2^{m-3}+1)$ \\ $2^{m-1}+2^{(m-1)/2}$ & $(2^m-1)\cdot 2^{(m-3)/2}\cdot (2^{(m-1)/2}-1)\cdot (5\cdot 2^{m-1}+4)/3$ \\ $2^{m-1}+2^{(m+1)/2}$ & $(2^m-1)\cdot 2^{(m-5)/2}\cdot (2^{(m-3)/2}-1)\cdot (2^{m-1}-1)/3$ \\ \hline \end{tabular} \end{table} \begin{theorem}\label{thm-BCHcodeDual} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the dual code ${\mathcal{C}}_{m}^\perp$ has parameters $[2^m-1, 2^m-1-3m, 7]$, and its weight distribution is given by \begin{eqnarray*} 2^{3m}A^\perp_k &=& \binom{2^m-1}{k} +a U_a(k) + b U_b(k) + c U_c(k) + d U_d(k) + e U_e(k), \end{eqnarray*} where $0 \leq k \leq 2^m-1$, \begin{eqnarray*} a &=& (2^m-1) 2^{(m-5)/2} (2^{(m-3)/2}+1) (2^{m-1}-1)/3, \\ b &=& (2^m-1) 2^{(m-3)/2} (2^{(m-1)/2}+1) (5\times 2^{m-1}+4)/3, \\ c &=& {(2^m-1)} (9\times 2^{2m-4}+3\times 2^{m-3}+1), \\ d &=& (2^m-1) 2^{(m-3)/2} (2^{(m-1)/2}-1) (5\times 2^{m-1}+4)/3, \\ e &=& (2^m-1) 2^{(m-5)/2} (2^{(m-3)/2}-1) (2^{m-1}-1)/3, \end{eqnarray*} and \begin{eqnarray*} U_a(k) = \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m+1)/2} \\ 0\le j \le 2^{m-1}+2^{(m+1)/2}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m+1)/2}} {i} \binom{2^{m-1}+2^{(m+1)/2}-1}{j}, \end{eqnarray*} \begin{eqnarray*} U_b(k) = \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m-1)/2} \\ 0\le j \le 2^{m-1}+2^{(m-1)/2}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m-1)/2}} {i} \binom{2^{m-1}+2^{(m-1)/2}-1}{j}, \end{eqnarray*} \begin{eqnarray*} U_c(k) = \sum_{\substack{0 \le i \le 2^{m-1} \\ 0\le j \le 2^{m-1}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}} {i}\binom{2^{m-1}-1} {j}, \end{eqnarray*} \begin{eqnarray*} U_d(k) = \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m-1)/2} \\ 0\le j \le 2^{m-1}-2^{(m-1)/2}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m-1)/2}}{i}\binom{2^{m-1}-2^{(m-1)/2}-1}{j}, \end{eqnarray*} \begin{eqnarray*} U_e(k) &=& \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m+1)/2} \\ 0\le j \le 2^{m-1}-2^{(m+1)/2}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m+1)/2}}{i}\binom{2^{m-1}-2^{(m+1)/2}-1}{j}. \end{eqnarray*} \end{theorem} \begin{proof} By assumption, the weight enumerator of ${\mathcal{C}}_{m}$ is given by $$ A(z)=1+az^{2^{m-1}-2^{(m+1)/2}} +bz^{2^{m-1}-2^{(m-1)/2}} +cz^{2^{m-1}}+dz^{2^{m-1}+2^{(m-1)/2}} +ez^{2^{m-1}+2^{(m+1)/2}}. $$ It then follows from Theorem \ref{thm-MI} that the weight enumerator of ${\mathcal{C}}_{m}^\perp$ is given by \begin{eqnarray*} 2^{3m}A^\perp(z) &=& (1+z)^{2^m-1}\left[ 1 + a\left(\frac {1-z} {1+z}\right)^{2^{m-1}-2^{\frac{m+1}{2}}} + b\left(\frac {1-z} {1+z}\right)^{2^{m-1}-2^{ \frac{m-1}{2} }} \right] + \\ & & (1+z)^{2^m-1}\left[ c\left(\frac {1-z} {1+z}\right)^{2^{m-1}}+ d\left(\frac {1-z} {1+z}\right)^{2^{m-1}+2^{\frac{m-1}{2} }} + e\left(\frac {1-z} {1+z}\right)^{2^{m-1}+2^{ \frac{m+1}{2} }} \right]. \end{eqnarray*} Hence, we have \begin{eqnarray*} 2^{3m}A^\perp(z) &=& (1+z)^{2^m-1} + \\ & & a(1-z)^{2^{m-1}-2^{(m+1)/2}}(1+z)^{2^{m-1}+2^{(m+1)/2}-1} + \\ & & b(1-z)^{2^{m-1}-2^{(m-1)/2}}(1+z)^{2^{m-1}+2^{(m-1)/2}-1} + \\ & & c(1-z)^{2^{m-1}}(1+z)^{2^{m-1}-1} + \\ & & d(1-z)^{2^{m-1}+2^{(m-1)/2}}(1+z)^{2^{m-1}-2^{(m-1)/2}-1} + \\ & & e(1-z)^{2^{m-1}+2^{(m+1)/2}}(1+z)^{2^{m-1}-2^{(m+1)/2}-1}. \end{eqnarray*} Obviously, we have \begin{eqnarray*} (1+z)^{2^m-1} &=& \sum_{k=0}^{2^m-1} \binom{2^m-1}{k}z^k. \end{eqnarray*} It is easily seen that \begin{eqnarray*} (1-z)^{2^{m-1}-2^{(m+1)/2}}(1+z)^{2^{m-1}+2^{(m+1)/2}-1} = \sum_{k=0}^{2^m-1} U_a(k) z^k \end{eqnarray*} and \begin{eqnarray*} (1-z)^{2^{m-1}-2^{(m-1)/2}}(1+z)^{2^{m-1}+2^{(m-1)/2}-1} = \sum_{k=0}^{2^m-1} U_b(k) z^k. \end{eqnarray*} Similarly, \begin{eqnarray*} (1-z)^{2^{m-1}+2^{(m-1)/2}}(1+z)^{2^{m-1}-2^{(m-1)/2}-1} = \sum_{k=0}^{2^m-1} U_d(k) z^k \end{eqnarray*} and \begin{eqnarray*} (1-z)^{2^{m-1}+2^{(m+1)/2}}(1+z)^{2^{m-1}-2^{(m+1)/2}-1} = \sum_{k=0}^{2^m-1} U_e(k) z^k. \end{eqnarray*} Finally, we have \begin{eqnarray*} (1-z)^{2^{m-1}}(1+z)^{2^{m-1}-1}= \sum_{k=0}^{2^m-1} U_c(k) z^k. \end{eqnarray*} Combining these formulas above yields the weight distribution formula for $A_k^\perp$. The weight distribution in Table \ref{tab-zhou3} tells us that the dimension of ${\mathcal{C}}_{m}$ is $3m$. Therefore, the dimension of ${\mathcal{C}}_{m}^\perp$ is equal to $2^m-1-3m$. Finally, we prove that the minimum distance of ${\mathcal{C}}_{m}^\perp$ equals $7$. We now prove that $A_k^\perp=0$ for all $k$ with $1 \leq k \leq 6$. Let $x=2^{(m-1)/2}$. With the weight distribution formula obtained before, we have \begin{eqnarray*} \binom{2^m-1}{1} &=& 2x^2 - 1,\\ a U_a(1) &=& 1/3 x^7 + 7/12 x^6 - 2/3 x^5 - 7/8 x^4 + 5/12 x^3 + 7/24 x^2 - 1/12 x, \\ b U_b(1) &=& 10/3x^7 + 5/3x^6 - 2/3x^5 + 1/2x^4 - 11/6x^3 - 2/3x^2 + 2/3x, \\ c U_c(1) &=& -9/2x^6 + 3/4x^4 - 5/4x^2 + 1, \\ d U_d(1) &=& -10/3x^7 + 5/3x^6 + 2/3x^5 + 1/2x^4 + 11/6x^3 - 2/3x^2 - 2/3x, \\ e U_e(1) &=& -1/3x^7 + 7/12x^6 + 2/3x^5 - 7/8x^4 - 5/12x^3 + 7/24x^2 + 1/12x. \end{eqnarray*} Consequently, \begin{eqnarray*} 2^{3m} A_1^\perp = \binom{2^m-1}{1} +a U_a(1) + b U_b(1) + c U_c(1) + d U_d(1) + e U_e(1)=0. \end{eqnarray*} Plugging $k=2$ into the weight distribution formula above, we get that \begin{eqnarray*} \binom{2^m-1}{2} &=& 2x^4 - 3x^2 + 1,\\ a U_a(2) &=& 7/12x^8 + 5/6x^7 - 35/24x^6 - 13/12x^5 + 7/6x^4 + 1/6x^3 - 7/24x^2 + 1/12x,\\ b U_b(2) &=& 5/3x^8 - 5/3x^7 - 7/6x^6 + 7/6x^5 - 7/6x^4 + 7/6x^3 + 2/3x^2 - 2/3x, \\ c U_c(2) &=& -9/2x^8 + 21/4x^6 - 2x^4 + 9/4x^2 - 1, \\ d U_d(2) &=& 5/3x^8 + 5/3x^7 - 7/6x^6 - 7/6x^5 - 7/6x^4 - 7/6x^3 + 2/3x^2 + 2/3x, \\ e U_e(2) &=& 7/12x^8 - 5/6x^7 - 35/24x^6 + 13/12x^5 + 7/6x^4 - 1/6x^3 - 7/24x^2 - 1/12x. \end{eqnarray*} Consequently, \begin{eqnarray*} 2^{3m} A_2^\perp = \binom{2^m-1}{2} +a U_a(2) + b U_b(2) + c U_c(2) + d U_d(2) + e U_e(2)=0. \end{eqnarray*} After similar computations with the weight distribution formula, one can prove that $A_k^\perp=0$ for all $k$ with $3 \leq k \leq 6$. Plugging $k=7$ into the weight distribution formula above, we arrive at \begin{eqnarray*} A_7^\perp = \frac{(x^2-1) (2x^2 - 1) (x^4 - 5x^2 + 34)}{630} . \end{eqnarray*} Notice that $x^4 - 5x^2 + 34=(x^2-5/2)^2 +34-25/4 >0$. We have $A_7^\perp >0$ for all odd $m \geq 5$. This proves the desired conclusion on the minimum distance of ${\mathcal{C}}_{m}^\perp$. \end{proof} \begin{theorem}\label{thm-lastcode} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. The code $\overline{{\mathcal{C}}_{m}^\perp}^\perp$ has parameters $$ \left[2^m, \, 3m+1, \, 2^{m-1}-2^{(m+1)/2}\right], $$ and its weight enumerator is given by \begin{eqnarray}\label{eqn-doubledual} \overline{A^\perp}^\perp (z) = 1+uz^{2^{m-1}-2^{\frac{m+1}{2} }} + vz^{2^{m-1}-2^{ \frac{m-1}{2}}} + wz^{2^{m-1}} + vz^{2^{m-1}+2^{ \frac{m-1}{2}}} + uz^{2^{m-1}+2^{\frac{m+1}{2} }} +z^{2^m}, \end{eqnarray} where \begin{eqnarray*} u &=& \frac{2^{3m-4} - 3 \times 2^{2m-4} + 2^{m-3}}{3}, \\ v &=& \frac{5\times 2^{3m-2} + 3 \times 2^{2m-2} - 2^{m+1}}{3} , \\ w &=& {2(2^m-1)} (9\times 2^{2m-4}+3\times 2^{m-3}+1). \end{eqnarray*} \end{theorem} \begin{proof} It follows from Theorem \ref{thm-allcodes} that the code has all the weights given in (\ref{eqn-doubledual}). It remains to determine the frequencies of these weights. The weight distribution of the code ${\mathcal{C}}_{m}$ given in Table \ref{tab-zhou3} and the generator matrix of the code $\overline{{\mathcal{C}}_{m}^\perp}^\perp$ documented in the proof of Theorem \ref{thm-allcodes} show that $$ \overline{A^\perp}^\perp_{2^{m-1}}=2c=w, $$ where $c$ was defined in Theorem \ref{thm-BCHcodeDual}. We now determine $u$ and $v$. Recall that ${\mathcal{C}}_{m}^\perp$ has minimum distance $7$. It then follows from Theorem \ref{thm-allcodes} that $\overline{{\mathcal{C}}_{m}^\perp}$ has minimum distance $8$. The first and third Pless power moments say that \begin{eqnarray*} \left\{ \begin{array}{lll} \sum_{i=0}^{2^m} \overline{A^\perp}^\perp_{i} &=& 2^{3m+1}, \\ \sum_{i=0}^{2^m} i^2 \overline{A^\perp}^\perp_{i} &=& 2^{3m-1} 2^m(2^m+1). \end{array} \right. \end{eqnarray*} These two equations become \begin{eqnarray*} \left\{ \begin{array}{l} 1+u+v+c = 2^{3m}, \\ (2^{2m-2} +2^{m+1}) u + (2^{2m-2} +2^{m-1}) v + 2^{2m-2}c + 2^{2m-1} = 2^{4m-2}(2^m+1). \end{array} \right. \end{eqnarray*} Solving this system of equations proves the desired conclusion on the weight enumerator of this code. \end{proof} Finally, we settle the weight distribution of the code $\overline{{\mathcal{C}}_{m}^\perp}$. \begin{theorem}\label{thm-3rdcode} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. The code $\overline{{\mathcal{C}}_{m}^\perp}$ has parameters $[2^m, 2^m-1-3m, 8]$, and its weight distribution is given by \begin{eqnarray} 2^{3m+1}\overline{A^\perp}_k &=& \left(1+(-1)^k \right) \binom{2^m}{k} + w E_0(k) + u E_1(k) + v E_2(k) + v E_3(k) +u E_4(k), \end{eqnarray} where $w, u, v$ are defined in Theorem \ref{thm-lastcode}, and \begin{eqnarray*} E_0(k) = \frac{1+(-1)^k}{2} (-1)^{\lfloor k/2 \rfloor} \binom{2^{m-1}}{\lfloor k/2 \rfloor}, \end{eqnarray*} \begin{eqnarray*} E_1(k) = \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m+1)/2} \\ 0\le j \le 2^{m-1}+2^{(m+1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m+1)/2}} {i} \binom{2^{m-1}+2^{(m+1)/2}}{j}, \end{eqnarray*} \begin{eqnarray*} E_2(k) = \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m-1)/2} \\ 0\le j \le 2^{m-1}+2^{(m-1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m-1)/2}} {i} \binom{2^{m-1}+2^{(m-1)/2}}{j}, \end{eqnarray*} \begin{eqnarray*} E_3(k) = \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m-1)/2} \\ 0\le j \le 2^{m-1}-2^{(m-1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m-1)/2}} {i} \binom{2^{m-1}-2^{(m-1)/2}}{j}, \end{eqnarray*} \begin{eqnarray*} E_4(k) = \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m+1)/2} \\ 0\le j \le 2^{m-1}-2^{(m+1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m+1)/2}} {i} \binom{2^{m-1}-2^{(m+1)/2}}{j}, \end{eqnarray*} and $0 \leq k \leq 2^m$. \end{theorem} \begin{proof} By definition, $$ \dim\left( \overline{{\mathcal{C}}_{m}^\perp} \right) = \dim\left({\mathcal{C}}_{m}^\perp \right)=2^m-1-3m. $$ It has been showed in the proof of Theorem \ref{thm-BCHcodeDual} that the minimum distance of $\overline{{\mathcal{C}}_{m}^\perp}$ is equal to $8$. We now prove the conclusion on the weight distribution of this code. By Theorems \ref{thm-MI} and \ref{thm-lastcode}, the weight enumerator of $\overline{{\mathcal{C}}_{m}^\perp}$ is given by \begin{eqnarray}\label{eqn-july0} 2^{3m+1}\overline{A^\perp}(z) &=& (1+z)^{2^m}\left[ 1 + \left(\frac{1-z}{1+z}\right)^{2^m} + w\left(\frac {1-z} {1+z}\right)^{2^{m-1}} \right] + \nonumber \\ & & (1+z)^{2^m}\left[ u\left(\frac {1-z} {1+z}\right)^{2^{m-1}-2^{\frac{m+1}{2} }}+ v \left(\frac {1-z} {1+z}\right)^{2^{m-1}-2^{\frac{m-1}{2} }} \right] + \nonumber \\ && (1+z)^{2^m}\left[ v \left(\frac {1-z} {1+z}\right)^{2^{m-1}+2^{\frac{m-1}{2} }} + u\left(\frac {1-z} {1+z}\right)^{2^{m-1}+2^{\frac{m+1}{2}}} \right]. \end{eqnarray} Consequently, we have \begin{eqnarray}\label{eqn-j18-1} 2^{3m+1}\overline{A^\perp}(z) &=& (1+z)^{2^m} + (1-z)^{2^m} + w (1-z^2)^{2^{m-1}} + \nonumber \\ & & u(1-z)^{2^{m-1}-2^{(m+1)/2}}(1+z)^{2^{m-1}+2^{(m+1)/2}} + \nonumber \\ & & v(1-z)^{2^{m-1}-2^{(m-1)/2}}(1+z)^{2^{m-1}+2^{(m-1)/2}} + \nonumber \\ & & v(1-z)^{2^{m-1}+2^{(m-1)/2}}(1+z)^{2^{m-1}-2^{(m-1)/2}} + \nonumber \\ & & u(1-z)^{2^{m-1}+2^{(m+1)/2}}(1+z)^{2^{m-1}-2^{(m+1)/2}}. \end{eqnarray} We now treat the terms in (\ref{eqn-j18-1}) one by one. We first have \begin{eqnarray}\label{eqn-j18-2} (1+z)^{2^m} + (1-z)^{2^m} = \sum_{k=0}^{2^m} \left(1+(-1)^k \right) \binom{2^m}{k}. \end{eqnarray} One can easily see that \begin{eqnarray}\label{eqn-j18-3} (1-z^2)^{2^{m-1}} = \sum_{i=0}^{2^{m-1}} (-1)^i \binom{2^{m-1}}{i} z^{2i} = \sum_{k=0}^{2^{m}} \frac{1+(-1)^k}{2} (-1)^{\lfloor k/2 \rfloor} \binom{2^{m-1}}{\lfloor k/2 \rfloor} z^{k}. \end{eqnarray} Notice that \begin{eqnarray*} (1-z)^{2^{m-1}-2^{(m+1)/2}}=\sum_{i=0}^{2^{m-1}-2^{(m+1)/2}} \binom{2^{m-1}-2^{(m+1)/2}}{i} (-1)^i z^i \end{eqnarray*} and \begin{eqnarray*} (1+z)^{2^{m-1}+2^{(m+1)/2}}=\sum_{i=0}^{2^{m-1}+2^{(m+1)/2}} \binom{2^{m-1}+2^{(m+1)/2}}{i} z^i \end{eqnarray*} We have then \begin{eqnarray}\label{eqn-j18-4} (1-z)^{2^{m-1}-2^{(m+1)/2}} (1+z)^{2^{m-1}+2^{(m+1)/2}} = \sum_{k=0}^{2^m} E_1(k) z^k. \end{eqnarray} Similarly, we have \begin{eqnarray}\label{eqn-j18-5} (1-z)^{2^{m-1}-2^{(m-1)/2}} (1+z)^{2^{m-1}+2^{(m-1)/2}} = \sum_{k=0}^{2^m} E_2(k) z^k, \end{eqnarray} \begin{eqnarray}\label{eqn-j18-6} (1-z)^{2^{m-1}+2^{(m-1)/2}} (1+z)^{2^{m-1}-2^{(m-1)/2}} = \sum_{k=0}^{2^m} E_3(k) z^k, \end{eqnarray} \begin{eqnarray}\label{eqn-j18-7} (1-z)^{2^{m-1}+2^{(m+1)/2}} (1+z)^{2^{m-1}-2^{(m+1)/2}} = \sum_{k=0}^{2^m} E_4(k) z^k. \end{eqnarray} Plugging (\ref{eqn-j18-2}), (\ref{eqn-j18-3}), (\ref{eqn-j18-4}), (\ref{eqn-j18-5}), (\ref{eqn-j18-6}), and (\ref{eqn-j18-7}) into (\ref{eqn-j18-1}) proves the desired conclusion. \end{proof} \section{Infinite families of $2$-designs from ${\mathcal{C}}_{m}^\perp$ and ${\mathcal{C}}_{m}$}\label{sec-2designs} \begin{theorem}\label{thm-2designs} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords of ${\mathcal{C}}_{m}$ with weight $k$, where $A_k \neq 0$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{k(k-1)A_k}{(2^m-1)(2^m-2)}, \end{eqnarray*} where $A_k$ is given in Table \ref{tab-zhou3}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}^\perp$ be the set of the supports of the codewords of ${\mathcal{C}}_{m}^\perp$ with weight $k$ and $A_k^\perp \neq 0$. Then $({\mathcal{P}}, {\mathcal{B}}^\perp)$ is a $2$-$(2^m-1, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{k(k-1)A_k^\perp}{(2^m-1)(2^m-2)}, \end{eqnarray*} where $A_k^\perp$ is given in Theorem \ref{thm-BCHcodeDual}. \end{theorem} \begin{proof} The weight distribution of ${\mathcal{C}}_{m}^\perp$ is given in Theorem \ref{thm-BCHcodeDual} and that of ${\mathcal{C}}_{m}$ is given in Table \ref{tab-zhou3}. By Theorem \ref{thm-BCHcodeDual}, the minimum distance $d^\perp$ of ${\mathcal{C}}_{m}^\perp$ is equal to $7$. Put $t=2$. The number of $i$ with $A_i \neq 0$ and $1 \leq i \leq 2^m-1 -t$ is $s=5$. Hence, $s=d^\perp-t$. The desired conclusions then follow from Theorem \ref{thm-AM1} and the fact that two binary vectors have the same support if and only if they are equal. \end{proof} \begin{example} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the BCH code ${\mathcal{C}}_{m}$ holds five $2$-designs with the following parameters: \begin{itemize} \item $(v,\, k, \, \lambda)=\left(2^m-1,\ 2^{m-1}-2^{\frac{m+1}{2}}, \ \frac{2^{\frac{m-5}{2}} \left(2^{\frac{m-3}{2}}+1\right) \left(2^{m-1} - 2^{\frac{m+1}{2}} \right)\left(2^{m-1} - 2^{\frac{m+1}{2}} -1\right)}{6} \right).$ \item $(v,\, k, \, \lambda)=\left(2^m-1,\ 2^{m-1}-2^{\frac{m-1}{2}}, \ \frac{2^{m-2} \left(2^{m-1} - 2^{\frac{m-1}{2} } -1 \right)\left( 5 \times 2^{m-1} + 4 \right)}{6} \right).$ \item $(v, \, k, \, \lambda)=\left(2^m-1, \ 2^{m-1}, \ 2^{m-2} (9 \times 2^{2m-4}+ 3 \times 2^{m-3} +1) \right).$ \item $(v,\, k, \, \lambda)=\left(2^m-1,\ 2^{m-1}+2^{\frac{m-1}{2}}, \ \frac{2^{m-2} \left(2^{m-1} + 2^{\frac{m-1}{2} } -1 \right)\left( 5 \times 2^{m-1} + 4 \right)}{6} \right).$ \item $(v,\, k, \, \lambda)=\left(2^m-1,\ 2^{m-1}+2^{\frac{m+1}{2}}, \ \frac{2^{\frac{m-5}{2}} \left(2^{\frac{m-3}{2}}-1\right) \left(2^{m-1} + 2^{\frac{m+1}{2}} \right)\left(2^{m-1} + 2^{\frac{m+1}{2}} -1\right)}{6} \right).$ \end{itemize} \end{example} \begin{example} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the supports of all codewords of weight $7$ in ${\mathcal{C}}_{m}^\perp$ give a $2$-$(2^m-1, 7, \lambda)$ design, where $$ \lambda=\frac{ 2^{2(m-1)} - 5 \times 2^{m-1} + 34 }{30}. $$ \end{example} \begin{proof} By Theorem \ref{thm-BCHcodeDual}, we have $$ A^\perp_7=\frac{(2^{m-1}-1) (2^m-1) (2^{2(m-1)} - 5 \times 2^{m-1} + 34)}{630}. $$ The desired conclusion on $\lambda$ follows from Theorem \ref{thm-2designs}. \end{proof} \begin{example} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the supports of all codewords of weight $8$ in ${\mathcal{C}}_{m}^\perp$ give a $2$-$(2^m-1, 8, \lambda)$ design, where $$ \lambda=\frac{ (2^{m-1}-4)(2^{2(m-1)} - 5 \times 2^{m-1} + 34) }{90}. $$ \end{example} \begin{proof} By Theorem \ref{thm-BCHcodeDual}, we have $$ A^\perp_8=\frac{(2^{m-1}-1) (2^{m-1}-4) (2^m-1) (2^{2(m-1)} - 5 \times 2^{m-1} + 34)}{2520}. $$ The desired conclusion on $\lambda$ follows from Theorem \ref{thm-2designs}. \end{proof} \begin{example} Let $m \geq 7$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the supports of all codewords of weight $9$ in ${\mathcal{C}}_{m}^\perp$ give a $2$-$(2^m-1, 9, \lambda)$ design, where $$ \lambda=\frac{ (2^{m-1}-4)(2^{m-1}-16)(2^{2(m-1)} - 2^{m-1} + 28) }{315}. $$ \end{example} \begin{proof} By Theorem \ref{thm-BCHcodeDual}, we have $$ A^\perp_9=\frac{(2^{m-1}-1) (2^{m-1}-4) (2^{m-1}-16) (2^m-1) (2^{2(m-1)} - 2^{m-1} + 28)}{11340}. $$ The desired conclusion on $\lambda$ follows from Theorem \ref{thm-2designs}. \end{proof} \section{Infinite families of $3$-designs from $\overline{{\mathcal{C}}_{m}^\perp}$ and $\overline{{\mathcal{C}}_{m}^\perp}^\perp$}\label{sec-3designs} \begin{theorem}\label{thm-newdesigns2} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-1\}$, and let $\overline{{\mathcal{B}}^\perp}^\perp$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}_{m}^\perp}^\perp$ with weight $k$, where $\overline{A^\perp}^\perp_k \neq 0$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}^\perp}^\perp)$ is a $3$-$(2^m, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{\overline{A^\perp}^\perp_k\binom{k}{3}}{\binom{2^m}{3}}, \end{eqnarray*} where $\overline{A^\perp}^\perp_k$ is given in Theorem \ref{thm-lastcode}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-1\}$, and let $\overline{{\mathcal{B}}^\perp}$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}_{m}^\perp}$ with weight $k$ and $\overline{A^\perp}_k \neq 0$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}^\perp})$ is a $3$-$(2^m, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{\overline{A^\perp}_k \binom{k}{3}}{\binom{2^m}{3}}, \end{eqnarray*} where $\overline{A^\perp}_k$ is given in Theorem \ref{thm-3rdcode}. \end{theorem} \begin{proof} The weight distributions of $\overline{{\mathcal{C}}_{m}^\perp}^\perp$ and $\overline{{\mathcal{C}}_{m}^\perp}$ are described in Theorems \ref{thm-lastcode} and \ref{thm-3rdcode}. Notice that the minimum distance $\overline{d^\perp}$ of $\overline{{\mathcal{C}}_{m}^\perp}^\perp$ is equal to $8$. Put $t=3$. The number of $i$ with $\overline{A^\perp}_i \neq 0$ and $1 \leq i \leq 2^m -t$ is $s=5$. Hence, $s=\overline{d^\perp}-t$. Clearly, two binary vectors have the same support if and only if they are equal. The desired conclusions then follow from Theorem \ref{thm-AM1}. \end{proof} \begin{example} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then $\overline{{\mathcal{C}}_{m}^\perp}^\perp$ holds five $3$-designs with the following parameters: \begin{itemize} \item $(v,\, k, \, \lambda)=\left(2^m,\ 2^{m-1}-2^{\frac{m+1}{2}}, \ \frac{ \left(2^{m-1} - 2^{\frac{m+1}{2}} \right) \left(2^{m-1} - 2^{\frac{m+1}{2}} -1\right) \left(2^{m-1} - 2^{\frac{m+1}{2}} -2\right)}{48} \right).$ \item $(v,\, k, \, \lambda)=\left(2^m,\ 2^{m-1}-2^{\frac{m-1}{2}}, \ \frac{ 2^{\frac{m-1}{2} } \left(2^{m-1} - 2^{\frac{m-1}{2} } -1 \right) \left( 2^{\frac{m-1}{2} } -2 \right)\left( 5 \times 2^{m-3} + 1 \right)}{3} \right).$ \item $(v, \, k, \, \lambda)=\left(2^m, \ 2^{m-1}, \ (2^{m-2}-1) (9 \times 2^{2m-4}+ 3 \times 2^{m-3} +1) \right).$ \item $(v,\, k, \, \lambda)=\left(2^m,\ 2^{m-1}+2^{\frac{m-1}{2}}, \ \frac{ 2^{\frac{m-1}{2} } \left(2^{m-1} + 2^{\frac{m-1}{2} } -1 \right) \left( 2^{\frac{m-1}{2} } +2 \right)\left( 5 \times 2^{m-3} + 1 \right)}{3} \right).$ \item $(v,\, k, \, \lambda)=\left(2^m,\ 2^{m-1}+2^{\frac{m+1}{2}}, \ \frac{ \left(2^{m-1} + 2^{\frac{m+1}{2}} \right) \left(2^{m-1} + 2^{\frac{m+1}{2}} -1\right) \left(2^{m-1} + 2^{\frac{m+1}{2}} -2\right)}{48} \right).$ \end{itemize} \end{example} \begin{example} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the supports of all codewords of weight $8$ in $\overline{{\mathcal{C}}_{m}^\perp}$ give a $3$-$(2^m, 8, \lambda)$ design, where $$ \lambda=\frac{2^{2(m-1)} - 5 \times 2^{m-1}+34}{30}. $$ \end{example} \begin{proof} By Theorem \ref{thm-3rdcode}, we have $$ \overline{A^\perp}_8=\frac{2^m(2^{m-1}-1)(2^m-1)(2^{2(m-1)} - 5 \times 2^{m-1}+34)}{315}. $$ The desired value of $\lambda$ follows from Theorem \ref{thm-newdesigns2}. \end{proof} \begin{example} Let $m \geq 7$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the supports of all codewords of weight $10$ in $\overline{{\mathcal{C}}_{m}^\perp}$ give a $3$-$(2^m, 10, \lambda)$ design, where $$ \lambda=\frac{(2^{m-1}-4) (2^{m-1}-16) (2^{2(m-1)}-2^{m-1}+28)}{315}. $$ \end{example} \begin{proof} By Theorem \ref{thm-3rdcode}, we have $$ \overline{A^\perp}_{10}=\frac{2^{m-1}(2^{m-1}-1)(2^m-1)(2^{m-1}-4) (2^{m-1}-16) (2^{2(m-1)}-2^{m-1}+28) }{4 \times 14175}. $$ The desired value of $\lambda$ follows from Theorem \ref{thm-newdesigns2}. \end{proof} \begin{example} Let $m \geq 5$ be an odd integer and let ${\mathcal{C}}_m$ be a binary code with the weight distribution of Table \ref{tab-zhou3}. Then the supports of all codewords of weight $12$ in $\overline{{\mathcal{C}}_{m}^\perp}$ give a $3$-$(2^m, 12, \lambda)$ design, where $$ \lambda=\frac{(2^{h-2}-1) (2 \times 2^{5h} - 55 \times 2^{4h} + 647 \times 2^{3h} - 2727 \times 2^{2h} + 11541 \times 2^{h} - 47208)}{2835} $$ and $h=m-1$. \end{example} \begin{proof} By Theorem \ref{thm-3rdcode}, we have $$ \overline{A^\perp}_{12}=\frac{\epsilon^2 (\epsilon^2-1) (\epsilon^2-4) (2 \epsilon^2-1) (2 \epsilon^{10} - 55 \epsilon^8 + 647 \epsilon^6 - 2727 \epsilon^4 + 11541 \epsilon^2 - 47208) }{8 \times 467775}, $$ where $\epsilon=2^{(m-1)/2}$. The desired value of $\lambda$ follows from Theorem \ref{thm-newdesigns2}. \end{proof} \section{Two families of binary cyclic codes with the weight distribution of Table \ref{tab-zhou3}}\label{sec-examplecodes} To prove the existence of the $2$-designs in Section \ref{sec-2designs} and the $3$-designs in Section \ref{sec-3designs}, we present two families of binary codes of length $2^m-1$ with the weight distribution of Table \ref{tab-zhou3}. Let $n=q^m-1$, where $m$ is a positive integer. Let $\alpha$ be a generator of ${\mathrm{GF}}(q^m)^*$. For any $i$ with $0 \leq i \leq n-1$, let $\mathbb{M}_i(x)$ denote the minimal polynomial of $\beta^i$ over ${\mathrm{GF}}(q)$. For any $2 \leq \delta \leq n$, define \begin{eqnarray}\label{eqn-BCHgeneratorPolyn} g_{(q,n,\delta,b)}(x)={\mathrm{lcm}}(\mathbb{M}_{b}(x), \mathbb{M}_{b+1}(x), \cdots, \mathbb{M}_{b+\delta-2}(x)), \end{eqnarray} where $b$ is an integer, ${\mathrm{lcm}}$ denotes the least common multiple of these minimal polynomials, and the addition in the subscript $b+i$ of $\mathbb{M}_{b+i}(x)$ always means the integer addition modulo $n$. Let ${\mathcal{C}}_{(q, n, \delta,b)}$ denote the cyclic code of length $n$ with generator polynomial $g_{(q, n,\delta, b)}(x)$. ${\mathcal{C}}_{(q, n, \delta, b)}$ is called a \emph{primitive BCH code} with \emph{designed distance} $\delta$. When $b=1$, the set ${\mathcal{C}}_{(q, n, \delta, b)}$ is called a \emph{narrow-sense primitive BCH code}. Although primitive BCH codes are not good asymptotically, they are among the best linear codes when the length of the codes is not very large \cite[Appendix A]{DingBook}. So far, we have very limited knowledge of BCH codes, as the dimension and minimum distance of BCH codes are in general open, in spite of some recent progress \cite{Ding,DDZ15}. However, in a few cases the weight distribution of a BCH code can be settled. The following theorem introduces such a case. \begin{theorem}\label{thm-BCHcode} Let $m \geq 5$ be an odd integer and let $\delta=2^{m-1}-1-2^{(m+1)/2}$. Then the BCH code ${\mathcal{C}}_{(2, 2^m-1, \delta, 0)}$ has length $n=2^m-1$, dimension $3m$, and the weight distribution in Table \ref{tab-zhou3}. \end{theorem} \begin{proof} A proof can be found in \cite{DFZ}. \end{proof} It is known that the dual of a BCH code may not be a BCH code. The following theorem describes a family of cyclic codes having the weight distribution of Table \ref{tab-zhou3}, which may not be BCH codes. \begin{theorem} Let $m \geq 5$ be an odd integer. Let ${\mathcal{C}}_m$ be the dual of the narrow-sense primitive BCH code ${\mathcal{C}}_{(2, 2^m-1, 7, 1)}$. Then ${\mathcal{C}}_m$ has the weight distribution of Table \ref{tab-zhou3}. \end{theorem} \begin{proof} A proof can be found in \cite{Kasa69}. \end{proof} \section{Summary and concluding remarks} In this paper, with any binary linear code of length $2^m-1$ and the weight distribution of Table \ref{tab-zhou3}, a huge number of infinite families of $2$-designs and $3$-designs with various block sizes were constructed. These constructions clearly show that the coding theory approach to constructing $t$-designs are in fact promising, and may stimulate further investigations in this direction. It is open if the codewords of a fixed weight in a family of linear codes can hold an infinite family of $t$-designs for some $t \geq 4$. It is noticed that the technical details of this paper are tedious. However, one has to settle the weight distribution of a linear code and the minimum distance of its dual at the same time, if one would like to employ the Assmus-Mattson Theorem for the construction of $t$-designs. Note that it could be very difficult to prove that a linear code has minimum weight $7$. This explains why the proofs of some of the theorems are messy and tedious, but necessary. \section*{Acknowledgments} The research of C. Ding was supported by the Hong Kong Research Grants Council, under Project No. 16300415.
83044e23b964c47a09237f88b03a8e180a05cc03
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Proof of Theorem~\ref{the:gl}}\label{sec:algorithm} In this section, we prove Theorem~\ref{the:gl}. Throughout this section, $G:[0,1]^E \to \mathbb{R}_+$ denotes the multilinear extension of $g$, while $L:[0,1]^E \to \mathbb{R}_+$ and $W:[0,1]^E \to \mathbb{R}_+$ denote the following linear functions: \begin{align*} L({\bm x}) = \sum_{e \in E}{\bm x}(e)\ell(e) \quad \text{and} \quad W({\bm x}) = \sum_{e \in E}{\bm x}(e)w(e). \end{align*} Furthermore, we define $d_g = \max_{e \in E} g(e)$, $d_\ell = \max_{e \in E} \ell(e)$, and $d_{g,\ell} = \max(d_g,d_\ell)$. Note that we have $g(O) \leq nd_g \leq nd_{g,\ell}$, $\ell(O) \leq nd_\ell \leq nd_{g,\ell}$, and $d_{g,\ell} \leq g(O)+\ell(O) \leq 2n d_{g,\ell}$. Recall that $n = |E|$. In Section~\ref{subsec:small}, we argue that we need to deal with ``small'' and ``large'' elements separately in order to guarantee that we get a set satisfying the knapsack constraint after rounding. Our algorithm updates a vector ${\bm x} \in [0,1]^E$ in several iterations and then rounds it. In Section~\ref{subsec:small-continous-greedy}, we present an algorithm that updates ${\bm x}$ by adding a vector supported on small elements. In Section~\ref{subsec:general-continuous-greedy}, we present the entire algorithm that computes the vector ${\bm x}$ by taking large elements into account. Then, in Section~\ref{subsec:rounding}, we describe our rounding procedure. We need to guess several parameters when running the algorithms in Section~\ref{subsec:small-continous-greedy} and~\ref{subsec:general-continuous-greedy}. Our final algorithm with the guessing process is presented in Section~\ref{subsec:final}. \input{small} \input{small-algorithm} \input{general-algorithm} \input{rounding} \input{entire} \section{The Budget Allocation Problem}\label{sec:applications} In this section, we bound the curvature of the submodular function that represents the budget allocation problem, and we confirm that our algorithm can be applied to the budget allocation problem in order to obtain an approximation factor better than $1-1/e$. We formally define the budget allocation problem. The input consists of a bipartite graph with the bipartition $A \cup B$, a weight function $w:A \to [0,1]$, a capacity function $c: A \to \mathbb{N}$, and a probability function $p:A \to [0,1]$. Intuitively speaking, the sets $A$ and $B$ correspond to media channels and customers, respectively. Each edge $(a,b)$ in the bipartite graph represents the potential influence of media channel $a$ on customer $b$. Consider a budget allocation ${\bm b} \in \mathbb{Z}_+^A$ to $A$ with ${\bm b}(a) \leq c(a)$ and $\sum_{a \in A}{\bm b}(a) w(s) \leq 1$. If a node $a$ is allocated a budget of ${\bm b}(a)$, it makes ${\bm b}(a)$ independent trials to activate each adjacent node $b$. The probability that $b$ is activated by $a$ in each trial is $p(a)$. Thus, the probability that $b$ becomes active is \[ 1 - \prod_{a \in \Gamma(b)}p(a)^{{\bm b}(a)}, \] where $\Gamma(b)$ denotes the set of nodes in $A$ adjacent to $b$. Hence, the expected number of activated target nodes is \[ \sum_{b \in B}\Bigl(1 - \prod_{a \in \Gamma(b)}p(a)^{{\bm b}(a)}\Bigr). \] The objective of this problem is to find the budget allocation that maximizes the expected number of activated target nodes. We can recast the problem using a submodular function. For each $a \in A$, let $E_a = \set{(a,i) \mid i \in c(a)}$, and let $E = \bigcup_{a \in A}E_a$. Then, we define $f:2^E \to \mathbb{R}_+$ as \[ f(S) = \sum_{b \in B}\Bigl(1 - \prod_{a \in \Gamma(b)}p(a)^{|S \cap E_a|}\Bigr). \] Further, we define $w':2^E \to [0,1]$ to be $w'((a,i)) = w(a)$. Then, the budget allocation problem is equivalent to maximizing $f(S)$ subject to $w'(S) \leq 1$. We now observe several properties of $f$. \begin{lemma}\label{lem:ba-marginal-gain} Let $S \subsetneq E$ and $(a,i) \in E \setminus S$. Then, \[ f_S((a,i)) = \sum_{b \in B: a \in \Gamma(b)}(1-p(a))\prod_{a' \in \Gamma(b)}p(a')^{|S \cap E_{a'}|}. \] \end{lemma} \begin{proof} For each $b \in B$, we define a function $g^b:2^E \to \mathbb{R}_+$ as $g^b(T) = 1 - \prod_{a \in \Gamma(b)}p(a)^{|T \cap E_a|}$. Note that $f_S((a,i)) = \sum_{b \in B}g^b_S((a,i))$. If $a \not \in \Gamma(b)$, then we clearly have $g^b_S((a,i)) = 0$. If $a \in \Gamma(b)$, then we have \[ g^b_S((a,i)) = (1-p(a))\prod_{a' \in \Gamma(b)}p(a')^{|S \cap E_{a'}|}. \] Summing $g^b_S((a,i))$ over all $b \in B$, we obtain the claim. \end{proof} \begin{corollary}\label{cor:ba-submodular} The function $f$ is submodular. \end{corollary} \begin{proof} From Lemma~\ref{lem:ba-marginal-gain}, it is easy to see that $f_S((a,i)) \geq f_T((a,i))$ holds for $S \subseteq T\subsetneq E$ and $(a,i) \in E \setminus T$. \end{proof} \begin{corollary}\label{cor:ba-curvature} The curvature $c_f$ of $f$ satisfies \[ c_f \leq 1- \min_{a \in A}\min_{b \in B: a \in \Gamma(b)}p(a)^{c(a)-1}\prod_{a' \in \Gamma(b) \setminus \set{a}}p(a')^{c(a')}. \] \end{corollary} \begin{proof} From Lemma~\ref{lem:ba-marginal-gain}, we have \begin{align*} f_{E\setminus (a,i)}((a,i)) & = \sum_{b \in B: a \in \Gamma(b)}(1-p(a))p(a)^{c(a)-1}\prod_{a' \in \Gamma(b) \setminus \set{a}}p(a')^{c(a')}, \\ f((a,i)) & = \sum_{b \in B: a \in \Gamma(b)}(1-p(a)). \end{align*} Hence, \begin{align*} c_f & = 1 - \min_{(a,i) \in E}\frac{f_{E\setminus (a,i)}((a,i))}{f((a,i))}\\ & = 1- \min_{a \in A}\frac{\sum_{b \in B: a \in \Gamma(b)}(1-p(a))p(a)^{c(a)-1}\prod_{a' \in \Gamma(b) \setminus \set{a}}p(a')^{c(a')}}{\sum_{b \in B: a \in \Gamma(b)}(1-p(a))} \\ & \leq 1- \min_{a \in A}\min_{b \in B: a \in \Gamma(b)}p(a)^{c(a)-1}\prod_{a' \in \Gamma(b) \setminus \set{a}}p(a')^{c(a')}. \qedhere \end{align*} \end{proof} From our main result (Theorem~\ref{the:intro}) and Corollaries~\ref{cor:ba-submodular} and~\ref{cor:ba-curvature}, when the capacity of each node $a \in A$ is bounded by a constant and the number of vertices adjacent to each node $b\in B$ is bounded by a constant, we obtain a polynomial-time algorithm whose approximation ratio is strictly better than $1-1/e$. \subsection{Putting things together}\label{subsec:final} Now, we present our entire algorithm. The idea is to simply guess $v_g$, $v_\ell$, $m$, $\set{\gamma^t_i}$, $\set{\lambda_i}$, $\set{\gamma^t_{\rm S}}$, and $\lambda_{\rm S}$, run Algorithm~\ref{alg:greedy-general} with the guessed values, and then round the obtained vectors using Algorithm~\ref{alg:rounding}. Naively, we have $O(|V_{\epsilon,n}(g,\ell)|^{O(1/\epsilon)})=O((\log (n/\epsilon)/\epsilon)^{O(1/\epsilon)} )$ choices for the sequence $\set{\gamma^t_{\rm S}}$. We can decrease the number of choices since $g$ has a bounded curvature. If we have a guess $\gamma_{\rm S}^0$ such that $\gamma_{\rm S}^0 \geq g(O_{\rm S}) \geq (1-\epsilon)\gamma_{\rm S}^0$, then we must have $\gamma_{\rm S}^0 \geq g_{S}(O_{\rm S}) \geq (1-\epsilon)(1-c_g) \gamma_{\rm S}^0$ for any set $S \subseteq E$. Hence, it suffices to consider sequences whose maximum and minimum values are within a factor of $(1-\epsilon)(1-c_g)$. Let $V_{\epsilon,n,\gamma_{\rm S}^0}(g,\ell) := \set{v \in V_{\epsilon,n}(g,\ell) \mid v \geq (1-\epsilon)(1-c_g)\gamma_{\rm_S}^0}$. Then, the number of such sequences is at most $|V_{\epsilon,n}(g,\ell)| \cdot |V_{\epsilon,n,\gamma_{\rm S}^0}(g,\ell)|^{O(1/\epsilon)}$, which is much smaller than $O((\log (n/\epsilon)/\epsilon)^{O(1/\epsilon)})$. A detailed description of our algorithm is given in Algorithm~\ref{alg:knapsack}. \begin{algorithm}[t!] \caption{$\textsc{Knapsack}$}\label{alg:knapsack} \begin{algorithmic}[1] \Require{A monotone submodular function $g:2^E \to \mathbb{R}_+$, a linear function $\ell:2^E \to \mathbb{R}$, a weight function $w :E \to \mathbb{R}_+$, and $\epsilon \in (0,1)$.} \Ensure{A set $S \subseteq E$ satisfying $w(S) \leq 1$.} \For{each choice of $v_g,v_\ell \in V_{\epsilon,n}(g,\ell)$} \State{$E_{\rm L} \leftarrow$ the set of large elements with respect to $v_g$ and $v_\ell$.} \State{$E_{\rm S} \leftarrow$ the set of small elements with respect to $v_g$ and $v_\ell$.} \State{$\mathcal{S} \leftarrow \emptyset$.} \State{$M := \lfloor \frac{1}{(1-c_g)\epsilon^6}\rfloor$} \For{each choice of $m$ from $\set{0,1,\ldots,M}$} \For{each choice of $\set{\gamma^t_i},\set{\lambda_i}$ from $ V_{\epsilon,m}(g,\ell)/m$} \For{each choice of $\gamma^0_{\rm S}, \lambda_{\rm S}$ from $V_{\epsilon,n}(g,\ell)$} \For{each choice of $\set{\gamma_{\rm S}^\epsilon,\ldots,\gamma_{\rm S}^{1-\epsilon}}$ from $V_{\epsilon,n,\gamma_{\rm S}^0}(g,\ell)$} \State{$({\bm y}_1,\ldots,{\bm y}_m,{\bm z}) \leftarrow \textsc{GuessingContinuousGreedy}_{\epsilon,\epsilon}(g,\ell,w,E_{\rm L},E_{\rm S},m,\set{\gamma^t_i}, \set{\lambda_i},\set{\gamma^t_{\rm S}},\lambda_{\rm S})$.} \State{$S \leftarrow \textsc{Rounding}_\epsilon(w,E_{\rm L}, E_{\rm S}, m,\set{{\bm y}_i},{\bm z})$.} \State{$\mathcal{S} \leftarrow \mathcal{S} \cup \set{S}$.} \EndFor \EndFor \EndFor \EndFor \EndFor \State{\Return $\arg\max_{S \in \mathcal{S}}g(S)+\ell(S)$.} \end{algorithmic} \end{algorithm} \begin{proof}[Proof of Theorem~\ref{the:gl}] Consider the case that $v_g$ and $v_\ell$ satisfy~\eqref{eq:guess-g(O)-and-l(O)}, $m = |O_{\rm L}|$, and $\set{\gamma^t_i},\set{\lambda_i}$, $\set{\gamma^t_{\rm S}}$, and $\lambda_{\rm S}$ are good guesses. Let $S$ be the (random) set obtained with these guesses. By Lemma~\ref{lem:rounding}, we have \begin{align} \mathop{\mathbf{E}}[g(S) + \ell(S)] \geq (1-\epsilon)(\widehat{G}({\bm x}) + L({\bm x})) - O(\epsilon)(g(O)+\ell(O) + v_g + v_\ell). \label{eq:g(S)+l(S)} \end{align} Conditioned on the event that \textsf{GuessingContinuousGreedy} succeeds, by (i) and (ii) of Lemma~\ref{lem:continuous-greedy}, we get \begin{align} \eqref{eq:g(S)+l(S)} & \geq (1-\epsilon)(1-1/e-O(\epsilon))g(O) + (1-\epsilon)(1-O(\epsilon))\ell(O) - O(\epsilon)(g(O)+\ell(O)+v_g+v_\ell + 8 d_{g,\ell}) \nonumber \\ & \geq (1-1/e)g(O) + \ell(O) - O(\epsilon)(g(O)+\ell(O)). \label{eq:g(S)+l(S)-good} \end{align} Since \textsf{GuessingContinuousGreedy} succeeds with probability at least $1-\epsilon$, we get \begin{align*} \mathop{\mathbf{E}}[g(S) + \ell(S)] & \geq (1-\epsilon)\cdot \eqref{eq:g(S)+l(S)-good} \geq (1-1/e)g(O) + \ell(O) - O(\epsilon)(g(O)+\ell(O)). \end{align*} Since Algorithm~\ref{alg:knapsack} outputs the set with the maximum objective, we have the desired property on the objective value. It is clear that the output of Algorithm~\ref{alg:knapsack} has weight at most $1$ because \textsf{Rounding} always outputs a set of weight at most $1$. For arbitrary $\gamma \in V_{\epsilon,n}(g,\ell)$, the time complexity of Algorithm~\ref{alg:knapsack} is \begin{align*} & O\Bigl(\frac{nM^2}{\epsilon^3}\log \frac{n M}{\epsilon}+\frac{n^4}{\epsilon}+\frac{n^2}{\epsilon^3}\log \frac{1}{\epsilon}\Bigr) \cdot |V_{\epsilon,n}(g,\ell)|^{O(1)} \cdot |V_{\epsilon,n,\gamma}(g,\ell)|^{O(1/\epsilon)} \cdot |V_{\epsilon,m}(g,\ell)|^{O(M/\epsilon)}\\ & = O\Bigl(\frac{nM^2}{\epsilon^3}\log \frac{n M}{\epsilon}+\frac{n^4}{\epsilon}+\frac{n^2}{\epsilon^3}\log \frac{1}{\epsilon}\Bigr) \cdot \Bigl(\frac{\log (n/\epsilon)}{\epsilon} \Bigr)^{O(1)}\cdot \Bigl(\frac{1}{\epsilon}\log \frac{1}{1-c_g} \Bigr)^{O(1/\epsilon)} \cdot \Bigl(\frac{\log (M/\epsilon)}{\epsilon} \Bigr)^{O(M/\epsilon)}\\ & = O\Bigl(\frac{n}{(1-c_g)^2\epsilon^{15}}\log \frac{n}{(1-c_g)\epsilon}+\frac{n^4}{\epsilon}+\frac{n^2}{\epsilon^3}\log \frac{1}{\epsilon}\Bigr) \cdot \Bigl(\frac{\log n}{\epsilon} \Bigr)^{O(1)} \cdot \\ & \qquad \Bigl(\frac{1}{\epsilon}\log \frac{1}{1-c_g} \Bigr)^{O(1/\epsilon)} \cdot \Bigl(\frac{1}{\epsilon}\log \frac{1}{1-c_g} \Bigr)^{O(1/((1-c_g)\epsilon^7))}\\ & = O\Bigl(\frac{n^4\polylog(n)}{(1-c_g)^2}\Bigr) \cdot \Bigl(\frac{1}{\epsilon}\log \frac{1}{1-c_g} \Bigr)^{\poly(1/\epsilon)/(1-c_g)}. \end{align*} Hence, we have the desired time complexity. By replacing $\epsilon$ with $\epsilon/C$ for a large constant $C$ (to change $O(\epsilon)$ to $\epsilon$), we have the desired result. \end{proof} \subsection{Continuous greedy algorithm with guessing}\label{subsec:general-continuous-greedy} In this section, we present an algorithm whose goal is to output a vector ${\bm x} \in [0,1]^E$ such that (i) $G({\bm x}) \geq (1-1/e)g(O)$, (ii) $L({\bm x}) \geq \ell(O)$, and (iii) $W({\bm x}) \leq w(O)$. Our algorithm is a variant of the continuous greedy algorithm~\cite{Calinescu:2011ju} but differs in the following aspects: we consider two functions $g$ and $\ell$, and we handle large and small elements separately. Let $m$ be an integer given as a parameter, which is a guessed value of $|O_{\rm L}|$ (with respect to the current values of $v_g$ and $v_\ell$). Then, we make copies $E_1,\ldots,E_{m}$ of $E$ and define a set $\widehat{E} = \bigcup_{i \in [m]}E_i \cup E_{\rm S}$. Then, we define a function $\widehat{g} : 2^{\widehat{E}} \rightarrow \mathbb{R}_+$ as $\widehat{g}(S_1,S_2,\ldots,S_{m},S_{\rm S}) = g(S_1 \cup \cdots \cup S_{m} \cup S_{\rm S})$. We note that $\widehat{g}$ is a monotone submodular function. Let $\widehat{G}$ be the multilinear extension of $\widehat{g}$. We introduce a vector ${\bm y}_i \in [0,1]^E$ for each $i \in [m]$ and another vector ${\bm z} \in [0,1]^E$. We always guarantee that ${\bm y}_i\;(i \in [m])$ is supported on $E_i$ and ${\bm z}$ is supported on $E_{\rm S}$. Our algorithm runs in $1/\epsilon$ iterations and updates the vectors ${\bm y}_i\;(i \in [m])$ and ${\bm z}$ in each iteration. Here, we assume that $1/\epsilon$ is an integer; otherwise, we slightly decrease $\epsilon$. The final output is the sequence of vectors $({\bm y}_1,\ldots,{\bm y}_{m},{\bm z})$, and their sum ${\bm x} := \sum_{i \in [m]}{\bm y}_i + {\bm z}$ will satisfy the conditions stated initially in this section. We call the first iteration the iteration at time $0$, the second iteration the iteration at time $\epsilon$, and so on. To explain how we update the vectors, we introduce several notations. For $t = \set{0,\epsilon,\ldots,1}$, we define ${\bm y}^t_i\;(i \in [m])$ and ${\bm z}^t$ as the vectors ${\bm y}_i$ and ${\bm z}$ immediately before the iteration at time $t$. We note that ${\bm y}^0_i = {\bm 0}\;(i \in [m])$ and ${\bm z}^0 = {\bm 0}$ hold. We define ${\bm y}^1_i\;(i \in [m])$ and ${\bm z}^1$ as ${\bm y}_i\;(i \in [m])$ and ${\bm z}$, respectively, after the iteration at time $1-\epsilon$. Note that the algorithm outputs the sequence of vectors $({\bm y}^1_1,\ldots,{\bm y}^1_m,{\bm z}^1)$. Then, we define ${\bm x}^t = \sum_{i \in [m]}{\bm y}^t_i + {\bm z}^t$. Further, for $t \in \set{0,\epsilon,\ldots,1-\epsilon}$ and $i \in \set{0,1,\ldots,m}$, we define ${\bm x}^t_i = \sum_{j \leq i}{\bm y}^{t+ \epsilon}_j + \sum_{j > i}{\bm y}^t_j + {\bm z}^t$, i.e., the vector obtained after the iteration at time $t-\epsilon$ followed by updating ${\bm y}_1,\ldots,{\bm y}_i$. Note that ${\bm x}^t_0 = {\bm x}^t$. As in the argument in Section~\ref{subsec:small}, we try all possible values for guessing $|O_{\rm L}|$. Hence, in what follows, we assume that the guessed value $m$ is correct, i.e., $m = |O_{\rm L}|$. Let $o_1,\ldots,o_{m}$ be the large elements in $O$, i.e., $O_{\rm L} = \set{o_1,\ldots,o_m}$. For $i \in [m]$, let $\widehat{o}_i$ be the copy of $o_i$ in $E_i$. Then, we define $\widehat{O}_{\rm L} = \set{\widehat{o}_1,\ldots,\widehat{o}_m}$ and $\widehat{O} = \widehat{O}_{\rm L} \cup O_{\rm S} \subseteq \widehat{E}$. For each $i \in [m]$, we update the vector ${\bm y}_i^t$ to ${\bm y}_i^{t+\epsilon}$ by finding an element $e_i^t \in E_i$ and adding the vector $\epsilon {\bm 1}_{e_i^t}$. Here, we want the element $e_i^t$ to satisfy (i) $\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{i-1})}(e_i^t)] \geq \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{i-1})}(\widehat{o}_i)]$, (ii) $\ell(e_i^t) \geq \ell(o_i)$, and (iii) $w(e_i^t) \leq w(o_i)$. As we do not know the values of $\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{i-1})}(\widehat{o}_i)]$ and $\ell(o_i)$, the algorithm requires their guessed values $\gamma_i^t$ and $\lambda_i$, respectively. We do not have to guess $w(o_i)$ because we will choose the element with the minimum weight satisfying (i) and (ii). Then, we update the vector ${\bm z}^t$ to ${\bm z}^{t+\epsilon}$ by finding a vector ${\bm v}^t$ and adding the vector $\epsilon {\bm v}^t$. Here, we want the vector ${\bm v}$ to satisfy (i) $\sum_{e \in E_{\rm S}}{\bm v}(e)\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x})}(e)] \geq \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{i-1})}(O_S)]$ (note that $O_S \subseteq E_S \subseteq \widehat{E}$), (ii) $L({\bm v}) \geq \ell(O_{\rm S})$, and (iii) $W({\bm v}) \leq w(O_{\rm S})$. Such a vector can be found by calling \textsc{SmallElements} with the guessed values $\gamma^t_{\rm S}$ and $\lambda_{\rm S}$ for $\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{i-1})}(O_S)]$ and $\ell(O_{\rm S})$, respectively. A detailed description of the algorithm is given in Algorithm~\ref{alg:greedy-general}. We will show that $\widehat{G}({\bm x}^{t+\epsilon})-G({\bm x}^t) \geq \epsilon (g(O) - \widehat{G}({\bm x}^{t+\epsilon}))$, which is sufficient to show that $\widehat{G}({\bm x}^{1})$ is close to the $(1-1/e)$-approximation to $g(O)$. We will also show that $L({\bm x}^1) \geq \ell(O)$ and $W({\bm x}) \leq w(O)$. \begin{algorithm}[t!] \caption{$\textsc{GuessingContinuousGreedy}_{\epsilon,\delta}(g,\ell,w,E_{\rm L}, E_{\rm S},m,\set{\gamma^t_i}, \set{\gamma^t_{\rm S}}, \set{\lambda_i},\lambda_{\rm S})$}\label{alg:greedy-general} \begin{algorithmic}[1] \Require{A monotone submodular function $g:2^E \to \mathbb{R}_+$, a monotone linear function $\ell:2^E \to \mathbb{R}$, a weight function $w :E \to [0,1]$, $\epsilon,\delta\in (0,1)$, a set of large and small elements $E_{\rm L}$ and $E_{\rm S}$, an integer $m\in \mathbb{N}$, guessed values $\set{\gamma^t_i}_{i\in [m],t \in \set{0,\epsilon,\ldots,1-\epsilon}}$, $\set{\gamma^t_{\rm S}}_{t \in \set{0,\epsilon,\ldots,1-\epsilon}}$, $\set{\lambda_i}_{i \in [m]}$, and $\lambda_{\rm S}$.} \Ensure{A vector ${\bm x} \in [0,1]^E$.} \State{${\bm y}_i\leftarrow {\bm 0}\in [0,1]^{\widehat{E}}$ for $i \in [m]$ and ${\bm z}\leftarrow {\bm 0}\in [0,1]^{\widehat{E}}$.} \For{$(t\leftarrow 0$; $t \leq 1-\epsilon$; $t\leftarrow t+\epsilon$)} \For{($i \leftarrow 1$; $i \leq m$; $i \leftarrow i + 1$)} \State{${\bm \theta}_i(e) \leftarrow \textsc{Estimate}_{\epsilon,\epsilon/m,\epsilon\delta/(2nm)}(\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x})}](e))$ for each $e \in E_i$.} \State{Let $e = \arg\min\set{w(e) \mid e \in E_i, {\bm \theta}_i(e) \geq (1-\epsilon)\gamma^t_i - \epsilon d/m, \ell(e) \geq \lambda_i}$.} \State{${\bm y}_i \leftarrow {\bm y}_i + \epsilon {\bm 1}_e$.} \EndFor \State{${\bm v} \leftarrow \textsc{SmallElements}_{\epsilon,\epsilon\delta/2}(\widehat{g},\ell,w,E_{\rm S},\gamma^t_{\rm S},\lambda_{\rm S},\sum_{i\in[m]}{\bm y}_i + {\bm z})$.} \State{${\bm z} \leftarrow {\bm z} + \epsilon {\bm v}$.} \EndFor \State{\Return $({\bm y}_1,\ldots,{\bm y}_{m},{\bm z})$.} \end{algorithmic} \end{algorithm} For $t \in \set{0,\epsilon,\ldots,1-\epsilon}$ and $i \in [m]$, let ${\bm \theta}_i^t$ be the ${\bm \theta}_i$ used in the iteration at time $t$. From Lemma~\ref{lem:chernoff} and the union bound, we immediately have the following: \begin{proposition}\label{pro:general-chernoff} With probability at least $1-\delta/2$, we have \begin{align*} (1- \epsilon )\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(e)] - \frac{\epsilon d_{g,\ell}}{m} & \leq {\bm \theta}^t_i(e) \leq (1 + \epsilon )\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(e)] + \frac{\epsilon d_{g,\ell}}{m} \end{align*} for every $t \in \set{0,\epsilon,\ldots,1-\epsilon}$, $i \in [m]$, and $e \in E_i$. \end{proposition} We formalize the concept that $\gamma_i^t$ and $\lambda_i$ are sufficiently accurate. \begin{definition}\label{def:good-guess-large} For $t \in \set{0,\epsilon,\ldots,1-\epsilon}$ and $i \in [m]$, we say that $\gamma^t_i$ is a \emph{good guess} if \[ \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(\widehat{o}_i)] \geq \gamma^t_i \geq (1-\epsilon) \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(\widehat{o}_i)] - \frac{\epsilon d_{g,\ell}}{m} \] holds. For $i \in [m]$, we say that $\lambda_i$ is a \emph{good guess} if \[ \ell(o_i) \geq \lambda_i \geq (1-\epsilon) \ell(o_i) - \frac{\epsilon d_{g,\ell}}{m} \] holds. \end{definition} Since $\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(o_i)] \leq d_{g,\ell}$ and $\ell(o_i) \leq d_{g,\ell}$ hold for every $i \in [m]$, we can find good guesses by trying all the values in the set $V_{\epsilon,m}(g,\ell)/m := \set{v / m\mid v \in V_{\epsilon,m}(g,\ell)}$. \begin{lemma}\label{lem:large-guess} Suppose that the consequence of Proposition~\ref{pro:general-chernoff} holds and that $\set{\gamma_i^t}$ and $\set{\lambda_i}$ are good guesses. Then, for every $t \in \set{0,\epsilon,\ldots,1-\epsilon}$, we have the following: \begin{itemize} \item[(i)] $\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(e^t_i)] \geq (1 - \epsilon)^3\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(\widehat{o}_i)] - 3\epsilon d_{g,\ell}/m$, \item[(ii)] $\ell(e^t_i) \geq (1-\epsilon)\ell(o_i) - \epsilon d_{g,\ell}/m$ for $i \in [m]$, and \item[(iii)] $w(e^t_i) \leq w(o_i)$ for $i \in [m]$. \end{itemize} \end{lemma} \begin{proof} Fix $t \in \set{0,\epsilon,\ldots,1-\epsilon}$ and $i \in [m]$. Note that we have \[ {\bm \theta}^t_i(o_i) \geq (1-\epsilon)\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(\widehat{o}_i)]-\frac{\epsilon d_{g,\ell}}{m} \geq (1-\epsilon)\gamma^t_i- \frac{\epsilon d_{g,\ell}}{m} \] and $\ell(o_i) \geq \lambda_i$. Since $o_i$ (in $E_i$) is a candidate for $e^t_i$, the element $e^t_i$ is well defined. In particular, we have $w(e^t_i) \leq w(o_i)$ because $e^t_i$ is chosen as the element with the minimum element satisfying the conditions. We have \[ (1+\epsilon)\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t}_{i-1})}(e^t_i)]+\frac{\epsilon d_{g,\ell}}{m} \geq {\bm \theta}^t_i(e^t_i) \geq (1-\epsilon)\gamma^t_i - \frac{\epsilon d_{g,\ell}}{m} \geq (1-\epsilon)^2\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_i)}(o_i)]-\frac{2\epsilon d_{g,\ell}}{m}. \] Rearranging this inequality, we get (i). Further, (ii) is immediate from the fact that $\ell(e_i^t) \geq \lambda_i$. \end{proof} We say that $\gamma_{\rm S}^t$ for $t \in \set{0,\epsilon,\ldots,1-\epsilon}$ and $\lambda_{\rm S}$ are \emph{good guesses} if they are good guesses in the sense of Definition~\ref{def:good-guess-small}. Then, we have the following: \begin{lemma}\label{lem:continuous-greedy} Suppose that $\set{\gamma_i^t},\set{\lambda_i}$, $\set{\gamma_{\rm S}^t}$, and $\lambda_{\rm S}$ are good guesses. Then, Algorithm~\ref{alg:greedy-general} returns vectors ${\bm y}_1,\ldots,{\bm y}_{m},{\bm z}$ such that ${\bm x} := \sum_{i \in [m]}{\bm y}_i + {\bm z}$ satisfies the following: \begin{itemize} \item[(i)] $\widehat{G}({\bm x}) \geq (1-1/e-O(\epsilon))g(O)-6\epsilon d_{g,\ell}$, \item[(ii)] $L({\bm x}) \geq (1-O(\epsilon))\ell(O)- 2 \epsilon d_{g,\ell}$, and \item[(iii)] $W({\bm x}) \leq w(O)$, \end{itemize} with probability at least $1-\delta$. The running time is $O(\frac{nm^2}{\epsilon^3}\log \frac{n m}{\epsilon \delta}+\frac{n^4}{\epsilon}+\frac{n^2}{\epsilon^3}\log \frac{1}{\epsilon \delta})$. \end{lemma} \begin{proof} With probability $1-\delta/2$, the consequence of Proposition~\ref{pro:general-chernoff} holds. Further, with probability $1-\delta/2$, all the invocations of \textsc{SmallElements} succeed in outputting vectors with the guarantees in Lemma~\ref{lem:small-element-guess}. By the union bound, all these occur with probability at least $1-\delta$. In what follows, we assume that this occurs. First, we check~(i). For each $t \in \set{0,\epsilon,\ldots,1-\epsilon}$, we have \begin{align*} & \widehat{G}({\bm x}^t_{m}) - \widehat{G}({\bm x}^t) = \sum_{j=1}^{m}\Bigl(\widehat{G}({\bm x}^t_{j-1} + \epsilon {\bm 1}_{e^t_j}) - \widehat{G}({\bm x}^t_{j-1})\Bigr)\\ & \geq \epsilon\sum_{j=1}^{m}\Bigl(\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{j-1})}(e^t_j)]\Bigr) \tag{By concavity of $\widehat{G}$}\\ & \geq \epsilon(1 - \epsilon)^3\sum_{j=1}^{m}\Bigl(\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{j-1})}(\widehat{o}_j)] - \frac{3\epsilon d_{g,\ell}}{m}\Bigr) \tag{By (i) of Lemma~\ref{lem:large-guess}}\\ & \geq \epsilon(1 - \epsilon)^3\sum_{j=1}^{m} \Bigl(\mathop{\mathbf{E}}\bigl[\widehat{g}_{R({\bm x}^t_{m}) \cup \set{\widehat{o}_k \mid k \in [j-1] }}(\widehat{o}_j) \bigr]\Bigr) -3\epsilon^2 d_{g,\ell} \\ & = \epsilon(1 - \epsilon)^3\Bigl(\mathop{\mathbf{E}}\bigl[\widehat{g}(R({\bm x}^{t}_{m}) \cup \widehat{O}_{\rm L}) - \widehat{g}(R({\bm x}^{t}_{m}))\bigr]\Bigr)- 3\epsilon^2 d_{g,\ell} \\ & \geq \epsilon(1 - \epsilon)^3(\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{m})}(\widehat{O}_{\rm L})] ) - 3\epsilon^2 d_{g,\ell}. \end{align*} For each $t \in \set{0,\epsilon,\ldots,1-\epsilon}$, we have \begin{align*} & \widehat{G}({\bm x}^{t+\epsilon}) - \widehat{G}({\bm x}^t_{m}) = \widehat{G}({\bm x}^{t}_{m} + \epsilon {\bm v}) - \widehat{G}({\bm x}^t_{m}) \\ & \geq \epsilon \sum_{e \in E}{\bm v}(e) \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t+\epsilon})}(e)] \tag{by Lemma~\ref{lem:discretization}} \\ & \geq \epsilon \bigl((1-\epsilon)^3 \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t+\epsilon})}(O_{\rm S})] - 3 \epsilon d_{g,\ell} \bigr) \tag{by (i) of Lemma~\ref{lem:small-element-guess}}\\ & \geq \epsilon (1-\epsilon)^3 \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t+\epsilon})}(O_{\rm S})] - 3\epsilon^2 d_{g,\ell}. \end{align*} Combining these two inequalities, we get \begin{align*} & \widehat{G}({\bm x}^{t+\epsilon}) - \widehat{G}({\bm x}^t) \\ & \geq \epsilon(1 - \epsilon)^3(\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^t_{m})}(\widehat{O}_{\rm L})] )- 3\epsilon^2 d_{g,\ell} + \epsilon (1-\epsilon)^3 \mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t+\epsilon})}(O_{\rm S} )] - 3 \epsilon^2 d_{g,\ell}\\ & \geq \epsilon (1 - \epsilon)^3\mathop{\mathbf{E}}[\widehat{g}_{R({\bm x}^{t+\epsilon})}(\widehat{O})] - 6 \epsilon^2 d_{g,\ell}\\ & \geq \epsilon (1 - \epsilon)^3\bigl(\widehat{g}(\widehat{O}) - \widehat{G}({\bm x}^{t+\epsilon})\bigr) - 6\epsilon^2 d_{g,\ell} = \epsilon (1 - \epsilon)^3\bigl(g(O) - \widehat{G}({\bm x}^{t+\epsilon})\bigr) - 6\epsilon^2 d_{g,\ell}. \end{align*} Rewriting the above inequality, we get \[ g(O) - \frac{\beta}{\alpha} - \widehat{G}({\bm x}^{t + \epsilon}) \leq \frac{g(O) - \frac{\beta}{\alpha} - \widehat{G}({\bm x}^t)}{1+\alpha}, \] where $\alpha = \epsilon(1 - \epsilon)^3$ and $\beta = 6\epsilon^2 d_{g,\ell}$. Then, by induction, we can prove that \[ g(O) - \frac{\beta}{\alpha} - \widehat{G}({\bm x}^t) \leq \frac{1}{(1+\alpha)^{t/\epsilon}}\Bigl(g(O) - \frac{\beta}{\alpha}\Bigr). \] Substituting $t = 1$ and rewriting once again, we get \[ \widehat{G}({\bm x}^1) \geq \Bigl(1 - \frac{1}{(1+\alpha)^{1/\epsilon}}\Bigr)(g(O)- \frac{\beta}{\alpha}) \geq \Bigl(1-\frac{1}{e}-O(\epsilon )\Bigr)\Bigl(g(O)-\frac{\beta}{\alpha}\Bigr) \geq \Bigl(1-\frac{1}{e}-O(\epsilon )\Bigr)g(O) - 6\epsilon d_{g,\ell}, \] assuming that $\epsilon$ is sufficiently small, say, less than $1/2$. To see (ii), we have for $t \in \set{0,\ldots,1-\epsilon}$ that \begin{align*} L({\bm x}^{t+\epsilon}) - L({\bm x}^{t}) & = L\Bigl({\bm x}^t + \epsilon \sum_{i=1}^{m} {\bm 1}_{e^t_i} + \epsilon {\bm v}\Bigr) - L({\bm x}^t) \\ & = \epsilon \sum_{i=1}^{m}\ell(e^t_i) + \epsilon (1-\epsilon)\ell(O_{\rm S}) - \epsilon^2 d_{g,\ell} \tag{by (ii) of Lemma~\ref{lem:small-element-guess}}\\ & \geq \epsilon((1-\epsilon)\ell(O_{\rm L})-\epsilon d_{g,\ell}) + \epsilon (1-\epsilon)\ell(O_{\rm S}) - \epsilon^2 d_{g,\ell} \tag{by (ii) of Lemma~\ref{lem:large-guess}}\\ & \geq \epsilon(1-\epsilon)\ell(O)- 2\epsilon^2 d_{g,\ell}. \end{align*} By induction, we get $L({\bm x}^1) \geq (1-\epsilon)\ell(O) - 2\epsilon d_{g,\ell}$. To see (iii), we have for $t \in \set{0,\ldots,1-\epsilon}$ that \begin{align*} W({\bm x}^{t+\epsilon}) - W({\bm x}^t) & = W\Bigl({\bm x}^t + \epsilon \sum_{i=1}^{m} {\bm 1}_{e^t_i} + \epsilon {\bm v}\Bigr) - W({\bm x}^t) \\ & \leq \epsilon (w(O_{\rm L})+ w(O_{\rm S}) ) \tag{By (iii)'s of Lemmas~\ref{lem:small-element-guess} and~\ref{lem:large-guess}}\\ & = \epsilon w(O). \end{align*} By induction, we get $W({\bm x}^1) \leq w(O)$. Finally, we analyze the time complexity. For estimating the ${\bm \theta}_i$'s, we need $O(\frac{nm}{\epsilon} \cdot \frac{m}{\epsilon^2}\log \frac{nm}{\epsilon \delta}) = O(\frac{nm^2}{\epsilon^3}\log \frac{n m}{\epsilon \delta})$ time. The time complexity of \textsc{SmallElements} is at most $O(\frac{1}{\epsilon} \cdot \bigl(n^4+\frac{n^2}{\epsilon^2}\log \frac{1}{\epsilon \delta}) = O(\frac{n^4}{\epsilon}+\frac{n^2}{\epsilon^3}\log \frac{1}{\epsilon \delta})$. Hence, the running time is as desired. \end{proof} \section{Introduction} In this paper, we consider the problem of maximizing a monotone submodular function under a knapsack constraint. Specifically, given a monotone submodular function $f:2^E \to \mathbb{R}_+$ and a weight function $w:E \to [0,1]$, we aim to solve the following optimization problem: \begin{align*} \text{maximize }f(S) \quad \text{subject to } w(S) \leq 1 \text{ and } S \subseteq E, \end{align*} where $w(S) = \sum_{e \in S}w(e)$. This problem has wide applications in machine learning tasks such as sensor placement~\cite{Krause:2008vo}, document summarization~\cite{Lin:2010wpa,Lin:2011wt}, maximum entropy sampling~\cite{Lee:2006cm}, and budget allocation~\cite{Soma:2014tp}. Although this problem is NP-hard in general, it is known that we can achieve $(1-1/e)$-approximation in polynomial time~\cite{Sviridenko:2004hq}, and this approximation ratio is indeed tight~\cite{Feige:1998gx}. Although it is useful to know that we can always obtain $(1-1/e)$-approximation in polynomial time, it is observed that a simple greedy method outputs even better solutions in real applications (see, e.g.,~\cite{Krause:2008vo}), and it is more desirable if we can guarantee a better approximation ratio by making assumptions on the input function. One such assumption is the notion of curvature, introduced by Conforti and Cornu{\'e}jols~\cite{Conforti:1984ig}. For a monotone submodular function $f:2^E \to \mathbb{R}_+$, the \emph{(total) curvature} of $f$ is defined as \[ c_f = 1 - \min_{e \in E}\frac{f_{E-e}(e)}{f(e)}. \] Intuitively speaking, the curvature measures how close $f$ is to a linear function. To see this, note that $c_f \in [0,1]$ and $c_f = 0$ if and only if $f$ is a linear function. It was shown in~\cite{Conforti:1984ig} that, for maximizing a monotone submodular function under a cardinality constraint, the greedy algorithm achieves an approximation ratio $(1-e^{-c_f})/c_f$, and the result was extended to a matroid constraint~\cite{Vondrak:2010zz}. Recently, Sviridenko~\emph{et~al.}~\cite{Sviridenko:2015ur} obtained a polynomial-time algorithm for a matroid constraint with an approximation ratio $1-c_f/e$, and showed that this approximation ratio is indeed tight for every $c_f \in [0,1]$ even under a cardinality constraint (note that $1-c_f/e$ is strictly larger than $(1-e^{-c_f})/c_f$ except when $c_f = 0$ or $c_f = 1$). In this paper, we extend these results to a knapsack constraint and present a polynomial-time algorithm under a knapsack constraint with an approximation ratio $1-c_f/e$. More specifically, we show the following: \begin{theorem}\label{the:intro} There is an algorithm that, given a monotone submodular function $f: 2^E \to \mathbb{R}_+$, a weight function $w: E \to [0,1]$, and $\epsilon \in (0,1)$, outputs a (random) set $S \subseteq E$ with $w(S) \leq 1$ (with probability one) satisfying \[ \mathop{\mathbf{E}}[f(S)] \geq \Bigl(1-\frac{c_f}{e} - \epsilon\Bigr)f(O). \] Here, $O\subseteq E$ is an optimal solution to the problem, i.e., $O$ is a set with $w(O) \leq 1$ that maximizes $f$. The running time is $O\Bigl(n^5+n^4\polylog(n) \cdot \bigl(\frac{1}{\epsilon} \bigr)^{\poly(1/\epsilon)}\Bigr)$, where $n = |E|$. \end{theorem} Note that the approximation ratio $1-c_f/e$ is indeed tight for every $c_f \in [0,1]$ because the lower bound given by~\cite{Sviridenko:2015ur} holds even for a cardinality constraint. To the best of our knowledge, this is the first result for a knapsack constraint that incorporates the curvature to obtain an approximation ratio better than $1-1/e$, which is tight for general submodular functions. We can apply our algorithm to all the above-mentioned applications to obtain a better solution when the input function has a small curvature. As a representative example, we consider the budget allocation problem~\cite{Alon:2012em}, which models a marketing process that allocates a given budget among media channels, such as TV, newspapers, and the Web, in order to maximize the impact on customers. We model the process using a bipartite graph on a vertex set $A \cup B$, where $A$ and $B$ correspond to the media channels and the customers, respectively, and an edge $(a,b) \in A \times B$ represents the potential influence of media channel $a$ on customer $b$. In a simplified setting where we can use each channel at most once, each media channel $a \in A$ can activate a customer with a predetermined probability $p_a \in [0,1]$. Then, we have to find a set $S \subseteq A$ that maximizes the expected number of activated customers subject to $w(S) := \sum_{a \in S}w(a) \leq 1$, where $w(a)$ is the cost of using media channel $a$. We can formulate this problem as the maximization of a monotone submodular function $f:2^A \to \mathbb{R}_+$ under the knapsack constraint $w(S) \leq 1$. We can show that $c_f \leq 1 - \min_{b \in B}p^{|\Gamma(b)|-1}$, where $\Gamma(b)$ denotes the set of neighbors of $b$ in the bipartite graph. By Theorem~\ref{the:intro}, this immediately gives the approximation ratio $1-c_f/e-\epsilon$ for this problem. The actual model is more general and discussed in detail in Section~\ref{sec:applications}. \subsection{Proof technique} Now, we present the outline of our proof. Let $f:2^E \to \mathbb{R}_+$ be the input function and $O \subseteq E$ be the optimal solution, i.e., $O$ is the set that maximizes $f$ among the sets with weight at most one. We assume that $c_f = 1-\Omega(\epsilon)$; otherwise, we can use a standard algorithm~\cite{Sviridenko:2004hq} to achieve the desired approximation ratio. Using the argument in~\cite{Sviridenko:2015ur}, we can decompose the input function $f$ into a monotone submodular function $g:2^E \to \mathbb{R}_+$ and a linear function $\ell:2^E \to \mathbb{R}_+$ such that, if we can compute a set $S \subseteq E$ with $w(S) \leq 1$ and $f(S) = g(S) + \ell(S) \geq (1-1/e)g(O)+\ell(O)$, then $S$ is a $(1-c_f/e)$-approximate solution. Moreover, by slightly changing the argument in~\cite{Sviridenko:2015ur}, we can also assume that $c_g=1-\Omega(\epsilon(1-c_f)) = 1-\Omega(\epsilon^2)$. In order to find the desired set $S \subseteq E$, we use a variant of the continuous greedy algorithm~\cite{Calinescu:2011ju} that simultaneously optimizes $g$ and $\ell$. In this algorithm, we consider continuous versions of $g$, $\ell$, and $w$, denoted by $G:[0,1]^E \to \mathbb{R}_+$, $L:[0,1]^E \to \mathbb{R}_+$, and $W:[0,1]^E \to \mathbb{R}_+$, respectively. We note that the function $G$ is called the multilinear extension of $g$, and that $L$ and $W$ are linear functions. We start with the zero vector ${\bm x} \in [0,1]^E$ and then iteratively update it. The algorithm consists of $1/\epsilon$ iterations, and roughly speaking, in each iteration, we find a vector ${\bm v} \in [0,1]^E$ with the following properties: (i) $G({\bm x} + \epsilon {\bm v}) - G({\bm x}) \geq \epsilon (G({\bm x} \vee {\bm 1}_{O}) - G({\bm x}))$, (ii) $L(\epsilon {\bm v}) = \epsilon L({\bm v}) \geq \epsilon \ell(O)$, and (iii) $W(\epsilon {\bm v}) = \epsilon W({\bm v})\leq w(O)$. Then, we update ${\bm x}$ by adding $\epsilon {\bm v}$. Here, ${\bm 1}_O$ is the characteristic vector of the set $O$ and $\vee$ is the coordinate-wise maximum. Intuitively speaking, these conditions mean that moving along the direction ${\bm v}$ from ${\bm x}$ is no worse than moving towards ${\bm x} \vee {\bm 1}_O$. We can find such a vector ${\bm v}$ by linear programming. Then, after $1/\epsilon$ iterations, we get a vector ${\bm x} \in [0,1]^E$ such that $G({\bm x}) \geq (1-1/e)G({\bm 1}_O) = (1-1/e)g(O)$, $L({\bm x}) \geq \ell(O)$, and $W({\bm x}) \leq w(O)$. Finally, we obtain a set $S \subseteq E$ by rounding the vector ${\bm x}$, where each element $e \in E$ is added with probability ${\bm x}(e)$. Unfortunately, this strategy does not work as is. Here, a crucial issue is that we cannot show the concentration of the weight in the rounding step. To address this issue, by borrowing an idea from~\cite{Badanidiyuru:2013jc}, we split the elements into large and small ones, where an element is said to be \emph{small} if $g(e) \leq \epsilon^6 g(O)$ and $\ell(e) \leq \epsilon^6 \ell(O)$, and is said to be \emph{large} otherwise (in our analysis, it is more convenient to define large and small elements in terms of $g$ and $\ell$ instead of $w$). Then, since the curvature of $g$ is bounded away from one, we can bound the number of large elements in $O$ by a function of $\epsilon$.\footnote{Although it is claimed in~\cite{Badanidiyuru:2013jc} that the number of large elements is bounded for any submodular function, it is not true in general.} Let $O_{\rm L},O_{\rm S} \subseteq O$ be the set of large and small elements in $O$, respectively. Further, we let $O_{\rm L} = \set{o_1,\ldots,o_m}$. Then, in each iteration, we do the following: For each $i \in \set{1,\ldots,m}$, we find an element $e_i$ such that (i) $G({\bm x} \vee {\bm 1}_{e_i}) - G({\bm x}) \geq G({\bm x} \vee {\bm 1}_{o_i}) - G({\bm x})$, (ii) $\ell(e_i) \geq \ell(o_i)$, and (iii) $w(e_i) \leq w(o_i)$. Then, we update ${\bm x}$ by adding $\epsilon{\bm 1}_{e_i}$. Here, ${\bm 1}_e$ is a characteristic vector of the element $e\in E$. Intuitively speaking, adding $e_i$ to the current solution is no worse than adding $o_i$. For small items, we find a vector ${\bm v}$ as before by considering the characteristic vector ${\bm 1}_{O_{\rm S}}$; then, we update ${\bm x}$ by adding $\epsilon {\bm v}$. In the rounding step, we handle large and small elements separately. Note that, for each $i \in \set{1,\ldots,m}$, we have computed $1/\epsilon$ elements (through $1/\epsilon$ iterations). Then, we chose one of them uniformly at random and add it to the output set. An advantage of this rounding procedure is that we can guarantee that the chosen element for $i \in \set{1,\ldots,m}$ has weight at most $w(o_i)$. For small elements, we apply the previous rounding procedure with a minor tweak to guarantee that the output set has weight at most one. In order to realize this idea, we need to address several additional issues. First, as we do not know the set $O$, we do not know values related to $O$, such as $g(O)$, $\ell(O)$, $G({\bm x} \vee {\bm 1}_O)$, and $G({\bm x} \vee {\bm 1}_{o_i})$. Hence, we cannot determine whether an element is small or large, and we cannot find the desired vector or element in each iteration. We address this issue by guessing these values. For example, we can show a lower bound and an upper bound on $g(O)$ that are $O(n)$ times apart. This means that we can find a $(1-\epsilon)$-approximation to $g(O)$ in the geometric sequence of length $O(\log_{1+\epsilon} n)=O((\log n)/\epsilon)$ between the lower and upper bounds. If we naively guess all the values, as we have $1/\epsilon$ iterations, the resulting time complexity will be $\mathrm{poly}(n) \cdot \bigl((\log n)/\epsilon\bigr)^{\mathrm{poly}(1/\epsilon)}$. However, since the function $g$ has curvature $1-\Omega(\epsilon^2)$, we can reduce the number of candidate values and thus improve the time complexity to $\mathrm{poly}(n)\cdot (1/\epsilon)^{\mathrm{poly}(1/\epsilon)}$. \subsection{Related work} As mentioned earlier, it has been shown that the greedy method achieves $(1-e^{-c_f})/c_f$ approximation for a cardinality constraint~\cite{Conforti:1984ig}. The result was extended to a matroid constraint by Vondr{\'a}k\xspace~\cite{Vondrak:2010zz}. He showed that the result actually holds if we replace $c_f$ with the \emph{curvature to the optimum} $c_f^*$, and the approximation ratio $(1-e^{-c_f^*})/c_f^*$ is tight. Sviridenko~\emph{et~al.}~\cite{Sviridenko:2015ur} improved the approximation ratio to $1-c_f/e-\epsilon$ for a matroid constraint (and hence, a cardinality constraint), which is unattainable with $c_f^*$, and showed that the approximation ratio $1-c_f/e$ is tight even for a cardinality constraint. Curvature has been used to explain the empirical performance of the greedy method. Sharma~\emph{et~al.}~\cite{Sharma:2015ur} considered maximum entropy sampling on Gaussian radial basis functions (RBF) kernels, which can be modeled as the maximization of a monotone submodular function, and showed that the curvature of this problem is close to zero. The maximization of a submodular function under a knapsack constraint has been studied extensively. Sviridenko obtained a $(1-1/e)$-approximation algorithm with time complexity $O(n^5)$~\cite{Sviridenko:2004hq}. We can also obtain $(1-1/e-\epsilon)$-approximation with a constant number of knapsack constraints; however, the time complexity blows up to $n^{\mathrm{poly}(1/\epsilon)}$~\cite{Kulik:2013ix}. It has been claimed in~\cite{Badanidiyuru:2013jc} that, for any fixed $\epsilon > 0$, there is a $(1-1/e-\epsilon)$-approximation algorithm with time complexity $\tilde{O}(n^2)$. However, as mentioned in the footnote, their argument has a drawback. Several approximation guarantees have been achieved in~\cite{Iyer:2013tl} using various parameters of the input function. However, none of them has an approximation ratio better than $1-1/e$ based solely on the assumption that the curvature is bounded. \subsection{Organization} The remainder of this paper is organized as follows. Section~\ref{sec:pre} introduces the definitions used throughout the paper and reviews the basic properties of submodular functions. Section~\ref{sec:reduction} explains the reduction to a joint approximation of a monotone submodular function and a monotone linear function. Section~\ref{sec:algorithm} presents a joint approximation algorithm. Section~\ref{sec:applications} describes an application to the budget allocation problem. \section*{Acknowledgments} We thank Takanori Maehara for providing us with the problem and insightful comments. \bibliographystyle{abbrv} \section{Preliminaries}\label{sec:pre} For an integer $n \in \mathbb{N}$, let $[n]$ denote the set $\set{1,\ldots,n}$. In this paper, the symbol $E$ always denotes a (finite) domain of a function. For a function $w:E \to \mathbb{R}$ and a subset $S \subseteq E$, we define $w(S) = \sum_{e \in S}w(e)$. Similarly, for a vector ${\bm x} \in \mathbb{R}^E$ and a set $S \subseteq E$, we define ${\bm x}(S) = \sum_{e \in S}{\bm x}(e)$. For an element $e \in E$, we define ${\bm 1}_e$ as the unit vector whose $e$-th element is $1$. For a set $S \subseteq E$, we define ${\bm 1}_S$ as $\sum_{e \in S}{\bm 1}_e$. Let $f:2^E \to \mathbb{R}$ be a function. For an element $e \in E$, we simply write $f(e)$ to denote $f(\set{e})$. For a set $S \subseteq E$, we define a function $f_S:2^{E} \to \mathbb{R}$ as $f_S(T) = f(S \cup T) - f(S)$. We say that $f$ is \emph{submodular} if, for any $S,T \subseteq E$, \[ f(S) + f(T) \geq f(S \cap T) + f(S \cup T). \] An equivalent condition is the \emph{diminishing return property}, which requires $f_S(e) \geq f_T(e)$ for any $S \subseteq T \subsetneq E$ and $e \in E\setminus T$. We say that $f$ is \emph{linear} if $f(S) = \sum_{e \in S}f(e)$ holds for every $S \subseteq E$. Note that, if $f$ is submodular (resp., linear), then $f_S$ is also submodular (resp., linear). For a vector ${\bm x} \in [0,1]^E$, let $R({\bm x})$ denote a random set, where each element $e\in E$ is included in the set with probability ${\bm x}(e)$. For a submodular function $f:2^E \to \mathbb{R}$, the \emph{multilinear extension} $F:[0,1]^E \to \mathbb{R}$ of $f$ is defined as \[ F({\bm x}) := \mathop{\mathbf{E}}[f(R({\bm x}))] = \sum_{S \subseteq E}f(S)\prod_{e \in S} {\bm x}(e) \prod_{e \in E \setminus S}(1-{\bm x}(e)). \] For an element $e\in E$ and a vector ${\bm x} \in [0,1]^E$, we define $\partial_e F({\bm x})$ as the slope of $F$ at ${\bm x}$ in the direction of ${\bm 1}_e$. The following fact is well known (see, e.g.,~\cite{Feldman:2013Tq}): \begin{align} \partial_e F({\bm x}) = \frac{F({\bm x} \vee {\bm 1}_e) - F({\bm x})}{1-x_e}= \frac{\mathop{\mathbf{E}}[f_{R({\bm x})}(e)]}{1-x_e}.\label{eq:partial-derivative} \end{align} The following lemma bounds the marginal gain of $F$ when adding $\epsilon {\bm y}$ to ${\bm x}$: \begin{lemma}\label{lem:discretization} Let $f:2^E \to \mathbb{R}_+$ be a monotone submodular function and ${\bm x},{\bm y} \in [0,1]^E$ be vectors such that ${\bm x} + \epsilon {\bm y} \in [0,1]^E$. Then, \[ F({\bm x} + \epsilon {\bm y}) - F({\bm x}) \geq \epsilon \sum_{e\in E}{\bm y}(e)\mathop{\mathbf{E}}[f_{R({\bm x}+\epsilon {\bm y})}(e)]. \] \end{lemma} \begin{proof} Let $e_1,\ldots,e_n$ be an arbitrary ordering of elements in $E$. For $i \in \set{0,1,\ldots,n}$, let ${\bm x}^i = {\bm x} + \sum_{j=1}^i \epsilon {\bm y}(e) {\bm 1}_e$. Note that ${\bm x}^0 = {\bm x}$ and ${\bm x}^n = {\bm x}+\epsilon {\bm y}$. Then, we have \begin{align*} F({\bm x} + \epsilon {\bm y}) - F({\bm x}) & = \sum_{i \in [n]} \partial_{e_i} F({\bm x}^{i-1}) \epsilon {\bm y}(e_i) \tag{by multilinearity of $F$}\\ & \geq \sum_{i \in [n]} \mathop{\mathbf{E}}[f_{R({\bm x}^{i-1})}(e)] \epsilon {\bm y}(e_i) \tag{by~\eqref{eq:partial-derivative}}\\ & \geq \epsilon\sum_{i \in [n]}{\bm y}(e_i) \mathop{\mathbf{E}}[f_{R({\bm x}^n)}(e)]. \tag{by submodularity of $f$}\\ & = \epsilon\sum_{i \in [n]}{\bm y}(e_i) \mathop{\mathbf{E}}[f_{R({\bm x}+\epsilon {\bm y})}(e)]. \qedhere \end{align*} \end{proof} We frequently use the following form of Chernoff's bound. \begin{lemma}[Relative+Additive Chernoff's bound~\cite{Badanidiyuru:2013jc}]\label{lem:chernoff} Let $X_1 , \ldots , X_n$ be independent random variables such that $X_i \in [0,1]$ for every $i \in [n]$. Let $X = \frac{1}{n} \sum_{i \in [n]} X_i$ and $\mu = \mathop{\mathbf{E}}[X]$. Then, for any $\alpha \in (0,1)$ and $\beta > 0$, we have \begin{align*} \Pr[|X - \mu| > \alpha \mu+\beta] \leq 2\exp\Bigl(-\frac{n\alpha \beta}{3}\Bigr). \end{align*} \end{lemma} This immediately gives the following sampling algorithm: \begin{corollary}\label{cor:chernoff} Suppose that we can obtain independent samples of a random variable $X$ bounded in $[0,d]$. Let $\mu = \mathop{\mathbf{E}}[X]$. Then, there exists an algorithm, denoted by $\textsc{Estimate}_{\alpha,\beta,\delta}(X)$, that, given $\alpha,\beta,\delta\in(0,1)$, outputs a value $\hat{\mu}$ such that $|\hat{\mu} - \mu| \leq \alpha \mu+\beta d$ with probability at least $1-\delta$. The number of samples used by the algorithm is $O(\log(1/\delta)/(\alpha\beta))$. \end{corollary} \section{Reduction}\label{sec:reduction} In this section, we prove Theorem~\ref{the:intro} using the following theorem, which gives a joint approximation of a monotone submodular function and a monotone linear function. \begin{theorem}\label{the:gl} There is an algorithm that, given a monotone submodular function $g : 2^E \to \mathbb{R}_+$, a monotone linear function $\ell : 2^E \to \mathbb{R}_+$, a weight function $w: E \to [0,1]$, and $\epsilon \in (0,1)$, outputs a (random) set $S \subseteq E$ with $w(S) \leq 1$ satisfying \[ \mathop{\mathbf{E}}[g(S) + \ell(S)] \geq \Bigl(1 - \frac{1}{e}\Bigr) g ( O ) + \ell(O) - \epsilon \bigl(g(O)+\ell(O)\bigr). \] Here, $O = \arg\max_{T\subseteq E:w(T) \leq 1}\bigl(g(T)+\ell(T)\bigr)$ is an optimal solution. The running time is $O\Bigl(\frac{n^4\polylog(n)}{(1-c_g)^2}\Bigr) \cdot \Bigl(\frac{1}{\epsilon}\log \frac{1}{1-c_g} \Bigr)^{\poly(1/\epsilon)/(1-c_g)}$, where $n = |E|$. \end{theorem} The proof of Theorem~\ref{the:gl} is given in Section~\ref{sec:algorithm}. In the remainder of this section, we prove Theorem~\ref{the:intro} using Theorem~\ref{the:gl}. The argument is similar to that used in~\cite{Sviridenko:2015ur}, but it is more subtle here because the running time in Theorem~\ref{the:gl} depends on the curvature of $g$. We use the following lemma. \begin{lemma}[Lemma 2.1 of~\cite{Sviridenko:2015ur}]\label{lem:2.1-of-Sviridenko} If $f : 2^E \to \mathbb{R}_+$ is a monotone submodular function, then $\sum_{e \in E} f_{E-e}(e) \geq (1 - c_f)f(S)$ for all $S \subseteq E$. \end{lemma} \begin{theorem}\label{the:f} There is an algorithm that, given a monotone submodular function $f: 2^E \to \mathbb{R}_+$, a weight function $w: E \to [0,1]$, and $\epsilon \in (0,1)$, outputs a (random) set $S \subseteq E$ with $w(S) \leq 1$ satisfying \[ \mathop{\mathbf{E}}[f(S)] \geq \Bigl(1-\frac{c_f}{e} - \epsilon\Bigr)f(O). \] Here, $O = \arg\max_{T \subseteq E: w(T) \leq 1}f(T)$ is an optimal solution. The running time is $O\Bigl(\frac{n^4}{(1-c_f)^2\epsilon^{17}}\log \frac{n}{(1-c_f)\epsilon}\Bigr) \cdot \Bigl(\frac{\log (n/\epsilon)}{\epsilon} \Bigr)^{O(1/\epsilon)} \Bigl(\frac{\log (1/((1-c_f)\epsilon))}{\epsilon} \Bigr)^{O(1/((1-c_f)\epsilon^8))}$, where $n = |E|$. \end{theorem} \begin{proof} Define the functions $g,\ell:2^E \to \mathbb{R}_+$ such that \begin{align*} \ell(S) = \Bigl(1-\frac{\epsilon}{2}\Bigr)\sum_{e \in S}f_{E-e}(e) \quad \text{and} \quad g(S) = f(S) - \ell(S) \end{align*} for every $S \subseteq E$. It is not hard to see that $\ell$ is a nonnegative monotone linear function and that $g$ is a nonnegative monotone submodular function. Moreover, the curvature of $g$ is \begin{align*} c_g = 1 - \min_{e \in E}\frac{g_{E-e}(e)}{g(e)} = 1 - \min_{e \in E}\frac{f_{E-e}(e) - (1-\epsilon/2)f_{E-e}(e)}{f(e) - (1-\epsilon/2)f_{E-e}(e)} \leq 1 - \frac{\epsilon}{2}\min_{e \in E}\frac{f_{E-e}(e)}{f(e)} = 1-\frac{\epsilon (1-c_f)}{2}. \end{align*} Further, Lemma~\ref{lem:2.1-of-Sviridenko} implies that for any set $S \subseteq E$, \[ \ell(S) = \Bigl(1-\frac{\epsilon}{2}\Bigr)\sum_{e\in S} f_{E-e}(e) \geq \Bigl(1-\frac{\epsilon}{2}\Bigr)(1 - c_f)f(S) \geq \Bigl(1 - c_f-\frac{\epsilon}{2}\Bigr)f(S). \] By applying Theorem~\ref{the:gl} to $g$, $\ell$, $w$, and $\epsilon/2$, we can find a (random) set $S \subseteq E$ with $w(S) \leq 1$ satisfying \begin{align*} \mathop{\mathbf{E}}[f(S)] & = \mathop{\mathbf{E}}[g(S) + \ell(S)] \geq \Bigl(1 - \frac{1}{e}\Bigr)g(O) + \ell(O) - \frac{\epsilon}{2} \bigl( g(O) + \ell(O)\bigr)\\ & = \Bigl(1 - \frac{1}{e}\Bigr) f(O) + \frac{1}{e}\ell(O) - \frac{\epsilon}{2} f(O)\\ & \geq \Bigl(1 - \frac{1}{e}\Bigr) f(O) + \frac{1-c_f-\epsilon/2}{e}f(O) - \frac{\epsilon}{2} f(O)\\ & \geq \Bigl(1 - \frac{c_f}{e} - \epsilon\Bigr) f(O). \end{align*} The running time is clearly as stated. \end{proof} Now, we prove our main theorem. \begin{proof}[Proof of Theorem~\ref{the:intro}] If $c_f < 1-e\epsilon$, then we run the algorithm in Theorem~\ref{the:f}. The approximation factor is $1-c_f/e-\epsilon$ and the running time is $O\Bigl(\frac{n^4}{(1-c_f)^2\epsilon^{17}}\log \frac{n}{(1-c_f)\epsilon}\Bigr) \cdot \Bigl(\frac{\log (n/\epsilon)}{\epsilon} \Bigr)^{O(1/\epsilon)} \Bigl(\frac{\log (1/((1-c_f)\epsilon))}{\epsilon} \Bigr)^{O(1/((1-c_f)\epsilon^8))}=O\Bigl(n^4\polylog(n) \cdot \bigl(\frac{1}{\epsilon} \bigr)^{\poly(1/\epsilon)}\Bigr)$. If $c_f \geq 1-e\epsilon$, then we simply run the $O(n^5)$-time $(1-1/e)$-approximation algorithm presented in~\cite{Sviridenko:2004hq}. Then, the approximation factor is \[ 1-\frac{1}{e} \geq 1-\frac{1-e\epsilon}{e}-\epsilon \geq 1-\frac{c_f}{e}-\epsilon. \] In both cases, the approximation factor is at least $1-c_f/e-\epsilon$ whereas the running time is as desired. \end{proof} \subsection{Rounding}\label{subsec:rounding} In this section, we explain how to round the vectors obtained by \textsc{GuessingContinuousGreedy} (Algorithm~\ref{alg:greedy-general}). Let $({\bm y}_1,\ldots,{\bm y}_m,{\bm z})$ be the vectors obtained by \textsc{GuessingContinuousGreedy}, and let ${\bm v}^{t}$ be the vector supported on $E_{\rm S}$ obtained in the iteration at time $t$ in \textsc{GuessingContinuousGreedy}. Note that ${\bm z} = \sum_{t \in \set{0,\epsilon,\ldots,1-\epsilon}}{\bm v}^t$. Our algorithm is summarized in Algorithm~\ref{alg:rounding}. \begin{algorithm}[t!] \caption{$\textsc{Rounding}_{\epsilon}(w,E_{\rm L}, E_{\rm S},m,\set{{\bm y}_i},{\bm z})$}\label{alg:rounding} \begin{algorithmic}[1] \Require{A weight function $w:E \to [0,1]$, $E_{\rm _L},E_{\rm S} \subseteq E$, a set of vectors $\set{{\bm y}_i}_{i \in [m]}$, and a vector ${\bm z}$.} \Ensure{A set $S \subseteq E$.} \State{$S_{\rm L} \leftarrow \emptyset$, $S_{\rm S} \leftarrow \emptyset$.} \State{Define ${\bm z}' \in [0,1]^{E_{\rm S}}$ as ${\bm z}'(e) = (1-\epsilon){\bm z}(e)$ if $w(e) < \epsilon^3 \max_t W({\bm v}^t )$ and ${\bm z}'(e) = 0$ otherwise.} \State{For each $e \in E_{\rm S}$, add it to $S_{\rm S}$ independently with probability ${\bm z}'(e)$.} \For{$i \in [m]$} \State{Add exactly one element in $E_i$ to $S_{\rm L}$ (as an element of $E$), where an element $e \in E_i$ is chosen with probability ${\bm y}_i(e)$.} \EndFor \If{$w(S_{\rm S} \cup S_{\rm L}) \leq 1$} \State{\Return $S_{\rm S} \cup S_{\rm L}$.} \Else \State{\Return $\emptyset$.} \EndIf \end{algorithmic} \end{algorithm} We use the following lemma to analyze the objective value of the output set. \begin{lemma}[Lemma~3.7 of~\cite{Calinescu:2011ju}]\label{lem:single-rounding} Let $E = E_1 \cup \cdots \cup E_k$, let $f : 2^E \to \mathbb{R}_+$ be a monotone submodular function, and for all $i \neq j$, we have $E_i \cap E_j = \emptyset$. Let ${\bm x} \in \mathbb{R}_+^E$ such that for each $E_i$ we have ${\bm x}(E_i) \leq 1$. If $T$ is a random set where we sample independently from each $E_i$ at most one random element, i.e., element $e$ with probability ${\bm x}(e)$, then \[ \mathop{\mathbf{E}}[f(T)] \geq F({\bm x}). \] \end{lemma} \begin{lemma}\label{lem:rounding-objective-values} We have $\mathop{\mathbf{E}}[g(S_{\rm L} \cup S_{\rm S})] \geq (1 - \epsilon)\widehat{G}({\bm x}) - \epsilon^3 v_g$ and $\mathop{\mathbf{E}}[\ell(S_{\rm L} \cup S_{\rm S})] \geq (1 - \epsilon)L({\bm x}) - \epsilon^3 v_\ell$. \end{lemma} \begin{proof} Let ${\bm x}' = \sum_{i\in [m]}{\bm y}_i + {\bm z}'$. First, let us relate the value of the vector ${\bm x}$ to that of ${\bm x}'$. \begin{align*} \widehat{G}({\bm x}')=\widehat{G}\Bigl(\sum_{i \in [m]} {\bm y}_i + {\bm z}' \Bigr) & \geq \widehat{G}\Bigl(\sum_{i \in [m]} {\bm y}_i + (1-\epsilon){\bm z}\Bigr) - \sum_{e \in E_{\rm S}: w(e) \geq \epsilon^3 \max_t W({\bm v}^t)}{\bm z}(e)g(e) \\ & \geq \widehat{G}\Bigl((1-\epsilon)(\sum_{i \in [m]} {\bm y}_i+{\bm z})\Bigr) - \max_{e\in E_{\rm S}} g(e) \sum_{e \in E_{\rm S}:w(e) \geq \epsilon^3 \max_t W({\bm v}^t)}{\bm z}(e) \\ & \geq (1-\epsilon)\widehat{G}({\bm x})- \epsilon^6 v_g \cdot \frac{1}{\epsilon^3} \geq (1-\epsilon)\widehat{G}({\bm x})-\epsilon^3 v_g. \end{align*} Next, we note that we get $S_{\rm L}$ by selecting exactly one random element from each $E_i$, which is a copy of $E_{\rm L}$, and we get $S_{\rm S}$ by sampling independently from ${\bm v}$. Hence, by applying Lemma~\ref{lem:single-rounding} with sets $E_1,\ldots,E_{m}$ and sets $\set{\set{e} \mid e \in S_{\rm S}}$, we get \[ \mathop{\mathbf{E}}[g(S_{\rm L} \cup S_{\rm S})] \geq (1-\epsilon)\widehat{G}({\bm x}) - \epsilon^3 v_g. \] By a similar argument, we get $L({\bm x}') \geq (1-\epsilon)L({\bm x}) - \epsilon^3 v_\ell$, and we have $\mathop{\mathbf{E}}[\ell(S_{\rm L} \cup S_{\rm S})] \geq (1-\epsilon)L({\bm x})-\epsilon^3 v_\ell$. \end{proof} Next, we show that the probability that the weight of the output set exceeds $w(O)$ decays exponentially. \begin{lemma}\label{lem:rounding-weight} For any $\gamma \geq 1$, we have $w(S_{\rm L} \cup S_{\rm S}) \leq \gamma w(O)$ with probability $1-\exp\bigl(-\Omega(\gamma/\epsilon^2)\bigr)$. \end{lemma} \begin{proof} Recall that, for each $i \in [m]$, the vector ${\bm y}_i$ is the sum of $1/\epsilon$ elements $e^0_i,\ldots,e^{1-\epsilon}_i$, and we pick one of them in Algorithm~\ref{alg:rounding}. By the condition $w(e^t_i) \leq w(o_i)$ for every $i \in [m]$ and $t \in \set{0,\epsilon,\ldots,1-\epsilon}$, the weight of the large elements after the rounding will be less than that of the large elements of the optimal solution. Hence, it is sufficient to prove that $w(S_{\rm S} ) \leq \gamma w(O_{\rm S})$ holds with probability $1-\exp(-\Omega(\gamma/\epsilon^2))$, where $S_{\rm S}$ is the set obtained by rounding ${\bm z}'$. First, note that \[ \mathop{\mathbf{E}}[w(S_{\rm S})] = \mathop{\mathbf{E}}[w(R({\bm z}'))] \leq (1-\epsilon)\mathop{\mathbf{E}}[w(R({\bm z}))] \leq (1-\epsilon)\max_{t} W({\bm v}^t) \leq (1-\epsilon)w(O_{\rm S}). \] For each $e \in E$, we set up a random variable $X_e$ to be $X_e = w(e)/(\epsilon^3w(O_{\rm S}))$ if $e \in S_{\rm S}$ and $X_e = 0$ otherwise. Note that each $X_e$ is bounded in $[0,1]$ because $\max_t W({\bm v}^t) \leq w(O_{\rm S})$. For $X = \sum_{e \in E_{\rm S}}X_e$, we have $\mu := \mathop{\mathbf{E}}[X] = \mathop{\mathbf{E}}[w(S_{\rm S})]/(\epsilon^3w(O_{\rm S})) \leq (1-\epsilon)/\epsilon^3$. Invoking Lemma~\ref{lem:chernoff} with $\alpha=\epsilon/2$ and $\beta=\gamma/(2\epsilon^3)$, we have \begin{align*} \Pr[w(S_{\rm S}) > \gamma w(O_{\rm S})] & = \Pr\Bigl[X \geq \frac{\gamma}{\epsilon^3}\Bigr] \leq \Pr\Bigl[X \geq (1+\alpha)\mu + \beta\Bigr] \\ & \leq 2\exp\Bigl(-\frac{\alpha \beta}{3}\Bigr) = \exp\Bigl(-\Omega\Bigl(\frac{\gamma}{\epsilon^2}\Bigr)\Bigr). \qedhere \end{align*} \end{proof} \begin{lemma}\label{lem:rounding} Algorithm~\ref{alg:rounding} outputs a (random) set $S$ with $w(S) \leq 1$ satisfying \[ \mathop{\mathbf{E}}[g(S)+\ell(S)] \geq (1 - \epsilon)(\widehat{G}({\bm x})+L({\bm x})) - O(\epsilon) \cdot (g(O)+\ell(O)+v_g+v_\ell). \] \end{lemma} \begin{proof} It is clear that we always have $w(S) \leq 1$. Now, we analyze the objective value attained by $S$. For any $\gamma \geq 1$, the probability that $w(S_{\rm L} \cup S_{\rm S}) > \gamma w(O)$ is at most $\exp\bigl(-C\gamma/\epsilon^2\bigr)$ for some $C > 0$ by Lemma~\ref{lem:rounding-weight}. Note that, if $T \subseteq E$ satisfies $w(T) \leq \gamma w(O)$, then $g(T)+\ell(T) \leq \gamma (g(O)+\ell(O))$ from the submodularity of $g+\ell$. By Lemma~\ref{lem:rounding-objective-values}, we have \begin{align*} & \mathop{\mathbf{E}}[g(S) + \ell(S)]\\ & \geq \mathop{\mathbf{E}}[g(S_{\rm L} \cup S_{\rm S})+\ell(S_{\rm L} \cup S_{\rm S})] - \int_1^\infty \gamma(g(O)+\ell(O))\exp(-C\gamma/\epsilon^2) \mathrm{d}\gamma \\ & \geq (1 - \epsilon)\widehat{G}({\bm x}) - \epsilon^3 v_g + (1 - \epsilon)L({\bm x}) - \epsilon^3 v_\ell - \frac{\epsilon^4+C\epsilon^2}{C^2}\exp(-C/\epsilon^2)(g(O)+\ell(O))\\ & = (1 - \epsilon)(\widehat{G}({\bm x})+L({\bm x})) - O(\epsilon) \cdot (g(O)+\ell(O)+v_g+v_\ell). \qedhere \end{align*} \end{proof} \subsection{Subroutine for handling small elements}\label{subsec:small-continous-greedy} Here, we explain a subroutine that finds a vector ${\bm v} \in [0,1]^E$ supported on the set $E_S$ of small elements (with respect to the current guesses $v_g$ and $v_\ell$) in order to update the current vector ${\bm x} \in \mathbb{R}^E$. We want ${\bm v}$ to satisfy the following properties: (i) $\sum_{e \in E_{\rm S}}{\bm v}(e) \mathop{\mathbf{E}}[g_{R({\bm x})}(e)] \geq \mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})]$, (ii) $L({\bm v}) \geq \ell(O_{\rm S})$, and (iii) $W({\bm v}) \leq w(O_{\rm S})$. There are several issues in finding such a vector ${\bm v}$: We cannot exactly calculate $\mathop{\mathbf{E}}[g_{R({\bm x})}(e)]\;(e \in E_{\rm S})$; hence, we need to estimate it. Further, we do not know the values $\mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})]$ and $\ell(O_{\rm S})$. In the subroutine presented here, we assume that their guessed values, denoted by $\gamma$ and $\lambda$, respectively, are given as a part of the input. Once we succeed in accurately estimating $\mathop{\mathbf{E}}[g_{R({\bm x})}(e)]\;(e \in E_{\rm S})$ and the given guessed values $\gamma$ and $\lambda$ are sufficiently accurate, we can find the desired vector ${\bm v}$ by solving a linear program. A detailed description of the subroutine is given in Algorithm~\ref{alg:small-elements}. \begin{algorithm}[t!] \caption{$\textsc{SmallElements}_{\epsilon,\delta}(g,\ell,w,E_{\rm S}, \gamma,\lambda,{\bm x})$}\label{alg:small-elements} \begin{algorithmic}[1] \Require{A monotone submodular function $g:2^E \to \mathbb{R}_+$, a monotone linear function $\ell:2^E \to \mathbb{R}$, a weight function $w :E \to [0,1]$, $\epsilon,\delta \in (0,1)$, a set of small elements $E_{\rm S}$, guessed values $\gamma,\lambda$, and a vector ${\bm x} \in [0,1]^E$.} \Ensure{A vector ${\bm v} \in [0,1]^E$.} \State{For each $e \in E_{\rm S}$, let ${\bm \theta}(e) = \textsc{Estimate}_{\epsilon,\epsilon/n,\delta/n}(g_{R({\bm x})}(e))$.} \State{Find a vector ${\bm v} \in [0,1]^E$ supported on $E_{\rm S}$ that minimizes $W({\bm v})$ subject to \[ {\bm v} \cdot {\bm \theta} \geq (1-\epsilon)\gamma - \epsilon d_{g,\ell} \quad \text{and} \quad L({\bm v}) \geq \lambda. \] by linear programming. } \State{\Return ${\bm v}$.} \end{algorithmic} \end{algorithm} Now, we analyze Algorithm~\ref{alg:small-elements}. From Corollary~\ref{cor:chernoff} and the fact that $g_{S}(e) \leq d_g \leq d_{g,\ell}$ for every $S \subseteq E$ and $e \in E$, we have the following: \begin{proposition}\label{pro:small-element-chernoff} With probability at least $1-\delta$, we have \[ (1- \epsilon)\mathop{\mathbf{E}}[g_{R({\bm x})}(e)] - \frac{\epsilon d_{g,\ell}}{n} \leq {\bm \theta}(e) \leq (1+ \epsilon)\mathop{\mathbf{E}}[g_{R({\bm x})}(e)] + \frac{\epsilon d_{g,\ell}}{n} \] for every $e \in E_{\rm S}$. \end{proposition} We formalize the concept that $\gamma$ and $\lambda$ are sufficiently accurate, and then show that Algorithm~\ref{alg:small-elements} outputs a desired vector with accurate $\gamma$ and $\lambda$. \begin{definition}\label{def:good-guess-small} We say that $\gamma$ and $\lambda$ are \emph{good guesses} if \[ \mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})] \geq \gamma \geq (1-\epsilon)\mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})] - \epsilon d_{g,\ell} \quad \text{and} \quad \ell(O_{\rm S}) \geq \lambda \geq (1-\epsilon)\ell(O_{\rm S}) - \epsilon d_{g,\ell} \] hold, respectively. \end{definition} Since $\mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})] \leq nd_{g,\ell}$ and $\ell(O_{\rm S}) \leq nd_{g,\ell}$ hold, we can find good guesses by trying all the values in the set $V_{\epsilon,n}(g,\ell)$. \begin{lemma}\label{lem:small-element-guess} Suppose that $\gamma$ and $\lambda$ are good guesses. Then, Algorithm~\ref{alg:small-elements} returns a vector ${\bm v} \in [0,1]^E$ supported on $E_{\rm S}$ such that \begin{itemize} \item[(i)] $\sum_{e \in E_{\rm S}}{\bm v}(e)\mathop{\mathbf{E}}[g_{R({\bm x})}(e)] \geq (1-\epsilon)^3\mathop{\mathbf{E}}[g_{ R({\bm x})}(O_{\rm S})] - 3\epsilon d_{g,\ell}$, \item[(ii)] $L({\bm v}) \geq (1-\epsilon)\ell(O_{\rm S})-\epsilon d_{g,\ell}$, and \item[(iii)] $W({\bm v}) \leq w(O_{\rm S})$, \end{itemize} with probability at least $1-\delta$. The time complexity of Algorithm~\ref{alg:small-elements} is $O(n^4+n^2\log (n/\delta )/\epsilon^2)$. \end{lemma} \begin{proof} With probability at least $1-\delta$, the consequence of Proposition~\ref{pro:small-element-chernoff} holds. In what follows, we assume that this occurs. The vector ${\bm 1}_{O_{\rm S}}$ satisfies \begin{align*} {\bm 1}_{O_{\rm S}}\cdot {\bm \theta} & = \sum_{e \in O_{\rm S}}{\bm \theta}(e) \geq \sum_{e \in O_{\rm S}}\Bigl((1-\epsilon)\mathop{\mathbf{E}}[g_{R({\bm x})}(e)] - \frac{\epsilon d_{g,\ell}}{n}\Bigr) \\ & \geq (1-\epsilon)\mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})] - \epsilon d_{g,\ell} \geq (1-\epsilon)\gamma - \epsilon d_{g,\ell}. \end{align*} Furthermore, we have $L({\bm 1}_{O_{\rm S}}) = \ell(O_{\rm S}) \geq \lambda$. Hence, the vector ${\bm v}$ is well defined, and in particular, we have $W({\bm v}) \leq W(O_{\rm S})$. Then, we have \begin{align*} \sum_{e \in E_{\rm S}}{\bm v}(e)\mathop{\mathbf{E}}[g_{R({\bm x})}(e)] & \geq (1-\epsilon)\sum_{e \in E_{\rm S}}{\bm v}(e) \Bigl({\bm \theta}(e)-\frac{\epsilon d_{g,\ell}}{n}\Bigr) \geq (1-\epsilon)\sum_{e \in E_{\rm S}}{\bm v}(e) {\bm \theta}(e)-\epsilon d_{g,\ell} \\ & \geq (1-\epsilon)\bigl((1-\epsilon)\gamma -\epsilon d_{g,\ell}\bigr)- \epsilon d_{g,\ell} \geq (1-\epsilon)^2\gamma- 2\epsilon d_{g,\ell} \\ & \geq (1-\epsilon)^2 \bigl((1-\epsilon)\mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})] - \epsilon d_{g,\ell}\bigr) - 2\epsilon d_{g,\ell} \\ & \geq (1-\epsilon)^3\mathop{\mathbf{E}}[g_{R({\bm x})}(O_{\rm S})] - 3\epsilon d_{g,\ell}. \end{align*} It is easy to confirm (ii) and (iii). The time complexity for computing ${\bm \theta}$ is $O(n^2\log (n/\delta )/\epsilon^2)$ from Corollary~\ref{cor:chernoff}, and the time complexity for solving the linear program is $O(n^4)$ by using the ellipsoid method. The total time complexity is bounded by $O(n^4+n^2\log (n/\delta )/\epsilon^2)$. \end{proof} \subsection{Small elements}\label{subsec:small} Our algorithm computes a vector ${\bm x} \in [0,1]^E$ with $W({\bm x}) \leq 1$ and then rounds it to a set. A natural rounding method is to simply output the random set $R({\bm x})$. Then, we can guarantee that the expected objective values $\mathop{\mathbf{E}}[g(R({\bm x}))]$ and $\mathop{\mathbf{E}}[ \ell(R({\bm x}))]$ are sufficiently large and the expected weight $\mathop{\mathbf{E}}[w(R({\bm x}))]$ is at most one. However, we cannot guarantee the concentration of $w(R({\bm x}))$ because some elements have large contributions to the weight. To resolve this issue, we say that elements in $E$ are \emph{small} if \[ g(e) \leq \epsilon^6 g(O) \quad \text{and} \quad \ell(e) \leq \epsilon^6 \ell(O). \] Then, we can freely remove some of small elements for decreasing the weight without decreasing the value significantly. Further, we can prove that the number of large elements is bounded by a polynomial in $\epsilon$ and $c_g$. An issue here is that we do not know $O$; hence, we cannot determine whether an element is small. To resolve this issue, we guess the values of $g(O)$ and $\ell(O)$. Without loss of generality, we can assume that $\epsilon/n = (1-\epsilon)^k$ for some integer $k$; otherwise, we slightly decrease the value of $\epsilon$. Then, we define a set $V_{\epsilon,n}(g,\ell) = \set{nd_{g,\ell},(1-\epsilon)nd,\ldots,\epsilon d_{g,\ell}, 0}$, and we use the values in $V_{\epsilon}(g,\ell)$ to guess $g(O)$ and $\ell(O)$. Since $g(O) \leq nd_{g,\ell}$ and $\ell(O) \leq nd_{g,\ell}$ hold, there exist some $v_g, v_\ell \in V_{\epsilon,n}(g,\ell)$ such that \begin{align} (1-\epsilon)v_g - \epsilon d_{g,\ell} \leq g(O) \leq v_g \quad \text{and} \quad (1-\epsilon)v_\ell - \epsilon d_{g,\ell} \leq \ell(O) \leq v_\ell. \label{eq:guess-g(O)-and-l(O)} \end{align} We say that an element $e \in E$ is \emph{small with respect to $(v_g,v_\ell)$} if \[ g(e) \leq \epsilon^6 v_g \quad \text{and} \quad \ell(e) \leq \epsilon^6 v_\ell. \] Otherwise, we say that an element $e \in E$ is \emph{large with respect to $(v_g,v_\ell)$}. Let $E_{\rm L}(v_g,v_\ell) \subseteq E$ and $E_{\rm S}(v_g,v_\ell) \subseteq E$ be the sets of large and small elements, respectively, with respect to $(v_g,v_\ell)$. Further, we define $O_{\rm L}(v_g,v_\ell) = E_{\rm L}(v_g,v_\ell) \cap O$ and $O_{\rm S}(v_g,v_\ell) = E_{\rm S}(v_g,v_\ell) \cap O$. We omit $v_g$ and $v_\ell$ from these notations if they are clear from the context. When $v_g$ and $v_\ell$ satisfy~\eqref{eq:guess-g(O)-and-l(O)}, we can upper bound the number of large elements in $O$: \begin{lemma}\label{lem:bound-on-large-elements} If $v_g$ and $v_\ell$ satisfy~\eqref{eq:guess-g(O)-and-l(O)}, then we have $|O_{\rm L}| = O\bigl(\frac{1}{(1-c_g)\epsilon^6}\bigr)$. \end{lemma} \begin{proof} Let $m_\ell$ be the number of elements $e \in O$ with $\ell(e) > \epsilon^6 v_\ell$. Then, we have $\epsilon^6 v_\ell m_\ell \leq \ell(O)$. Since $v_\ell \geq \ell(O)$, we have $m_\ell \leq 1/\epsilon^6$. Let $\set{o_1,\ldots,o_{m_g}}$ be the set of elements $e \in O$ with $g(e) > \epsilon^6 v_g$. Then, we have \[ g(O) \geq \sum_{i \in [m_g]}g_{\set{o_1,\ldots,o_{i-1}}}(o_i) \geq (1-c_g)\sum_{i \in [m_g]}g(o_i) \geq (1-c_g)\epsilon^6 v_g m_g. \] Since $v_g \geq g(O)$, we have $m_g \leq \frac{1}{(1-c_g)\epsilon^6}$. Then, we have $|O_{\rm L}| \leq m_g + m_\ell = O(\frac{1}{(1-c_g)\epsilon^6})$. \end{proof} In addition to the values of $g(O)$ and $\ell(O)$, the value of $|O_{\rm L}|$ is not also not known. However, we can easily guess it because there are only $O(\frac{1}{(1-c_g)\epsilon^6})$ choices from Lemma~\ref{lem:bound-on-large-elements}. We use the symbol $m$ to denote the guessed value of $|O_{\rm L}|$. For each choice of $v_g$, $v_\ell$, and $m$, we compute a (random) set that jointly maximizes $g$ and $\ell$, and the final output is the best one among them. Since $|V_{\epsilon,n}(g,\ell)| = O(\log_{1/(1-\epsilon)}(n/\epsilon)) = O(\log(n/\epsilon)/\epsilon)$, this guessing process makes the running time $O\Bigl(\bigl(\log(n/\epsilon)/\epsilon\bigr)^2 \cdot \frac{1}{(1-c_g)\epsilon^6}\Bigr) = O\Bigl(\frac{\log^2(n/\epsilon)}{(1-c_g)\epsilon^8}\Bigr)$ times larger. The details will be explained in Section~\ref{subsec:final}.
ae90ea0a8bacb9e427099c9cbd18306e787ccae3
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{{Introduction}} \label{sect:intro} Waves measured by a collection of nearby sensors, called an array of receivers, carry information about their source and the medium through which they travel. We consider a typical remote sensing regime with sources of small (point-like) support, and study the inverse problem of determining them from the array measurements. When the waves travel in a known and non-scattering (e.g. homogeneous) medium, the sources can be localized with reverse time migration \cite{Biondi,bleistein2001mathematics} also known as backprojection \cite{munson1983tomographic}. This estimates the source locations as the peaks of the image formed by superposing the array recordings delayed by travel times from the receivers to the imaging points. The accuracy of the estimates depends on the array aperture, the distance of the sources from the array, and the temporal support of the signals emitted by the sources. It may be improved under certain conditions by using $l_1$ optimization, which seeks to invert the linear mapping from supposedly sparse vectors of the discretized source amplitude on some mesh, to the array measurements. The fast growing literature of imaging with $l_1$ optimization in homogeneous media includes compressed sensing studies such as \cite{fannjiang2010compressed,fannjiang2010compressive}, synthetic radar imaging studies like \cite{anitori2013design,borcea2016synthetic}, array imaging studies like \cite{chai2013robust}, and the resolution study \cite{borcea2015resolution}. In this paper we assume that the waves travel in heterogeneous media with fluctuations of the wave speed caused by numerous inhomogeneities. The amplitude of the fluctuations is small, meaning that a single inhomogeneity is a weak scatterer. However, there are many inhomogeneities that interact with the waves on their way from the sources to the receivers, and their scattering effect accumulates. Because in applications it is impossible to know the inhomogeneities in detail, and these cannot be estimated from the array measurements as part of the inversion, the fluctuations of the wave speed are uncertain. We model this uncertainty with a random process, and thus study inversion in random media. In this stochastic framework, the actual heterogeneous medium in which the waves propagate is one realization of the random model. The data measured at the array are uncertain and the question is how to mitigate the uncertainty to get images that are robust with respect to arbitrary medium realizations, i.e., they are statistically stable. The mitigation of uncertainty in the wave propagation model becomes important when the sources are further than a few scattering mean free paths from the array. The scattering mean free path is the length scale on which the waves randomize \cite{van1999multiple}, meaning that their fluctuations from one medium realization to another are large in comparison with their coherent (statistical expectation) part. The random wave distortions registered at the array are very different from additive and uncorrelated noise assumed usually in inversion. They are more difficult to mitigate and lead to poor and unreliable source reconstructions by coherent methods like reverse time migration or standard $l_1$ optimization. The Coherent Interferometric (CINT) method \cite{borcea2006adaptive,borcea2011enhanced} is designed to deal efficiently with such random distortions, as long as there is some residual coherence in the array measurements. This holds when the sources are separated from the array by distances (ranges) that are large with respect to the scattering mean free path, but do not exceed a transport mean free path, which is the distance at which the waves forget their initial direction \cite{van1999multiple}. The transport mean free path defines the range limit of applicability of coherent inversion methods. Beyond it only incoherent methods based on transport or diffusion equations \cite{arridge2009optical} can be used. In this paper we assume a scattering regime where the CINT method is useful. It forms images by superposing cross-correlations of the array measurements, delayed by travel times between the receivers and the imaging points. As shown in \cite{borcea2006adaptive,borcea2007asymptotics,borcea2011enhanced}, the cross-correlations must be computed locally, in appropriate time windows, and over limited receiver offsets. This introduces a smoothing in the CINT image formation, which is essential for stabilizing statistically the images, at the expense of resolution. The larger the random distortions of the array measurements, the more smoothing is needed and the worse the resolution \cite{borcea2006adaptive,borcea2007asymptotics}. Thus, it is natural to ask if it is possible to improve the source localization by using the prior information that the sources have small support. We show that under generic conditions, the CINT imaging function is approximately a discrete convolution of the vector of source intensities discretized on the imaging mesh, with a blurring kernel. To reconstruct the sources we seek to undo the convolution using convex ($l_1$) optimization. We present an analysis of the method in a scattering regime where the random medium effects on the array measurements can be modeled by large wavefront distortions, as assumed in adaptive optics \cite{beckers1993adaptive}. We derive from first principles the CINT blurring kernel in this regime, and state the inversion problem as an $l_1$ optimization. We also quantify the quality of the reconstruction with error estimates that depend on the separation of the sources, or of clusters of sources, but are independent of the source placement on or off the imaging mesh, as long as the sources are sufficiently far apart. The analysis shows that we can expect almost exact reconstructions when the sources are further apart than the CINT resolution limits. This is similar to the super-resolution results in \cite{candes2014towards}, that show that one dimensional discrete convolutions can be undone by convex optimization, assuming that the minimum distance between the points in the support of the unknown vectors is $2/f_c$, where $f_c$ is the largest ``frequency'' in the Fourier transform of the convolution kernel. When the sources are clustered together, the $l_1$ reconstruction is not guaranteed to be close to the true vector of source intensities in the point-wise sense. However, we show that its support is in the vicinity of the clusters, and its entries are related to the average source intensities there. The paper is organized as follows: We begin in section \ref{sect:form} with the formulation of the inverse problem as an $l_1$ optimization. The analysis of the method is in sections \ref{sect:anal} - \ref{sect:res}, and we demonstrate its performance with numerical simulations in section \ref{sect:num}. We end with a summary in section \ref{sect:sum}. \section{Formulation of the inverse problem} \label{sect:form} \begin{figure}[t] \vspace{-0.15in} \centerline{\includegraphics[width = 0.7 \textwidth]{Figures/PARAX_SCH_v2.pdf}} \vspace{-0.45in} \caption{Schematic for the inversion setup with an array of receivers that is planar square of side $a$. The range direction is orthogonal to the array aperture. The sources are at ranges of order $L$ in an imaging region $\mathfrak{D}$ modeled as a rectangular prism with size $D_3$ in range and $D$ in cross-range. } \label{fig:setup} \end{figure} Consider the inversion setup illustrated in Figure \ref{fig:setup}, where $N_s$ sources located at $\vec{\bm{y}}_s$, for $s = 1, \ldots, N_s$, emit signals $f_s(t)$ that generate sound waves recorded at a remote array of receivers placed at $\vec{\bm{x}}_r$, for $r = 1, \ldots, N_r$. For simplicity we assume that the array aperture is planar and square, of side $a$. This allows us to introduce a system of coordinates centered at the array, with range direction orthogonal to the array, and cross-range plane parallel to the array. In this system of coordinates we have $ \vec{\bm{x}}_r = (\bm{x}_r,0)$, with cross-range vectors $\bm{x}_r = (x_{r,1},x_{r,2})$ satisfying $|x_{r,1}|,|x_{r,2}| \le a/2$. The sources are at $\vec{\bm{y}}_s = (\bm{y}_s,y_{s,3})$, with range coordinates $y_{s,3}$ of order $L$, satisfying $L \gg a$, and two dimensional cross-range vectors $\bm{y}_s$. In general, the signals $f_s(t)$ emitted by the sources may be pulses, chirps or even noise-like, with Fourier transforms \begin{equation} \hat f_s(\omega) = \int_{-\infty}^\infty dt \, f_s(t) e^{i \omega t} \label{eq:f2} \end{equation} supported in the frequency interval $[\omega_o-\pi B,\omega_o + \pi B]$, where $\omega_o$ is the central frequency and $B$ is the bandwidth. We denote the recorded waves by $p(\vec{\bm{x}}_r,t)$, and use the linearity of the wave equation to write \begin{equation} p(\vec{\bm{x}}_r,t) = \int_{\omega_o-\pi B}^{\omega_o + \pi B} \frac{d \omega}{2 \pi} \, \hat p(\vec{\bm{x}}_r, \omega) e^{-i \omega t}, \qquad \hat p(\vec{\bm{x}}_r,\omega) = \sum_{s=1}^{N_s} \hat f_s(\omega) \hat G(\vec{\bm{x}}_r,\vec{\bm{y}}_s,\omega) + \hat n(\vec{\bm{x}}_r,\omega), \label{eq:f1} \end{equation} for $r = 1, \ldots, N_r$. Here the Green's function $\hat G$ models the propagation of time harmonic waves in the medium, and $\hat n$ denotes additive and uncorrelated noise. The inverse problem is to determine the sources from these array measurements. \subsection{Imaging in homogeneous media} The Green's function in media with constant speed $c_o$ is \begin{equation} \hat G_o(\vec{\bm{x}}_r,\vec{\bm{y}}_s,\omega) = \frac{\exp[i \omega \tau(\vec{\bm{x}}_r,\vec{\bm{y}}_s)]}{4 \pi |\vec{\bm{x}}_r-\vec{\bm{y}}_s|}, \label{eq:Go} \end{equation} where \begin{equation} \tau(\vec{\bm{x}}_r,\vec{\bm{y}}_s) = {|\vec{\bm{x}}_r-\vec{\bm{y}}_s|}/{c_o} \label{eq:tau} \end{equation} is the travel time from the source at $\vec{\bm{y}}_s$ to the receiver at $\vec{\bm{x}}_r$. The measurements are of the form \begin{equation} p_o(\vec{\bm{x}}_r,t) = \sum_{s=1}^{N_s} \frac{f_s\left(t - \tau(\vec{\bm{x}}_r,\vec{\bm{y}}_s)\right)}{4 \pi |\vec{\bm{x}}_r-\vec{\bm{y}}_s|} + n(\vec{\bm{x}}_r,t), \label{eq:f4} \end{equation} and in reverse time migration they are synchronized using travel time delays with respect to a presumed source location at a search point $\vec{\bm{y}}$, and then superposed to form an image \begin{align} \mathcal{J}_o(\vec{\bm{y}}) = \sum_{r=1}^{N_r} p_o\left(\vec{\bm{x}}_r,\tau(\vec{\bm{x}}_r,\vec{\bm{y}})\right). \label{eq:f5} \end{align} The resolution limits of the imaging function \eqref{eq:f5} are well known. The cross-range resolution is of order $\lambda_o L/a$, where $\lambda_o = 2 \pi c_o/\omega_o$ is the central wavelength, and the range resolution is inverse proportional to the temporal support of $f_s(t)$, which determines the precision of the travel time estimation. If the signals $f_s(t)$ are pulses, their temporal support is of order $1/B$, and the range resolution is of order $c_o/B$. If they are chirps or other long signals that are known, then they must be compressed in time by cross-correlation (matched filtering) with the time reversed $f_s(t)$ to achieve the $c_o/B$ range resolution \cite{Curlander}. If the signals are unknown and noise-like, then imaging must be based on cross-correlations of the array measurements, like in CINT. We refer to \cite{borcea2015resolution} for the formulation of the inverse source problem as an $l_1$ optimization, and recall from there that the resolution limits $\lambda_o L/a$ and $c_o/B$ also play a role in the successful recovery of the presumed sparse source support. \subsection{Coherent interferometric imaging} The imaging function $\mathcal{J}_o(\vec{\bm{y}})$ does not work well in random media at ranges $L$ that exceed a few scattering mean free paths. This is because the measurements have large random distortions that are very different than the additive noise $n(\vec{\bm{x}}_r,t)$, and cannot be reduced by simply summing over the receivers as in \eqref{eq:f5}. To mitigate these distortions we image with the CINT function \begin{align} \mathcal{J}(\vec{\bm{y}}) = \int_{-\infty}^\infty d \omega \int_{-\infty}^\infty d \widetilde \omega \, \hat \phi\left(\frac{\widetilde \omega}{\Omega}\right) \sum_{r=1}^{N_r} \sum_{r' = 1}^{N_r} \psi\left(\frac{|\bm{x}_r-\bm{x}_{r'}|}{X(\omega)}\right) \hat p (\vec{\bm{x}}_r,\omega + \widetilde \omega/2) \overline{\hat p(\vec{\bm{x}}_{r'},\omega-\widetilde \omega/2)} \nonumber \\ \times \exp[-i (\omega + \widetilde \omega/2) \tau(\vec{\bm{x}}_r,\vec{\bm{y}}) + i (\omega - \widetilde \omega/2) \tau(\vec{\bm{x}}_{r'},\vec{\bm{y}})], \label{eq:f6} \end{align} where $\hat \phi$ and $\psi$ are smooth window functions of dimensionless argument and support of order one, and the domain of integration is restricted by the finite bandwidth that supports the measurements, \[ \omega \pm \widetilde \omega/2 \in [\omega_o-\pi B,\omega_o + \pi B]. \] As in reverse time migration, the travel times are used in \eqref{eq:f6} to synchronize the waves due to a presumed source at the search location $\vec{\bm{y}}$. However, the image is formed by superposing cross-correlations of the array measurements $p(\vec{\bm{x}}_r,t)$, instead of the measurements themselves. The cross-correlations are convolutions of $p(\vec{\bm{x}}_r,t)$ with the time reversed $p(\vec{\bm{x}}_{r'},t)$, for $r, r' = 1, \ldots, N_r$. The time reversal appears as complex conjugation in the frequency domain, denoted with the bar in \eqref{eq:f6}. The time window \begin{equation} \label{eq:defphi} \phi\left(\Omega t\right) = \frac{1}{2 \pi \Omega}\int_{-\infty}^\infty d \omega\,\hat \phi\left(\frac{\omega}{\Omega}\right)e^{-i \omega t} \end{equation} and spatial window $\psi(|\bm{x}|/X)$ ensure that the cross-correlations are computed locally, over receiver offsets that do not exceed the distance $X$, and over time offsets of order $1/\Omega$. These threshold parameters account for the fact that scattering in random media decorrelates statistically the waves at frequencies separated by more than $\Omega_d$, the decoherence frequency, and points separated by more than $X_d$, the decoherence length. We refer to \cite{borcea2007asymptotics,borcea2011enhanced} and the next section for more details. Here it suffices to recall that \eqref{eq:f6} is robust\footnote{Robustness refers to statistical stability of the image with respect to the realizations of the random medium. It means that the standard deviation of $\eqref{eq:f6}$ is small with respect to its expectation near the peaks.} when $X \lesssim X_d$ and $\Omega \lesssim \Omega_d$, and that the image $\mathcal{J}(\vec{\bm{y}})$ has a cross-range resolution of order $\lambda_o L /X$ and a range resolution of order $c_o/\Omega$. The best focus occurs at $X \approx X_d$ and $\Omega \approx \Omega_d$, so the decoherence parameters $X_d$ and $\Omega_d$ can be estimated with optimization, as explained in \cite{borcea2006adaptive}. To state the inverse problem as a convex optimization, we make the simplifying assumption that the sources emit the same known pulse $f(t)$, so that \begin{equation} \hat f_s(\omega) = \hat f(\omega) \rho(\vec{\bm{y}}_s), \qquad s = 1, \ldots, N_s, \label{eq:f7} \end{equation} for an unknown, complex valued amplitude $\rho(\vec{\bm{y}})$. Using \eqref{eq:f7} in \eqref{eq:f1} and substituting in \eqref{eq:f6} we obtain \begin{equation} \mathcal{J}(\vec{\bm{y}}) \approx \sum_{s=1}^{N_s} \sum_{s'=1}^{N_s} \rho(\vec{\bm{y}}_s) \overline{\rho(\vec{\bm{y}}_{s'})}\, \kappa(\vec{\bm{y}}, \vec{\bm{y}}_s,\vec{\bm{y}}_{s'}), \label{eq:f8} \end{equation} with kernel \begin{align} \kappa(\vec{\bm{y}},\vec{\bm{y}}_s,\vec{\bm{y}}_{s'}) =& \sum_{r=1}^{N_r} \sum_{r' = 1}^{N_r} \psi\left(\frac{|\bm{x}_r-\bm{x}_{r'}|}{X}\right) \int_{-\infty}^\infty d \omega \int_{-\infty}^\infty d \widetilde \omega \, \hat \phi\left(\frac{\widetilde \omega}{\Omega}\right) \hat f\left(\omega + {\widetilde \omega}/{2}\right) \overline{\hat f\left(\omega-{\widetilde \omega}/{2}\right)} \nonumber \\ & \hspace{-0.5in}\times \hat G\left(\vec{\bm{x}}_r,\vec{\bm{y}}_s,\omega+{\widetilde \omega}/{2}\right)\overline{\hat G(\vec{\bm{x}}_{r'},\vec{\bm{y}}_{s'},\omega-\widetilde \omega/2)}\exp[-i (\omega + \widetilde \omega/2) \tau(\vec{\bm{x}}_r,\vec{\bm{y}})] + i (\omega - \widetilde \omega/2)\tau(\vec{\bm{x}}_{r'},\vec{\bm{y}})], \label{eq:f9} \end{align} where the approximation is because we neglect the additive noise\footnote{Additive noise is considered in all the numerical simulations in section \ref{sect:num}, but for simplicity we neglect it in the analysis.}. In the analysis of the next section we take the Gaussian pulse \begin{equation} \hat f(\omega) = \left(\frac{\sqrt{2 \pi}}{B}\right)^{1/2} \exp \left[ -\frac{(\omega-\omega_o)^2}{4 B^2}\right], \label{eq:f10} \end{equation} normalized by \[ \|f \|_{_2} = \left[\int_{-\infty}^\infty dt \, |f(t)|^2 \right]^{1/2}= \left[\int_{-\infty}^\infty \frac{d \omega}{2 \pi}\, |\hat f(\omega)|^2\right]^{1/2} = 1. \] This choice allows us to obtain an explicit expression of the kernel \eqref{eq:f9}. Naturally, in practice the pulses may not be Gaussian, and the sources may emit different signals. The method described here still applies to such cases, with $\rho(\vec{\bm{y}}_s)$ replaced in \eqref{eq:f8} by $\|f_s\|_{_2}$, and the substitution \[ \hat f(\omega + \widetilde \omega/2) \overline{\hat f(\omega-\widetilde \omega/2)} \leadsto \frac{\hat f_s(\omega + \widetilde \omega/2)}{\|f_s\|} \frac{\overline{\hat f_{s'}(\omega-\widetilde \omega/2)}}{\|f_{s'}\|_{_2}} \] in \eqref{eq:f9}. Since $f_s$ is unknown in general, we may only estimate the kernel $\kappa$ up to unknown, constant multiplicative factors. This still allows the determination of the location of the sources, but does not give good estimates of their intensities. \subsection{The optimization problem} \label{sect:OPT} Let us consider a reconstruction mesh with $N_z$ points denoted generically by $\vec{\bm{z}}$, and name $\bm{\rho}$ the column vector with entries given by the unknown $\rho(\vec{\bm{z}})$. We sample the CINT image $\mathcal{J}(\vec{\bm{y}})$ at $N_y < N_z$ points, and gather these samples in the ``data'' vector $\bm{d}$. It is natural to take $N_y < N_z$, because we seek to super-resolve the CINT image, which is blurred by the kernel \eqref{eq:CINTK}. At first glance it appears that we may use equation \eqref{eq:f8} to formulate the inversion as an optimization problem for recovering the rank one matrix $\bm{\rho} \bm{\rho}^\star$, as in \cite{candes2013phaselift,chai2010array}. However, as shown in the next section, in strong random media where CINT is needed, the kernel $\kappa(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')$ is large only when $\vec{\bm{z}}$ and $\vec{\bm{z}}'$ are nearby. In fact, for reasonable mesh sizes on which we can expect to obtain unique reconstructions, the kernel satisfies \begin{equation} \kappa(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}') \approx \left\{ \begin{array}{ll} \kappa(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}) \quad &\mbox{if} ~ \vec{\bm{z}}'=\vec{\bm{z}}, \\ 0 & \mbox{if}~ \vec{\bm{z}} \ne \vec{\bm{z}}'. \end{array}\right. \label{eq:diagKappa} \end{equation} Thus, only the diagonal entries $|\rho(\vec{\bm{z}})|^2$ of $\bm{\rho} \bm{\rho}^\star$ play a role. These are the source intensities and we denote by $\bm{u} \in \mathbb{R}^{N_z}$ the vector of unknowns formed by them. Equation \eqref{eq:f8} gives \begin{equation} \bm{\mathcal{M}}\, \bm{u} \approx \bm{d}, \label{eq:f11} \end{equation} where $\bm{\mathcal{M}}$ is the $N_y \times N_z$ ``measurement'' matrix with entries $\kappa(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}).$ We formulate the inversion as the $l_1$ optimization problem \begin{equation} \min_{\bm{u} \in \mathbb{R}^{N_z}} \|\bm{u}\|_{_1} ~ ~ \mbox{such that} ~~ \|\bm{\mathcal{M}}\, \bm{u} - \bm{d}\|_{_2} \le ~\mbox{tolerance}, \label{eq:f12} \end{equation} where the tolerance accounts for additive noise effects and the random fluctuations of the CINT imaging function, which are small for large enough array aperture $a$ and bandwidth $B$, as shown in \cite{borcea2007asymptotics,borcea2011enhanced}. We prove in section \ref{sect:CINT} that the left hand side in \eqref{eq:f11} is approximately a discrete convolution. The $l_1$ optimization is useful in this context, and recovers well sources that are well separated, as expected from the results in \cite{candes2014towards}. We rediscover such results in this paper using a different analysis. We also consider cases where the sources are clustered together, and show that although we cannot expect good reconstructions in the point-wise sense, the $l_1$ minimizer is supported in the vicinity of the clusters, and estimates the average source intensity there. \section{Setup of the analysis} \label{sect:anal} We introduce in this section the random wave speed model and the scaling assumptions which define the relations between the wavelength $\lambda_o$, the typical size $\ell$ of the inhomogeneities in the medium, the standard deviation $\sigma$ of the fluctuations of the wave speed, the array aperture $a$, the range $L$, and the extent of the imaging region. The scaling allows us to describe the scattering effects of the random medium as large wavefront distortions. This is a simple wave propagation model that is convenient for analysis, and captures qualitatively all the important features of imaging with CINT. That is to say, equation \eqref{eq:f11} holds in general scattering regimes, and the CINT kernel $\kappa$ has a similar form, but the mathematical expression of the decoherence length $X_d$ and frequency $\Omega_d$, which quantify the blurring by the kernel, are expected to change. The expressions of $X_d$ and $\Omega_d$ in terms of $\lambda_o, \sigma, \ell$ and $L$ are needed for analysis, but they are unlikely to be known in practice. This is why one should determine the decoherence parameters directly form the data or adaptively, during the CINT image formation, as in \cite{borcea2006adaptive}. The model of the wave speed $c(\vec{\bm{x}})$ is \begin{equation} c(\vec{\bm{x}}) = c_o \left[ 1 + \sigma \mu \left(\frac{\vec{\bm{x}}}{\ell}\right) \right]^{-1/2}, \label{eq:a1} \end{equation} where $\mu$ is a mean zero, stationary random process of dimensionless argument. We suppose that $\mu$ is bounded almost surely and denote by $\mathcal{R}$ its auto-correlation, assumed isotropic and Gaussian for convenience \begin{equation} \mathcal{R}(\vec{\bm{x}}) = \mathbb{E} \left[ \mu(\vec{\bm{x}} + \vec{\bm{x}}') \mu(\vec{\bm{x}})\right] = e^{- |\vec{\bm{x}}|^2/2}. \label{eq:a2} \end{equation} Then, $\sigma \ll 1$ quantifies the small amplitude of the fluctuations of $c(\vec{\bm{x}})$, and $\ell$ is the correlation length, which characterizes the typical size of the inhomogeneities in the medium. We explain in section \ref{sect:sc1} how far the waves should propagate in media modeled by \eqref{eq:a1}, so that the cumulative scattering effects can be described by large wavefront distortions. We are interested in imaging with finite size arrays at long distances, so the waves propagate in the range direction, within a cone (beam) of small opening angle. This is the paraxial regime defined in section \ref{sect:sc2}. We describe in section \ref{sect:sc3} the wave randomization quantified by the scattering mean free path, and the statistical decorrelation quantified by the decoherence length and frequency. We end the section with a summary of the scaling assumptions. \subsection{Wave scattering regime} \label{sect:sc1} We use a geometrical optics (Rytov) wave propagation model that holds in high frequency regimes with separation of scales \begin{equation} \label{eq:a3} \lambda_o \ll \ell \ll L, \end{equation} and standard deviation $\sigma$ of the fluctuations satisfying \begin{equation} \sigma \ll \min \left\{ \left(\frac{\ell}{L}\right)^{3/2}, \frac{\sqrt{\ell \lambda_o}}{L}\right\}. \label{eq:a4} \end{equation} It is shown in \cite{rytov1989principle} that the first bound in \eqref{eq:a4} ensures that the rays remain straight and the variance of the amplitude of the Green's function is negligible, so we can use the same geometrical spreading factor as in the homogeneous medium. The second bound in \eqref{eq:a4} ensures that only first order (in $\sigma$) corrections of the travel time matter, so the propagation model is \begin{equation} \hat G(\vec{\bm{x}},\vec{\bm{y}},\omega) \approx \hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega) \exp \left[ i \omega \tau(\vec{\bm{x}},\vec{\bm{y}})\frac{\sigma}{2} \int_0^1 d \vartheta \, \mu \left(\frac{(1-\vartheta) \vec{\bm{y}} + \vartheta \vec{\bm{x}}}{\ell}\right) \right]. \label{eq:a6} \end{equation} Let us write the random phase correction in \eqref{eq:a6} as \begin{equation} \omega \delta \tau(\vec{\bm{x}},\vec{\bm{y}}) = \frac{(2 \pi)^{1/4}}{2} \sigma k \sqrt{\ell |\vec{\bm{x}}-\vec{\bm{y}}|} \, \nu(\vec{\bm{x}},\vec{\bm{y}}), \label{eq:a7p} \end{equation} where $k = \omega/c_o$ is the wavenumber and \begin{equation} \nu(\vec{\bm{x}},\vec{\bm{y}}) = \frac{1}{(2 \pi)^{1/4}} \sqrt{\frac{|\vec{\bm{x}}-\vec{\bm{y}}|}{\ell}} \int_0^1 d \vartheta \, \mu \left(\frac{(1-\vartheta) \vec{\bm{y}} + \vartheta \vec{\bm{x}}}{\ell}\right) \label{eq:a7} \end{equation} is defined by the integral of the fluctuations along the straight ray between $\vec{\bm{y}}$ and $\vec{\bm{x}}$. It is shown in \cite[Lemma 3.1]{borcea2011enhanced} that $\nu(\vec{\bm{x}},\vec{\bm{y}})$ converges in distribution as $\ell/|\vec{\bm{x}}-\vec{\bm{y}}| \sim \ell/L \to 0$ to a Gaussian process. Its mean \begin{equation} \mathbb{E} \left[ \nu(\vec{\bm{x}},\vec{\bm{y}})\right] = 0, \end{equation} and variance \begin{equation} \mathbb{E} \left[ \nu^2(\vec{\bm{x}},\vec{\bm{y}}) \right] = \frac{|\vec{\bm{x}}-\vec{\bm{y}}|}{\ell \sqrt{2 \pi}}\int_0^1 d \vartheta \int_0^1 d \vartheta' \, \mathcal{R} \left( \frac{(\vartheta'-\vartheta) (\vec{\bm{x}}-\vec{\bm{y}})}{\ell}\right) \approx 1, \label{eq:a8} \end{equation} are calculated from definition \eqref{eq:a7} and the expression \eqref{eq:a2} of $\mathcal{R}$. The approximation is for $\ell/L$ small. We conclude from \eqref{eq:a7p}, \eqref{eq:a8} and $k = O( k_o)$, with $k_o = 2 \pi /\lambda_o$, that the random phase fluctuations in \eqref{eq:a6} have standard deviation of order $\sigma \sqrt{\ell L}/\lambda_o$. When this is small, the random medium effects are negligible and any coherent imaging method works well. We are interested in the case of large fluctuations, so we ask that \begin{equation} \label{eq:a9} \sigma \gg \frac{\lambda_o}{\sqrt{\ell L}}. \end{equation} This is consistent with \eqref{eq:a4} when \[ \frac{\lambda_o/\sqrt{\ell L}}{(\ell/L)^{3/2}} = \frac{\lambda_o L}{\ell^2} \ll 1, \qquad \frac{\lambda_o/\sqrt{\ell L}}{\sqrt{\ell \lambda_o}/{L} } = \frac{\sqrt{\lambda_o L}}{\ell} \ll 1, \] so we tighten our assumption \eqref{eq:a3} on the correlation length as \begin{equation} \sqrt{\lambda_o L} \ll \ell \ll L. \label{eq:a10} \end{equation} \subsection{The paraxial regime} \label{sect:sc2} Suppose that the sources are contained in the search (imaging) region \begin{equation} \label{eq:imreg} \mathfrak{D} = \left[-D/2,D/2\right]\times\left[-D/2,D/2\right]\left[-D_3/2,D_3/2\right], \end{equation} which is a rectangular prism of sides $D$ in cross-range and $D_3$ in range, as illustrated in Figure \ref{fig:setup}. When $D$ and the array aperture $a$ are small with respect to the range scale $L$, the rays connecting the sources and the receivers are contained within a cone (beam) of small opening angle, and we can use the paraxial approximation to simplify the calculations. The paraxial regime is defined by the scaling relations \begin{equation} \lambda_o \ll D \lesssim a \ll L, \qquad D_3 \ll L, \qquad \frac{a^4}{\lambda_o L^3} \ll 1, \qquad \frac{a^2 D_3}{\lambda_o L^2} \ll 1, \label{eq:a11} \end{equation} so that for $\vec{\bm{y}} = (\bm{y},y_3)$ and $\vec{\bm{x}} = (\bm{x},0)$ we get \begin{align} k |\vec{\bm{x}}-\vec{\bm{y}}| &= k \left( y_3 + \frac{|\bm{x}|^2}{2 L}- \frac{\bm{x} \cdot \bm{y}}{L} + \frac{|\bm{y}|^2}{2 L} \right) + O \left( \frac{a^4}{\lambda_o L^3}\right) + O\left(\frac{a^2 D_3}{\lambda_o L^2}\right) \nonumber \\ &\approx k \left( y_3 + \frac{|\bm{x}|^2}{2 L}- \frac{\bm{x} \cdot \bm{y}}{L} + \frac{|\bm{y}|^2}{2 L} \right), \label{eq:a12} \end{align} and \begin{equation} \frac{1}{4 \pi |\vec{\bm{x}}-\vec{\bm{y}}|} = \frac{1}{4 \pi L} \left[ 1 + O\left(\frac{D_3}{L}\right) + O\left(\frac{a^2}{L^2} \right) \right]\approx \frac{1}{4 \pi L}. \label{eq:a13} \end{equation} These approximations are proved in appendix \ref{ap:DerPar}, and the deterministic factor in \eqref{eq:a6} becomes \begin{equation} \hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega) \approx \frac{1}{4 \pi L} \exp \left[ i k \left(y_3 + \frac{|\bm{x}|^2}{2 L}- \frac{\bm{x} \cdot \bm{y}}{L} + \frac{|\bm{y}|^2}{2 L} \right) \right]. \label{eq:a14} \end{equation} \subsection{Randomization and statistical decorrelation of the waves} \label{sect:sc3} Here we quantify the scattering mean free path $\mathcal{S}$, the length scale on which the waves randomize (lose coherence), and the decoherence length $X_d$ and frequency $\Omega_d$, which describe the statistical decorrelation of the waves due to scattering. These important scales appear in the definition of the CINT blurring kernel defined in section \ref{sect:CINT}. \vspace{0.05in} \begin{proposition} \label{prop.1} The expectation of the Green's function \eqref{eq:a6} is given by \begin{equation} \mathbb{E}\big[ \hat G(\vec{\bm{x}},\vec{\bm{y}},\omega)\big] \approx \hat G_o (\vec{\bm{x}},\vec{\bm{y}},\omega) e^{- \frac{|\vec{\bm{x}}-\vec{\bm{y}}|}{\mathcal{S}(\omega)}} \approx 0, \label{eq:prop1.1} \end{equation} where $\mathcal{S}(\omega)$ is the scattering mean free path defined by \begin{equation} \mathcal{S}(\omega) = \frac{8}{\sqrt{2 \pi} \sigma^2 k^2 \ell} \ll L. \label{eq:prop1.2} \end{equation} \end{proposition} \noindent This result, proved in appendix \ref{ap:moments}, shows that the wavefront distortions due to scattering in the random medium do not average out. The expectation of $\hat{G}$ is not the same as the Green's function in the homogeneous medium, but decays exponentially with the distance of propagation on the scale $\mathcal{S}(\omega)$, the scattering mean free path. The scaling assumption \eqref{eq:a9} and definition \eqref{eq:prop1.2} give \[ \frac{|\vec{\bm{x}}-\vec{\bm{y}}|}{\mathcal{S}(\omega)} = O \left(\frac{L}{\mathcal{S}(\omega)}\right) = O(\sigma^2 k^2 \ell L) \gg 1, \] which is why the expectation in \eqref{eq:prop1.1} is almost zero. The standard deviation of the fluctuations is approximately \[ \mbox{std}[\hat G(\vec{\bm{x}},\vec{\bm{y}},\omega)] \approx \sqrt{\Big|\hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega)\Big|^2 - \Big|\mathbb{E}\big[ \hat G(\vec{\bm{x}},\vec{\bm{y}},\omega)\big] \Big|^2} \approx \Big| \hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega)\Big|, \] where we used that $|\hat G(\vec{\bm{x}},\vec{\bm{y}},\omega)| \approx |\hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega)|$. Thus, the random fluctuations of the waves dominate their coherent part (the expectation) at the ranges considered in our analysis, \[ \frac{ \Big| \mathbb{E} \big[ \hat G(\vec{\bm{x}},\vec{\bm{y}},\omega) \big] \Big|}{\mbox{std}[\hat G(\vec{\bm{x}},\vec{\bm{y}},\omega)]} \approx e^{-\frac{|\vec{\bm{x}}-\vec{\bm{y}}|}{S(\omega)}} \approx 0, \] and the wave is randomized. Reverse time migration or standard $l_1$ optimization methods cannot mitigate these large random distortions, as we illustrate with numerical simulations. This is why we base our inversion on the CINT method. \vspace{0.05in} \begin{proposition} \label{prop.2} Consider two points $\vec{\bm{x}} = (\bm{x},0)$ and $\vec{\bm{x}}' = (\bm{x}',0)$ in the array aperture and two points $\vec{\bm{y}} = (\bm{y},y_3)$ and $\vec{\bm{y}}' = (\bm{y}',y_3')$ in the imaging region. Assume that the bandwidth $B$ is small with respect to the central frequency $\omega_o$, so that $|\omega-\omega_o|, |\omega'-\omega_o| \le \pi B \ll \omega_o$. Then, the second moments of \eqref{eq:a6} are \begin{align} \mathbb{E}\left[ \hat G(\vec{\bm{x}},\vec{\bm{y}},\omega) \overline{\hat G(\vec{\bm{x}}',\vec{\bm{y}}',\omega')}\right] \approx \hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega) \overline{\hat G_o(\vec{\bm{x}}',\vec{\bm{y}}',\omega')} e^{- \frac{|y_3-y_3'|}{\mathcal{S}}- \frac{(\omega-\omega')^2}{2 \Omega_d^2} - \frac{|\bm{y}-\bm{y}'|^2 + (\bm{y}-\bm{y}')\cdot(\bm{x}-\bm{x}') + |\bm{x}-\bm{x}'|^2}{2 X_d^2}}, \label{eq:prop2.1} \end{align} with short notation $\mathcal{S} = \mathcal{S}(\omega_o)$, and decoherence frequency $\Omega_d$ and decoherence length $X_d$ defined by \begin{equation} \Omega_d = \frac{2 \omega_o}{(2 \pi)^{5/4}} \left(\frac{\lambda_o}{\sigma \sqrt{\ell L}}\right) \ll \omega_o, \qquad X_d =\sqrt{3} \ell\, \frac{ \Omega_d}{\omega_o} \ll \ell. \label{eq:prop2.2} \end{equation} \end{proposition} \noindent Moment formula \eqref{eq:prop2.1} is proved in appendix \ref{ap:moments}, and the inequalities in \eqref{eq:prop2.2} are due to assumption \eqref{eq:a9}. The first exponential factor in \eqref{eq:prop2.1} accounts for the randomization due to the travel time fluctuations between the two ranges. In our scaling $|y_3-y_3'| \lesssim D_3$, and by the last inequality in \eqref{eq:a11}, and the definition of the scattering mean free path, we have \begin{equation} \frac{|y_3-y_3'|}{\mathcal{S}(\omega_o)} = O \left(D_3 \frac{\sigma^2\ell}{\lambda_o^2} \right) \ll O\left(\frac{\lambda_o L^2}{a^2} \frac{\sigma^2 \ell}{\lambda_o^2} \right) \ll O\left(\frac{\ell^2}{a^2} \right), \label{eq:a15} \end{equation} where we used the bound \eqref{eq:a4} on $\sigma$. It is shown in \cite[Section 4]{borcea2011enhanced} that the standard deviation of the CINT image is small with respect to the expectation of its peak value (i.e., the imaging function is statistically stable) when \begin{equation} a > \ell. \label{eq:a16} \end{equation} Stability is essential for imaging to succeed, so we ask that the array aperture satisfy \eqref{eq:a16}, and conclude from \eqref{eq:a15} that the second moments \eqref{eq:prop1.1} simplify as\footnote{Although this moment formula is derived here using the model \eqref{eq:a6}, the result is generic and can be obtained in other scattering regimes. The only difference is the definition of $X_d$ and $\Omega_d$. See for example \cite{papanicolaou2007self}.} \begin{align} \mathbb{E}\left[ \hat G(\vec{\bm{x}},\vec{\bm{y}},\omega) \overline{\hat G(\vec{\bm{x}}',\vec{\bm{y}}',\omega')}\right] &\approx \hat G_o(\vec{\bm{x}},\vec{\bm{y}},\omega) \overline{\hat G_o(\vec{\bm{x}}',\vec{\bm{y}}',\omega')}e^{ - \frac{(\omega-\omega')^2}{2 \Omega_d^2} - \frac{|\bm{y}-\bm{y}'|^2 + (\bm{y}-\bm{y}')\cdot(\bm{x}-\bm{x}') + |\bm{x}-\bm{x}'|^2}{2 X_d^2}}, \label{eq:a17} \end{align} with $\hat G_o$ given in \eqref{eq:a14}. The exponential decay in \eqref{eq:a17} models the statistical decorrelation of the waves due to scattering. In our context, the spatial decorrelation, modeled by the decay in $\bm{x}-\bm{x}'$ and $\bm{y}-\bm{y}'$, can be explained by the fact that rays connecting sources to far apart receivers traverse through different parts of the random medium. Because $\mu$ does not have long range correlations, the fluctuations of the travel time along such different rays are statistically uncorrelated. The waves at far apart frequencies are uncorrelated because they interact differently with the random medium. This gives the decay in $\omega-\omega'$ in equation \eqref{eq:a17}. Definition \eqref{eq:f9} of the CINT kernel involves the superposition of $\hat G(\vec{\bm{x}},\vec{\bm{y}},\omega) \overline{\hat G(\vec{\bm{x}}',\vec{\bm{y}}',\omega')}$ over the array elements and frequencies. If the array aperture $a$ is large with respect to $X_d$, the superposition stabilizes statistically because we sum many uncorrelated entries, as in the law of large numbers. This is why CINT is robust with respect to the uncertainty of the fluctuations of the wave speed, as shown in \cite{borcea2007asymptotics,borcea2011enhanced}. \subsection{Summary of the scaling assumptions} \label{sect:sc4} We gather here the scaling assumptions stated throughout the section, and complement them with extra assumptions on the bandwidth and the size of the imaging region. We refer to appendix \ref{ap:verify} for the verification of their consistency. The wavelength $\lambda_o$ is the smallest length scale, and the range $L$ is the largest. The assumptions \eqref{eq:a16} and \eqref{eq:a11} on the aperture are \begin{equation} \ell < a \ll (\lambda_o L^3)^{1/4}. \label{eq:as1} \end{equation} The upper bound ensures that we can use the paraxial approximation and the lower bound gives $a > \ell \gg X_d$, so that the CINT image is statistically stable. Assumption \eqref{eq:a10} combined with \eqref{eq:as1} gives that the correlation length of the wave speed fluctuations should satisfy \begin{equation} \sqrt{\lambda_o L} \ll \ell \ll (\lambda_o L^3)^{1/4} \ll L. \label{eq:as2} \end{equation} The standard deviation $\sigma$ of the fluctuations is bounded above as in \eqref{eq:a4}, and below as in \eqref{eq:a9}, \begin{equation} \frac{\lambda_o}{\sqrt{\ell L}} \ll \sigma \ll \frac{\sqrt{\ell \lambda_o}}{L}. \label{eq:as3} \end{equation} The cross-range and range sizes $D$ and $D_3$ of the imaging region should be large with respect to the CINT resolution limits of $\lambda_o L/X$ in cross-range and $c_o/\Omega$ in range (see next section), so we can observe the image focus. We take the threshold parameters \begin{equation} X/X_d = O(1), \qquad \Omega/\min\{\Omega_d,B\} = O(1), \label{eq:as4pp} \end{equation} and recalling the scaling assumptions \eqref{eq:a11} that allow us to use the paraxial approximation, we obtain \begin{equation} \frac{c_o}{\Omega} \ll D_3 \ll \frac{\lambda_o L^2}{a^2} \quad \mbox{and} \quad \frac{\lambda_o L}{X} \ll D \lesssim a. \label{eq:as5} \end{equation} In general, the CINT image is statistically stable if in addition to having $a \gg X_d$, which follows from \eqref{eq:prop2.2} and \eqref{eq:as1}, the bandwidth $B$ is larger than the decoherence frequency $\Omega_d$. However, for the propagation model \eqref{eq:a6} considered in this section, where the effect of the random medium consists only of wavefront distortions and no delay spread (reverberations), the bandwidth $B$ does not play a role in the stabilization of CINT, as shown in \cite[Section 4.4.4]{borcea2011enhanced}. Thus, we study imaging in both narrowband and broadband regimes: \vspace{0.05in} \noindent \textbf{The narrowband regime} is defined by $B$ satisfying \begin{equation} \omega_o\left(\frac{a}{L}\right)^2 \ll B \ll \omega_o \min\left\{1, \frac{\lambda_o L}{a X_d} \right\}= \omega_o \frac{\lambda_o L}{a X_d}. \label{eq:as4} \end{equation} As verified in Appendix \ref{ap:verify}, \begin{equation} \frac{\lambda_o L}{X_d} \ll \ell < a, \qquad \frac{\lambda_o L}{a X_d} \ll \frac{\Omega_d}{\omega_o}, \label{eq:as4p} \end{equation} so $B \ll \Omega_d$. This choice leads to a simpler expression of the CINT blurring kernel, but since $\Omega$ is of the order of $B$, the range resolution is the same as in the homogeneous medium, and cannot be improved with optimization unless the sources are very far apart in range. However, the optimization can improve the cross-range focusing. \vspace{0.05in}\noindent \textbf{The broadband regime} is defined by \begin{equation} \Omega_d \ll B \ll \omega_o, \label{eq:as4bb} \end{equation} so we may seek to improve the CINT resolution in both range and cross-range. The expression of the CINT kernel is more complicated in this case, but it simplifies slightly when \begin{equation} \frac{\lambda_o}{\sqrt{\ell L}} \ll \frac{\lambda_o^{2/3} \ell^{1/6}}{L^{5/6}} \ll \sigma \ll \frac{\sqrt{\ell \lambda_o}}{L}, \qquad \frac{\ell}{a} = O(1). \label{eq:as3bb} \end{equation} We present the analysis that uses these conditions, which say that the fluctuations in the random medium are even stronger than in \eqref{eq:as3}, but the correlation length is not much smaller than $a$. Extensions to larger apertures are possible, although the analysis is more complicated. \section{The CINT blurring kernel} \label{sect:CINT} Here we derive the CINT convolution model. To obtain an explicit expression of the kernel \eqref{eq:f9}, we use the Gaussian pulse \eqref{eq:f10} and the Gaussian threshold windows \begin{equation} \hat \phi\left(\frac{\widetilde \omega}{\Omega}\right) = e^{-\frac{\widetilde \omega^2}{2 \Omega^2}}, \qquad \psi\left(\frac{|\widetilde {\bx}|}{X}\right) = e^{-\frac{|\widetilde {\bx}|^2}{2 \Omega^2}}, \label{eq:R1} \end{equation} with $X$ and $\Omega$ satisfying \begin{equation} {X}/{X_d} = O(1), \qquad \Omega = \left\{ \begin{array}{ll} B \quad & \mbox{in narrowband regime} \\ \\ O(\Omega_d) & \mbox{in broadband regime}. \end{array} \right. \label{eq:R2} \end{equation} As stated previously, and shown in \cite{borcea2006adaptive}, $X_d$ and $\Omega_d$ can be estimated adaptively, by optimizing the focusing of the CINT image. This is why we can assume that $X_d$ is known approximately. The same holds for $\Omega_d$, if the bandwidth is big enough. The expression of the CINT kernel is simpler in the narrowband scaling \eqref{eq:as4}, where $B \ll \Omega_d$, as shown in section \ref{sect:CINT1}, and we take $\Omega = B$. The broadband regime is discussed in section \ref{sect:CINT2}. Typically, the receivers are separated by distances of order $\lambda_o$, so that they behave collectively as an array. Since $\lambda_o \ll a$, we have $ N_r = O\left(a^2/\lambda_o^2\right) \gg 1, $ and we can approximate the sums in \eqref{eq:f9} by integrals \[ \sum_{r=1}^N \leadsto \frac{N_r}{a^2} \int_{\mathcal{A}} d \bm{x}, \] where $\mathcal{A}$ denotes the array aperture, the square of side $a$. To avoid specifying the finite aperture in the integrals, and to simplify the calculations, we use a Gaussian apodization factor \begin{equation} \psi_\mathcal{A}(\bm{x}) = \exp\left[-\frac{|\bm{x}|^2}{2(a/6)^2}\right], \label{eq:R3} \end{equation} which is negligible outside the disk of radius $a/2$. \subsection{The CINT kernel in the narrowband regime} \label{sect:CINT1} The calculation of the kernel \eqref{eq:f9} is in appendix \ref{sect:apCINT}, and we state the results in the next proposition. \vspace{0.05in} \begin{proposition} \label{prop.3} Let $\vec{\bm{z}} = (\bm{z},z_3)$, $\vec{\bm{z}}'=(\bm{z}',z_3')$ and $\vec{\bm{y}} = (\bm{y},y_3)$ be three points in the imaging region and define the center and difference vectors \[ \frac{\vec{\bm{z}} + \vec{\bm{z}}'}{2} = (\overline{\bz},\overline{z}_3), \qquad \vec{\bm{z}}-\vec{\bm{z}}' = (\widetilde{\bz},\widetilde{z}_3). \] Under the assumptions \eqref{eq:as1}-\eqref{eq:as4}, the CINT kernel \eqref{eq:f9} is approximated by \begin{align} \kappa(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}') &\approx C\exp\left[-\frac{|\overline{\bz}-\bm{y}|^2}{2 R^2} - \frac{ (\overline{z}_3-y_3)^2}{2 R_3^2} \right]\, \mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}'), \label{eq:CINTK} \end{align} where $C$ is a constant, and \begin{equation} R = \frac{L}{k_o X_e}, \qquad R_3 = \frac{c_o}{\Omega_e}, \label{eq:prop3.1} \end{equation} with $X_e$ and $\Omega_e$ defined by \begin{equation} \frac{1}{X_e^2} = \frac{1}{X_d^2} + \frac{1}{X^2} + \frac{1}{4(a/6)^2}, \qquad \frac{1}{\Omega_e^2} = \frac{1}{\Omega_d^2} + \frac{1}{\Omega^2} + \frac{1}{4 B^2}. \label{eq:R4} \end{equation} The factor $\mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')$ is complex, with absolute value \begin{align} |\mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')| = \exp \left[-\frac{\widetilde{z}_3^2}{2 \widetilde{R}^2} - \frac{|\widetilde{\bz}|^2}{2} \left( \frac{1}{\gamma X_d^2} + \frac{1}{\widetilde{R}^2} \right) \right], \label{eq:prop3.2} \end{align} where \begin{equation} \frac{1}{\gamma} = 1 - \frac{X_e^2}{4 X_d^2} > \frac{3}{4}, \quad \widetilde{R}_3 = \frac{c_o}{B}, \quad \widetilde{R} = 6 \sqrt{2} \frac{L}{k_o a}. \end{equation} \end{proposition} The parameter $R$ defined in \eqref{eq:prop3.1} is the CINT cross-range resolution limit, the length scale of exponential decay of the kernel $\kappa(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')$ with $\overline{\bz}-\bm{y}$. Definitions \eqref{eq:R4}, \eqref{eq:prop2.2} and assumption \eqref{eq:R2} give that \[ X_e = O(X_d) \ll a, \] so the resolution is worse than in homogeneous media, \begin{equation} R \gg \frac{L}{k_o a}. \label{eq:R6} \end{equation} This is due to the smoothing needed to stabilize statistically the image \cite{borcea2006adaptive}. The goal of the convex optimization \eqref{eq:f12} is to overcome this blurring and localize better the sources in cross-range. The parameter $R_3$ is the CINT range resolution limit. Because we are in the narrowband regime, we obtain from definition \eqref{eq:R4} and \eqref{eq:R2} that $ \Omega_e \approx B,$ and therefore $R_3$ is similar to the range resolution in homogeneous media, \begin{equation} R_3 = \frac{c_o}{\Omega_e} \approx \frac{c_o}{B}. \end{equation} The results obtained in \cite{borcea2015resolution} for imaging with $l_1$ optimization in homogeneous media show that it is not possible to improve the $c_o/B$ range resolution, unless the sources are very far apart in range. We cannot expect to do better in random media, so we do not seek any super-resolution in range, in the narrowband regime. Note that by the first inequality in \eqref{eq:as4p} we have $\widetilde{R} \ll X_d$, so the kernel decays with the offsets $\widetilde{\bz}$ and $\widetilde{z}$ on the length scales $\widetilde{R}$ and $\widetilde{R}_3$. These scales are, up to a constant of order one, the resolution limits of imaging in homogeneous media. \subsection{The CINT kernel in the broadband regime} \label{sect:CINT2} The expression of the CINT kernel is stated in the next proposition, proved in appendix \ref{sect:apCINT}. \begin{proposition} \label{prop.3bb} Consider the same points and notation as in Proposition \ref{prop.3}. Under the assumptions \eqref{eq:as1}-\eqref{eq:as5} and \eqref{eq:as4bb}-\eqref{eq:as3bb}, the CINT kernel is given by \begin{align} \kappa (\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}') &\approx \frac{C}{\sqrt{1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}}}\exp\left[-\frac{|\overline{\bz}-\bm{y}|^2}{2 R^2} - \frac{ \big(\overline{z}_3-y_3 + \frac{|\overline{\bz}|^2-|\bm{y}|^2}{2 L}\big)^2}{2 R_3^2\big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]} \right]\, \mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}'), \label{eq:CINTKernBB} \end{align} where $C$ is a constant, \begin{equation} \theta = \frac{6\omega_o X_e}{\Omega_e a}, \label{eq:defTheta} \end{equation} and $\mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')$ is a complex multiplicative factor with absolute value \begin{align} \left|\mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')\right| = \exp \left\{ -\frac{\widetilde{z}_3^2}{2 \widetilde{R}_3^2} - \frac{|\widetilde{\bz}|^2}{2} \left[ \frac{1}{\gamma X_d^2 } + \frac{1}{\widetilde{R}^2} \right]+ \frac{\left|\frac{(\overline{\bz}-\bm{y})}{R} \cdot\frac{\widetilde{\bz}}{\widetilde{R}}\right|^2 }{4 \theta^2\big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]} \right\} \label{eq:KernBB} \end{align} \end{proposition} Because $X_e = O(X_d)$ and $\Omega_e = \Omega_d$, we obtain from definitions \eqref{eq:defTheta} and \eqref{eq:prop2.2} that \begin{equation} \theta = O\left(\frac{\ell}{a} \right) = O(1), \label{eq:KernBB1} \end{equation} where we used the assumption \eqref{eq:as3bb} on the aperture.\footnote{In the narrowband case the aperture may be much larger than $\ell$, as in \eqref{eq:as1}. It is only in the broadband case that we take $\ell = O(a)$ to simplify the expression of the CINT kernel. } The first term in the exponential in \eqref{eq:CINTKernBB} gives the focusing in cross-range, which is the same as in the narrowband case: $|\overline{\bz}-\bm{y}| = O(R)$. This means that the denominators in \eqref{eq:CINTKernBB} are order one, \[ 1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2} = O(1). \] The second term in the exponential in \eqref{eq:CINTKernBB} gives the focusing in range. In our setting we have by the paraxial approximation that \[ \overline{z}_3 - y_3 + \frac{|\overline{\bz}|^2-|\bm{y}|^2}{2L} \approx |(\overline{\bz},\overline{z}_3)|-|(\bm{y},y_3)|, \] so CINT estimates the distance from the center of the array to $(\overline{\bz},\overline{z}_3)$ with resolution of order $R_3$. Since $\Omega_e \ll B$, this resolution is worse than in homogeneous media \[ R_3 = \frac{c_o}{\Omega_e} = O\left(\frac{c_o}{B}\right) = O(\widetilde{R}_3), \] so in the broadband regime it makes sense to seek an improvement of both the range and cross-range resolution with optimization. Equations \eqref{eq:CINTKernBB} and \eqref{eq:KernBB} show that the kernel decays with the offset $\widetilde{z}_3$ on the same scale $\widetilde{R}_3$ as before. To see the decay with the offset $\widetilde{\bz}$, we note that the last two terms in \eqref{eq:KernBB} satisfy \begin{align*} \frac{|\widetilde{\bz}|^2}{2} \left[ \frac{1}{\gamma X_d^2 } + \frac{1}{\widetilde{R}^2} \right]- \frac{\left|\frac{(\overline{\bz}-\bm{y})}{R} \cdot\frac{\widetilde{\bz}}{\widetilde{R}}\right|^2 }{4 \theta^2\big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]} &= \frac{|\widetilde{\bz}|^2}{2 R^2 \big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]} \left[ 1 + \frac{\widetilde{R}^2 \big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]}{\gamma X_d^2} \right] + \frac{\frac{|\overline{\bz}-\bm{y}|^2}{R^2} \frac{|\widetilde{\bz}|^2}{\widetilde{R}^2} - \left|\frac{(\overline{\bz} - \bm{y})}{R}\cdot\frac{\widetilde{\bz}}{\widetilde {R}}\right|^2}{4 \theta^2 \big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]} \\ &\gtrsim \frac{|\widetilde{\bz}|^2}{2 R^2 \big[1 + \frac{|\overline{\bz}-\bm{y}|^2}{2 \theta^2 R^2}\big]}, \end{align*} where we used that $\widetilde{R} \ll X_d$, as explained in the previous section. This shows that the kernel decays with the cross-range offsets on the same scale $\widetilde{R}$ as before. \subsection{The approximate convolution model} \label{sect:CINTConv} Let us discretize the imaging region $\mathfrak{D}$ defined in \eqref{eq:imreg} on a mesh with size $\vec{\bm{h}} = (h,h,h_3)$. In principle, the steps $h$ and $h_3$ may be chosen arbitrarily small, to avoid discretization error due to sources being off the mesh. However, we know from \cite{borcea2015resolution} and the analysis below and the numerical simulations that we cannot expect reconstructions at scales that are finer than the resolution limits in homogeneous media. This motivates us to formulate the inversion using the assumption that the sources are further apart than $3 \widetilde{R}$ in cross-range and $3 \widetilde{R}_3$ in range. This leads to a simpler optimization problem because by Propositions \ref{prop.3} and \ref{prop.3bb} we have \[ |\mathfrak{M}(\vec{\bm{y}},\vec{\bm{z}},\vec{\bm{z}}')| \le \exp\left(-\frac{9}{2}\right) \ll 1, \quad \mbox{if} ~ |\bm{z}-\bm{z}'| \ge 3 \widetilde{R} ~ \mbox{or}~ |z_3-z_3'| \ge 3 \widetilde{R}_3, \] and we may work only with the diagonal part of the CINT kernel. We obtain the linear system of equations \eqref{eq:f11}, with vector $\bm{u}$ of components $|\rho(\vec{\bm{z}})|^2$ at the $N_z$ mesh points in $\mathfrak{D}$. The ``data'' vector $\bm{d}$ consists of the samples of the CINT image at $N_y < N_z$ equidistant points in $\mathfrak{D}$, and in the narrowband regime the $N_y \times N_z$ matrix $\bm{\mathcal{M}}$ has entries \begin{equation} m_{\vec{\bm{y}},\vec{\bm{z}}} = C\exp\left[-\frac{|\bm{z}-\bm{y}|^2}{2 R^2} - \frac{ (\overline{z}_3-y_3)^2}{2 R_3^2} \right], \qquad B \ll \Omega_d, \label{eq:R8} \end{equation} with constant $C$. This depends only on $\vec{\bm{y}}-\vec{\bm{z}}$, so we have a convolution as stated in section \ref{sect:OPT}. In the broadband regime, the entries of $\bm{\mathcal{M}}$ are \begin{equation} m_{\vec{\bm{y}},\vec{\bm{z}}} = \frac{C}{\sqrt{1+\frac{|\bm{z}-\bm{y}|^2}{2 \theta^2 R^2}}} \exp\left[-\frac{|\bm{z}-\bm{y}|^2}{2 R^2} - \frac{ | z_3-y_3 + \frac{|\bm{z}-\bm{y}|^2}{2 L } + \frac{\bm{y} \cdot (\bm{z}-\bm{y})}{L}|^2}{2 R_3^2\big[1+\frac{|\bm{z}-\bm{y}|^2}{2 \theta^2 R^2}\big]} \right], \qquad B \gg \Omega_d, \label{eq:R8bb} \end{equation} with constant $C$. Were it not for the last term in \eqref{eq:R8bb}, we would have a convolution. This term is large only at points $\vec{\bm{y}} = (\bm{y},y_3)$ with $\bm{y}$ near the boundary of the imaging region ($|\bm{y}| \sim D < a$), because by definition \eqref{eq:defTheta} and the assumption $\theta = O(1)$ we get \[ \left| \frac{\bm{y} \cdot (\bm{z}-\bm{y})/L}{R_3}\right| = O\left(\frac{|\bm{y}|}{L} \frac{R}{R_3}\right) = O\left(\frac{|\bm{y}|}{a \theta}\right) = O\left(\frac{|\bm{y}|}{a}\right). \] For points with $|\bm{y}| \ll D < a$ the right hand side in \eqref{eq:R8bb} is approximately a function of $\vec{\bm{y}}-\vec{\bm{z}}$, corresponding to a convolution model. \section{Resolution analysis} \label{sect:res} In this section we analyze the reconstruction of the vector $\bm{u}$ of source intensities using the convex optimization formulation described in section \ref{sect:OPT}. To simplify the analysis, we treat the approximation in \eqref{eq:f11} as an equality, and study the $l_1$ optimization \begin{equation} \min_{\bm{u} \in \mathbb{R}^{N_z}} \|\bm{u}\|_{_1} ~ ~ \mbox{such that} ~~ \bm{\mathcal{M}}\, \bm{u} = \bm{d}. \label{eq:Res1} \end{equation} This neglects additive noise and random fluctuations of the CINT function, which are small in our scaling. It also implies that the sources are on the reconstruction mesh, so that the equality constraint in \eqref{eq:Res1} holds for the true discretized source intensity. Naturally, in practice the sources may lie anywhere in $\mathfrak{D}$, and noise and distortions due to the random medium play a role. This is why we use the more robust formulation \eqref{eq:f12} in the numerical simulations in section \ref{sect:num}. We expect from the study \cite{candes2014towards} of deconvolution using $l_1$ optimization that the solution of \eqref{eq:Res1} should be a good approximation of the unknown vector of intensities if the sources are well separated. We show in this section that this is indeed the case. We also consider the case of clusters of nearby sources, and show that the $l_1$ solution is useful when the clusters are well separated. The analysis is built on our recent results in \cite{borcea2015resolution}. \subsection{Definitions} \label{sect:T1Def} Let $\bm{Y} = \left\{\vec{\bm{y}}_s, ~ ~ s = 1, \ldots, N_s\right\}$ be the set that supports the unknown, point-like sources in $\mathfrak{D}$. We quantify the spatial separation between them using the following definition: \vspace{0.05in}\begin{definition} \label{def.1} The points in $\bm{Y}$ are separated by at least $\vec{\bm{H}} = (H,H,H_3)$, if the intersection of $\bm{Y}$ with any rectangular prism of sides less then $H$ in cross-range and $H_3$ in range consists of at most one point. \end{definition} \vspace{0.05in} \noindent For example, if the sources are all in the same cross-range plane, and the minimum distance between any two of them is $H_{min}$, we may take $H = H_{min}$ and $H_3 = D_3$. We search the sources on a mesh with $N_z$ points denoted generically by $\vec{\bm{z}}$. The mesh discretizes $\mathfrak{D}$, and we call it $\mathfrak{D}_z$. For simplicity we let $\bm{Y} \subset \mathfrak{D}_z$. To any $\vec{\bm{z}} \in \mathfrak{D}_z$, we associate the column vector $\bm{m}_{\vec{\bm{z}}} \in \mathbb{R}^{N_y}$ of the matrix $\bm{\mathcal{M}}$. Its entries are given in \eqref{eq:R8} in the narrowband regime and by \eqref{eq:R8bb} in the broadband regime, for $N_y < N_z$ points $\vec{\bm{y}}$ at which we sample the CINT image. \vspace{0.05in} \begin{definition} \label{def.2} We quantify the interaction between two presumed sources at $\vec{\bm{z}},\vec{\bm{z}}' \in \mathfrak{D}_{\vec{\bm{z}}}$ using the cross-correlation of the associated column vectors in $\bm{\mathcal{M}}$, \begin{equation} \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'} = \frac{\left| \left< \bm{m}_{\vec{\bm{z}}}, \bm{m}_{\vec{\bm{z}}'} \right>\right|}{\|\bm{m}_{\vec{\bm{z}}}\|_2 \|\bm{m}_{\vec{\bm{z}}'}\|_2}. \label{eq:Res2} \end{equation} Here $\left< \cdot, \cdot \right>$ is the Euclidian inner product and $\| \cdot \|_2$ is the Euclidian norm. \end{definition} \vspace{0.05in} \noindent Note that \eqref{eq:Res2} is symmetric and non-negative, and attains its maximum at $\vec{\bm{z}}' = \vec{\bm{z}}$, where $\mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}} = 1$. We will show below that \eqref{eq:Res2} decreases as the points $\vec{\bm{z}}$ and $\vec{\bm{z}}'$ grow apart. This motivates the next definition which uses $\mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}$ to measure the distance between $\vec{\bm{z}}$ and $\vec{\bm{z}}'$. \vspace{0.05in} \begin{definition} \label{def.3} We define the semimetric $\Delta:\mathfrak{D}_z \times \mathfrak{D}_z \to [0,1]$ by \begin{equation} \Delta(\vec{\bm{z}},\vec{\bm{z}}') = 1 - \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}, \qquad \forall \vec{\bm{z}},\vec{\bm{z}}' \in \mathfrak{D}_z, \label{eq:Res3} \end{equation} and let \begin{equation} \mathscr{B}_r(\vec{\bm{z}}) = \left\{ \vec{\bm{z}}' \in \mathfrak{D}_z ~ ~ \mbox{s.t.} ~ ~ \Delta(\vec{\bm{z}},\vec{\bm{z}}') < r \right\} \label{eq:Res4} \end{equation} be the open balls defined by $\Delta$. \end{definition} \vspace{0.05in} \noindent We will show that $\mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}$ decays as $\|\vec{\bm{z}}-\vec{\bm{z}}'\|_2$ grows. Thus, we say that points $\vec{\bm{z}}'$ outside $\mathscr{B}_r(\vec{\bm{z}})$ have a weaker interaction with $\vec{\bm{z}}$ than points inside $\mathscr{B}_r(\vec{\bm{z}})$. Moreover, we may relate intuitively $\Delta(\vec{\bm{z}},\vec{\bm{z}}')$ to the Euclidian distance $\|\vec{\bm{z}}-\vec{\bm{z}}'\|_2$. \vspace{0.05in} \begin{definition} \label{def.4} We define the interaction coefficient of the set $\bm{Y}$ of source locations by \begin{equation} \mathcal{I}(\bm{Y}) = \max_{\vec{\bm{z}} \in \mathfrak{D}_z} \sum_{\vec{\bm{y}} \in \bm{Y}\setminus \mathscr{N}(\vec{\bm{z}})} \mathcal{I}_{\vec{\bm{z}},\vec{\bm{y}}}, \label{eq:Res5} \end{equation} where $\mathscr{N}(\vec{\bm{z}})$ is the closest point to $\vec{\bm{z}}$ in $\bm{Y}$, as measured with the semimetric $\Delta$. \end{definition} \vspace{0.05in} \noindent In general more than one point may be closest to $\vec{\bm{z}}$. If this is so, we let $\mathscr{N}(\vec{\bm{z}})$ be any such point. \subsection{Results} The results stated here describe the relation between the reconstruction $\bm{u}_\star$, the solution of the convex optimization problem \eqref{eq:Res1}, and the true unknown vector $\bm{u}$ of source intensities. The next theorem shows that $\bm{u}_\star$ is essentially supported in the set $\bm{Y}$, when the points there are well separated. \vspace{0.05in} \begin{theorem} \label{thm.1} Suppose that the source locations in $\bm{Y}$ are separated by at least $\vec{\bm{H}} = (H,H,H_3)$ in the sense of Definition \ref{def.1}, where \begin{equation} \frac{H}{R} = \alpha, ~ ~ \frac{H_3}{R_3} = \alpha_3, \label{eq:asH} \end{equation} for $\alpha, \alpha_3 > 1$. Take $r \in (0,1)$ small enough so that the balls $\mathscr{B}_r(\vec{\bm{y}}_s)$ centered at $\vec{\bm{y}}_s \in \bm{Y}$, for $s = 1, \ldots, N_s$, are disjoint. Let $\bm{u}_\star$ be the $l_1$ minimizer in \eqref{eq:Res1}, and decompose it as $ \bm{u}_\star = \bm{u}_\star^{(i)} + \bm{u}_\star^{(o)}, $ where $\bm{u}_\star^{(i)}$ is supported in the union $\bigcup_{s=1}^{N_s}\mathscr{B}_r(\vec{\bm{y}}_s)$ of balls centered at the points in $\bm{Y}$, and $\bm{u}_\star^{(o)}$ is supported in the complement of this union. Then, there exists a constant $C$ that is independent of $\alpha$ and $\alpha_3$, such that \begin{equation} \|\bm{u}_\star^{(o)}\|_1 \le \frac{C}{r} \mathfrak{F}(\alpha,\alpha_3) \|\bm{u}_\star\|_1, \label{eq:Res6} \end{equation} where $\mathfrak{F}(\alpha,\alpha_3)$ is function that decays with $\alpha$ and $\alpha_3$. In the narrowband case it is given by \begin{equation} \mathfrak{F}(\alpha,\alpha_3)= \frac{ \exp \left[- \left(\frac{\min\{\alpha,\alpha_3\}}{4} \right)^2\right]}{\alpha^2 \alpha_3}, \label{eq:defFnb} \end{equation} for arbitrary $\alpha, \alpha_3 > 1$. In the broadband case $\alpha_3 > 8 \alpha/\theta$, and \begin{equation} \mathfrak{F}(\alpha,\alpha_3)= \frac{ \exp \left[-\frac{1}{2} \left(\frac{\alpha}{4}\right)^2\right] + \exp[-\alpha]}{\alpha^3}. \label{eq:defFbb} \end{equation} \end{theorem} \vspace{0.0in} \noindent Note that the scales of separation between the sources are the resolution parameters $R$ and $R_3$ of CINT. The parameters $\alpha$ and $\alpha_3$ in the separation assumption may be any non-negative real numbers, but the statement of the theorem is useful only when the coefficient in front of $\|\bm{u}_\star\|_1$ in \eqref{eq:Res6} is smaller than one. This happens for large enough $\alpha$ and $\alpha_3$. The larger the separation between the sources, the smaller the right hand side in \eqref{eq:Res6} is, and the better the concentration of the support of $\bm{u}_\star$ near the points in $\bm{Y}$. The next corollary gives an estimate of the error of the reconstruction. \vspace{0.05in} \begin{corollary} \label{cor.1} Let $\bm{u} \in \mathbb{R}^{N_z}$ be the vector of true source intensities, and use the same assumptions and notation as in Theorem \ref{thm.1}. Denote the entries of the $l_1$ minimizer $\bm{u}_\star$ by $u_\star(\vec{\bm{z}})$, where $\vec{\bm{z}}$ are the $N_z$ points on the mesh $\mathfrak{D}_z$. Define the effective reconstructed source intensity vector $\overline{\bm{u}}_\star \in \mathbb{R}^{N_z}$, with entries \begin{equation} \overline{u}_\star(\vec{\bm{z}}) = \left\{ \begin{array}{ll} \sum_{\vec{\bm{z}}' \in \mathscr{B}_r(\vec{\bm{z}})} u_\star(\vec{\bm{z}}') \, \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}, \quad &\mbox{if} ~ \vec{\bm{z}} \in \bm{Y}, \\ 0, &\mbox{otherwise}. \end{array} \right. \label{eq:Res7} \end{equation} Then, we have the following estimate of the relative error \begin{equation} \frac{\| \bm{u} - \overline{\bm{u}}_\star\|_1}{\|\bm{u}\|_1} \le \frac{C}{ r} \mathfrak{F}(\alpha,\alpha_3), \label{eq:Res8} \end{equation} with constant $C$ independent of $\alpha$ and $\alpha_3$. \end{corollary} \vspace{0.05in} \noindent This result says that when the sources are far apart, the effective intensity vector $\overline{\bm{u}}_\star$ is close to the true solution $\bm{u}$. By definition, the support of $\overline{\bm{u}}_\star$ is at the source points in $\bm{Y}$. Its entries $\overline{u}_\star(\vec{\bm{z}})$ at $\vec{\bm{z}} \in \bm{Y}$ are weighted averages of the entries of $\bm{u}_\star$ at points $\vec{\bm{z}}' \in \mathscr{B}_r(\vec{\bm{z}})$, with weights $\mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}$. When $r$ is small, these weights are close to one, so $\overline{u}_\star(\vec{\bm{z}})$ is approximately the sum of the entries of $\bm{u}_\star$ supported in the ball $\mathscr{B}_r(\vec{\bm{z}})$. Theorem \ref{thm.1} and its corollary are not useful when the sources are clustered together. The next result deals with this case, when the clusters are well separated. \vspace{0.05in} \begin{theorem} \label{thm.2} Let $\epsilon \in (0,1)$ be such that there exists a subset $\bm{Y}_\epsilon$ of $\bm{Y}$, satisfying \begin{equation} \bm{Y} \subset \bigcup_{\vec{\bm{z}} \in \bm{Y}_\epsilon}\, \mathscr{B}_\epsilon(\vec{\bm{z}}), \qquad \mathscr{B}_\epsilon(\vec{\bm{z}}) \bigcap \mathscr{B}_\epsilon(\vec{\bm{z}}') = \emptyset, ~ ~ \forall \,\vec{\bm{z}},\vec{\bm{z}}' \in \bm{Y}_\epsilon, ~ ~ \vec{\bm{z}} \ne \vec{\bm{z}}'. \label{eq:Res9} \end{equation} Suppose that the points in $\bm{Y}_\epsilon$ are separated by at least least $\vec{\bm{H}} = (H,H,H_3)$ in the sense of Definition \ref{def.1}, where $H$ and $H_3$ satisfy \eqref{eq:asH}, for some $\alpha, \alpha_3 > 1$. Let $r$ satisfy $\epsilon < r < 1$, and decompose the $l_1$ minimizer $\bm{u}_\star$ in \eqref{eq:Res1} as $ \bm{u}_\star = \bm{u}_\star^{(i)} + \bm{u}_\star^{(o)}, $ where $\bm{u}_\star^{(i)}$ is supported in the union $\bigcup_{\vec{\bm{z}} \in \bm{Y}_\epsilon}\mathscr{B}_r(\vec{\bm{z}})$ of balls centered at the points in $\bm{Y}_\epsilon$, and $\bm{u}_\star^{(o)}$ is supported in the complement of this union. There exists a constant $C$ that is independent of $\alpha$ and $\alpha_3$, such that \begin{equation} \|\bm{u}_\star^{(o)}\|_1 \le \frac{C}{ r} \mathfrak{F}(\alpha,\alpha_3)\|\bm{u}_\star\|_1 + \frac{\epsilon}{r} \|\bm{u}\|_1, \label{eq:Res10} \end{equation} where $\bm{u}$ is the vector of true source intensities. \end{theorem} \vspace{0.05in} \noindent Equation \eqref{eq:Res9} says that we can cover the sources with disjoint balls of radius $\epsilon$, centered at the points in $\bm{Y}_\epsilon$. Thus, we call $\bm{Y}_\epsilon$ the effective support of the sources, and $\epsilon$ the radius of the clusters. The statement of the theorem is that when the clusters are well separated, and they have small radius, the $l_1$ minimizer will be supported in their vicinity. As expected, \eqref{eq:Res10} converges to \eqref{eq:Res6} in the limit $\epsilon \to 0$. \subsection{Proofs} \label{sect:proofs} We use \cite[Theorem 4.1 and Corollary 4.2]{borcea2015resolution} which state that for the decomposition of the $l_1$ minimizer $\bm{u}_\star$ as in Theorem \ref{thm.1}, and for the effective reconstructed source intensity vector $\overline{\bm{u}}_\star$ defined in \eqref{eq:Res7}, we have \begin{equation} \|\bm{u}_\star^{(o)}\|_1 \le \frac{2 \mathcal{I}(\bm{Y})}{r} \|\bm{u}_\star\|_1 ~ ~ \mbox{and} ~ ~ \frac{\|\bm{u} - \overline{\bm{u}}_\star\|_1}{\|\bm{u}\|_1} \le \frac{2 \mathcal{I}(\bm{Y})}{r}. \label{eq:Pf1} \end{equation} To determine the interaction coefficient $\mathcal{I}(\bm{Y})$ of the sources, we estimate first the cross-correlations $\mathcal{I}_{\vec{\bm{z}}, \vec{\bm{z}}'}$: \vspace{0.05in} \begin{lemma} \label{lem.1} The cross-correlations $\mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}$ defined in \eqref{eq:Res2} satisfy \begin{equation} \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'} \approx \exp\left[-\frac{|\bm{z}-\bm{z}'|^2}{4 R^2} -\frac{(z_3-z_3')^2 }{4R_3^2}\right], \label{eq:Pf2} \end{equation} in the narrowband regime and \begin{equation} \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'} \le C \exp\left[-\frac{|\bm{z}-\bm{z}'|^2}{8 R^2} -\frac{\theta\left|z_3-z_3' + \frac{|\bm{z}|^2-|\bm{z}'|^2}{2L}\right|}{R_3}\right], \label{eq:Pf2bb} \end{equation} in the broadband regime, for all $ \vec{\bm{z}},\vec{\bm{z}}' \in \mathfrak{D}_z$, with $\vec{\bm{z}} = (\bm{z},z_3)$ and $ \vec{\bm{z}}' = (\bm{z}',z_3')$. The constant $C$ in \eqref{eq:Pf2bb} is given by \begin{equation} C =\frac{3 e^{-\theta^2}}{2 \mbox{erfc}(\sqrt{2} \, \theta)}, \end{equation} with $\theta$ defined in \eqref{eq:defTheta} of order one. \end{lemma} \vspace{0.05in} \noindent The proof is in Appendix \ref{ap:PfLem1}. The next lemma, proved in sections \ref{sect:pfLemnb} and \ref{sect:pfLembb}, gives the estimate of $\mathcal{I}(\bm{Y})$, that combined with \eqref{eq:Pf1} proves Theorem \ref{thm.1} and Corollary \ref{cor.1}. \vspace{0.05in} \begin{lemma} \label{lem.2} Suppose that the points in $\bm{Y}$ are separated by at least $\vec{\bm{H}} = (H,H,H_3)$ in the sense of Definition \ref{def.1}, where $H$ and $H_3$ are as in \eqref{eq:asH}, for some $\alpha,\alpha_3 > 1$. The interaction coefficient satisfies \begin{equation} \mathcal{I}(\bm{Y}) \le C \mathfrak{F}(\alpha,\alpha_3) \label{eq:Pf3} \end{equation} for a constant $C$ that is independent of $\alpha$ and $\alpha_3$ and $\mathfrak{F}(\alpha,\alpha_3)$ as defined in Theorem \ref{thm.1}. \end{lemma} \vspace{0.05in} To prove Theorem \ref{thm.2}, we use \cite[Theorem 4.4]{borcea2015resolution} which states that \begin{equation} \|\bm{u}_\star^{(o)}\|_1 \le \frac{2 \mathcal{I}(\bm{Y}_\epsilon)}{r} \|\bm{u}_\star\|_1 + \frac{\|\bm{u}\|-\|\overline{\bm{u}}\|_1}{r}, \label{eq:PfT2.1} \end{equation} for $\overline{\bm{u}}$ defined by \begin{equation} \overline{\bm{u}}(\vec{\bm{z}}) = \left\{ \begin{array}{ll} \sum_{\vec{\bm{z}}' \in \mathscr{B}_\epsilon(\vec{\bm{z}})\cap \bm{Y}} u(\vec{\bm{z}}') \, \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}, \quad &\mbox{if} ~ \vec{\bm{z}} \in \bm{Y}_\epsilon, \\ 0, &\mbox{otherwise}. \end{array} \right. \label{eq:PfT2.2} \end{equation} The interaction coefficient of the effective support $\bm{Y}_\epsilon$ is as in Lemma \ref{lem.2}, so it remains to estimate the last term in \eqref{eq:PfT2.1}. Let us define the set \[ S_{\vec{\bm{z}},\epsilon} = \left\{ \vec{\bm{z}}' \in \bm{Y} ~ ~ \mbox{s.t.} ~ ~ \vec{\bm{z}}' \in \mathscr{B}_\epsilon(\vec{\bm{z}})\right\}, \] so that with definition \eqref{eq:PfT2.2} we can write \[ \|\overline{\bm{u}}\|_1 = \sum_{\vec{\bm{z}} \in \bm{Y}_\epsilon} \left| \sum_{\vec{\bm{z}}' \in S_{\vec{\bm{z}},\epsilon}} u(\vec{\bm{z}}') \, \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'} \right| = \sum_{\vec{\bm{z}} \in \bm{Y}_\epsilon} \sum_{\vec{\bm{z}}' \in S_{\vec{\bm{z}},\epsilon}} |u(\vec{\bm{z}}')| \, \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}, \] where we used that by definition $ \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'} \ge 0$ and $u(\vec{\bm{z}}') = |\rho(\vec{\bm{z}}')|^2 \ge 0$. Since the norm of the vector of the true source intensities is given by \[ \|\bm{u}\|_1 = \sum_{\vec{\bm{z}} \in \bm{Y}} |u(\vec{\bm{z}})| = \sum_{\vec{\bm{z}} \in \bm{Y}_\epsilon} \sum_{\vec{\bm{z}}' \in S_{\vec{\bm{z}}\,\epsilon}} |u(\vec{\bm{z}}')|, \] we obtain that \begin{equation} \|\bm{u}\|_1 - \|\overline{\bm{u}}\|_1 = \sum_{\vec{\bm{z}} \in \bm{Y}_\epsilon} \sum_{\vec{\bm{z}}' \in S_{\vec{\bm{z}},\epsilon}} |u(\vec{\bm{z}}')| \left(1 - \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'}\right) < \epsilon \sum_{\vec{\bm{z}} \in \bm{Y}_\epsilon} \sum_{\vec{\bm{z}}' \in S_{\vec{\bm{z}},\epsilon}} |u(\vec{\bm{z}}')| = \epsilon \|\bm{u}\|_1. \label{eq:PfT2.4} \end{equation} The inequality is because \[ \Delta(\vec{\bm{z}},\vec{\bm{z}}') = 1 - \mathcal{I}_{\vec{\bm{z}},\vec{\bm{z}}'} < \epsilon, \quad \forall \, \vec{\bm{z}}' \in \mathscr{B}_\epsilon(\vec{\bm{z}}). \] The statement of Theorem \ref{thm.2} follows by substitution of \eqref{eq:PfT2.4} in \eqref{eq:PfT2.1}, and using the estimate in Lemma \ref{lem.2}, with $\bm{Y}$ replaced by $\bm{Y}_\epsilon$. $\Box$. \subsubsection{Proof of Lemma \ref{lem.2} in the narrowband regime} \label{sect:pfLemnb} \vspace{0.05in} Recall Definition \ref{def.4} of $\mathcal{I}(\bm{Y})$, and let $\vec{\bm{z}} \in \mathfrak{D}_z$ be the maximizer of the sum in \eqref{eq:Res5}, so that \begin{equation} \mathcal{I}(\bm{Y}) = \sum_{\vec{\bm{y}} \in \bm{Y}\setminus \mathscr{N}(\vec{\bm{z}})} \mathcal{I}_{\vec{\bm{z}},\vec{\bm{y}}}. \label{eq:Pf4} \end{equation} We denote the components of $\vec{\bm{z}}$ by $z_j$, for $j=1,2,3$, and conclude from the source separation assumption in the lemma that the set \begin{equation} \mathscr{S}_{\vec{\bm{z}}} = \left\{ \vec{\bm{z}}'= (z_1',z_2',z_3') \in \mathfrak{D} ~ ~ \mbox{s.t.} ~ ~ |z_j-z_j'| < H, ~j = 1, 2, ~ ~ |z_3-z_3'| < H_3\right\} \label{eq:defSz} \end{equation} contains at most one point in $\bm{Y}$. This may be $\mathscr{N}(\vec{\bm{z}})$, the closest point in $\bm{Y}$ to $\vec{\bm{z}}$ with respect to the semimetric $\Delta$, satisfying \begin{equation} \mathcal{I}_{\vec{\bm{z}},\mathscr{N}(\vec{\bm{z}})} \ge \mathcal{I}_{\vec{\bm{z}},\vec{\bm{y}}}. \qquad \forall \, \vec{\bm{y}} \in \bm{Y}. \label{eq:Pf5} \end{equation} Alternatively, $\mathscr{S}_{\vec{\bm{z}}}$ may be empty or contain another point in $\bm{Y}$. In either case, we obtain from equations \eqref{eq:Pf4} and \eqref{eq:Pf5} that \[ \mathcal{I}(\bm{Y}) \le \sum_{\vec{\bm{y}} \in \bm{Y}\setminus \mathscr{S}_{\vec{\bm{z}}}} \mathcal{I}_{\vec{\bm{z}},\vec{\bm{y}}}, \] and from the bound in Lemma \ref{lem.1}, \begin{equation} \mathcal{I}(\bm{Y}) \le C \sum_{\vec{\bm{y}} \in \bm{Y}\setminus \mathscr{S}_{\vec{\bm{z}}}} \mathscr{E}_{\vec{\bm{z}},\vec{\bm{y}}}, ~ ~\mbox{for} ~ ~ \mathscr{E}_{\vec{\bm{z}},\vec{\bm{y}}} = \exp\left[-\frac{|\bm{z}-\bm{y}|^2}{8 R^2} -\frac{(z_3-y_3)^2}{R_3}\right]. \label{eq:Pf6} \end{equation} Using again the source separation assumption in the lemma, we conclude that for any $\vec{\bm{y}} \in \bm{Y}$, we can define a set $\mathscr{H}_{\vec{\bm{y}}}$, in the form of a rectangular prism of sides $H/2$ in cross-range and $H_3/2$ in range, satisfying \begin{equation} \vec{\bm{y}} \in \mathscr{H}_{\vec{\bm{y}}} ~ ~ \mbox{and} ~ ~ \mathscr{H}_{\vec{\bm{y}}} \bigcap \mathscr{H}_{\vec{\bm{y}}'} = \emptyset, \qquad \forall \, \vec{\bm{y}} \ne \vec{\bm{y}}' \in \bm{Y}. \label{eq:Pf.7} \end{equation} There are many such sets, but we make our choice so that $\vec{\bm{y}}$ is the furthermost point to $\vec{\bm{z}}$ in $\mathscr{H}_{\vec{\bm{y}}}$, satisfying \begin{equation} \mathscr{E}_{\vec{\bm{z}},\vec{\bm{y}}} \le \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'}, \qquad \forall \, \vec{\bm{z}}' \in \mathscr{H}_{\vec{\bm{y}}}. \label{eq:Pf.8} \end{equation} This allows us to write \begin{equation} \mathscr{E}_{\vec{\bm{z}},\vec{\bm{y}}} \le \frac{8}{H^2 H_3} \int_{\mathscr{H}_{\vec{\bm{y}}}} d \vec{\bm{z}}' \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'}, \qquad \forall \, \vec{\bm{y}} \in \bm{Y}, \label{eq:Pf.9} \end{equation} and obtain from \eqref{eq:Pf6} that \begin{equation} \mathcal{I}(\bm{Y}) \le \frac{8 C^2}{H^2 H_3} \sum_{\bm{y} \in \bm{Y} \setminus \mathscr{S}_{\vec{\bm{z}}}} \int_{\mathscr{H}_{\vec{\bm{y}}}} d \vec{\bm{z}}' \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} \le \frac{8 C^2}{H^2 H_3} \int_{\mathbb{R}^3 \setminus \mathscr{S}_{\vec{\bm{z}},\frac{1}{2}}} d \vec{\bm{z}}' \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'}, \label{eq:Pf10} \end{equation} with $\mathscr{S}_{\vec{\bm{z}}, \frac{1}{2}}$ defined as in \eqref{eq:defSz}, with half the values of $H$ and $H_3$, \begin{equation} \mathscr{S}_{\vec{\bm{z}},\frac{1}{2}} = \left\{ \vec{\bm{z}}'= (z_1',z_2',z_3') \in \mathfrak{D} ~ ~ \mbox{s.t.} ~ ~ |z_j-z_j'| < H/2, ~j = 1, 2, ~ ~ |z_3-z_3'| < H_3/2\right\}. \label{eq:defSzhalf} \end{equation} The last inequality in \eqref{eq:Pf10} is because the integrand is positive, the sets $\mathscr{H}_{\vec{\bm{y}}}$ are disjoint, and \[ \bigcup_{\vec{\bm{y}} \in \bm{Y}\setminus \mathscr{S}_{\vec{\bm{z}}}} \mathscr{H}_{\vec{\bm{y}}} \subset \mathbb{R}^3 \setminus \mathscr{S}_{\vec{\bm{z}},\frac{1}{2}}. \] We estimate the integral in \eqref{eq:Pf10} by decomposing the set $\mathscr{S}_{\vec{\bm{z}},\frac{1}{2}}^c = \mathbb{R}^3\setminus \mathscr{S}_{\vec{\bm{z}},\frac{1}{2}}$ in three components denoted by $\mathscr{C}_{\vec{\bm{z}},j}$, where \[ \mathscr{C}_{\vec{\bm{z}},j} = \left\{\vec{\bm{z}}' \in \mathbb{R}^3 ~ ~ \mbox{s.t.} ~ ~ |z_j-z_j'| \ge H/2 \right\}, \quad j = 1,2, \] and \[ \mathscr{C}_{\vec{\bm{z}},3} = \left\{\vec{\bm{z}}' \in \mathbb{R}^3 ~ ~ \mbox{s.t.} ~ ~ |z_j-z_j'| < H/2, ~ |z_3-z_3'| \ge H_3/2 \right\}. \] We have \begin{align} \frac{8}{H^2 H_3} \int_{\mathscr{C}_{\vec{\bm{z}},1}} d \vec{\bm{z}}' \, \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} &= \frac{8}{H^2 H_3} \int_{|z_1'-z_1| \ge H/2} d z_1'\, e^{-\frac{(z_1'-z_1)^2}{4 R^2}} \int_{-\infty}^\infty d z_2'\, e^{-\frac{(z_2'-z_2)^2}{4 R^2}} \int_{-\infty}^\infty d z_3' \, e^{-\frac{(z_3'-z_3)^2}{4 R^2_3}} \nonumber \\ &=\frac{32\pi R R_3}{H^2 H_3} 2 \int_{H/2}^\infty dt \, e^{-\frac{t^2}{4 R^2}} \nonumber \\ &= \frac{64 \sqrt{\pi}}{\alpha^2 \alpha_3} \mbox{erfc} \left(\frac{\alpha}{4}\right), \label{eq:Pf11} \end{align} where we evaluated the integrals over $z_3'$ and $z_2'$ in the second line, and used \eqref{eq:asH} in the last line. The integral over $\mathscr{C}_{\vec{\bm{z}},2}$ is the same, so it remains to estimate \begin{align} \frac{8}{H^2 H_3} \int_{\mathscr{C}_{\vec{\bm{z}},3}} d \vec{\bm{z}}' \, \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} = \frac{8}{H^2 H_3} \int_{|z_1'-z_1| < H/2} d z_1'\, e^{-\frac{(z_1'-z_1)^2}{4 R^2}} \int_{|z_2'-z_2| < H/2} d z_2'\, e^{-\frac{(z_2'-z_2)^2}{4 R^2}}\nonumber \\ \times \int_{|z_3'-z_3| \ge H_3/2} d z_3' \, e^{-\frac{(z_3'-z_3)^2}{4 R^2_3}}. \label{eq:Pf12} \end{align} We bound the integrals over $z_1'$ and $z_2'$ by those of the real line, and rewrite the integral over $z_3'$ in terms of the complementary error function, to obtain \begin{align} \frac{8}{H^2 H_3 } \int_{\mathscr{C}_{\vec{\bm{z}},3}} d \vec{\bm{z}}' \, \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} &\le \frac{64 \sqrt{\pi}}{\alpha^2 \alpha_3} \mbox{erfc} \left(\frac{\alpha_3}{4}\right). \label{eq:Pf14} \end{align} The statement of Lemma \ref{lem.2} follows from \eqref{eq:Pf10}, with right hand side given by the sum of the integrals over $\mathscr{C}_{\vec{\bm{z}},1}$ and $\mathscr{C}_{\vec{\bm{z}},2}$ estimated in \eqref{eq:Pf11}, and over $\mathscr{C}_{\vec{\bm{z}},3}$, estimated in \eqref{eq:Pf14}. $\Box$ \subsubsection{Proof of Lemma \ref{lem.2} in the broadband regime} \label{sect:pfLembb} We obtain from Lemma \ref{lem.1}, the same way as above, and with the same notation, that \begin{equation} \mathcal{I}(\bm{Y}) \le C \sum_{\vec{\bm{y}} \in \bm{Y}\setminus \mathscr{S}_{\vec{\bm{z}}}} \mathscr{E}_{\vec{\bm{z}},\vec{\bm{y}}}, ~ ~\mbox{for} ~ ~ \mathscr{E}_{\vec{\bm{z}},\vec{\bm{y}}} = \exp\left[-\frac{|\bm{z}-\bm{y}|^2}{8 R^2} -\frac{\theta\left|z_3-y_3 + \frac{|\bm{z}|^2-|\bm{y}|^2}{2L}\right|}{R_3}\right]. \label{eq:Pf6bb} \end{equation} We also define as before, using the source separation assumption in the lemma, the set $\mathscr{H}_{\vec{\bm{y}}}$, satisfying \eqref{eq:Pf.7} and \eqref{eq:Pf.8}. This leads us to the bound \eqref{eq:Pf10}, with the set $\mathscr{S}_{\vec{\bm{z}},\frac{1}{2}}$ defined in \eqref{eq:defSzhalf}. We estimate the integral in \eqref{eq:Pf10} by decomposing the set $\mathbb{R}^3\setminus \mathscr{S}_{\vec{\bm{z}},\frac{1}{2}}$ in three parts $S_{\vec{\bm{z}},j}$, where \[ S_{\vec{\bm{z}},j} = \left\{\vec{\bm{z}}' \in \mathbb{R}^3 ~ ~ \mbox{s.t.} ~ ~ |z_j-z_j'| \ge H/2 \right\}, \quad j = 1,2, \] and \[ S_{\vec{\bm{z}},3} = \left\{\vec{\bm{z}}' \in \mathbb{R}^3 ~ ~ \mbox{s.t.} ~ ~ |z_j-z_j'| < H/2, ~ |z_3-z_3'| \ge H_3/2 \right\}. \] We have \begin{align} \frac{8}{H^2 H_3} \int_{S_{\vec{\bm{z}},1}} d \vec{\bm{z}}' \, \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} &= \frac{8}{H^2 H_3} \int_{|z_1'-z_1| \ge H/2} d z_1'\, e^{-\frac{(z_1'-z_1)^2}{8 R^2}} \int_{-\infty}^\infty d z_2'\, e^{-\frac{(z_2'-z_2)^2}{8 R^2}} \int_{-\infty}^\infty d z_3' \, e^{-\frac{\theta}{ R_3} \left|z_3'-z_3 + \frac{|\bm{z}|^2-|\bm{z}'|^2}{2 L} \right|} \nonumber \\ &=\frac{128 \pi R^2 R_3}{H^2 H_3 \theta} \mbox{erfc}\left(\frac{H}{4 \sqrt{2} R}\right)\nonumber \\ &\le \frac{128\pi }{\alpha^2 \alpha_3 \theta}e^{-\left(\frac{h}{4 \sqrt{2} R}\right)^2} \nonumber \\ &\le \frac{16 \pi}{\alpha^3} e^{-\left(\frac{\alpha}{4 \sqrt{2}}\right)^2}, \label{eq:Pf11bb} \end{align} where we evaluated the integrals over $z_3'$ and $z_2'$ in the second line. The first inequality is because the complementary error function satisfies $\mbox{erfc}(x) \le \exp(-x^2)$, and the second inequality is by the assumption on $\alpha_3$. The integral over $S_{\vec{\bm{z}},2}$ is the same, so it remains to estimate \begin{align} \frac{8}{H^2 H_3} \int_{S_{\vec{\bm{z}},3}} d \vec{\bm{z}}' \, \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} \approx \frac{8}{H^2 H_3} \int_{|z_1'-z_1| < H/2} d z_1'\, e^{-\frac{(z_1'-z_1)^2}{8 R^2}} \int_{|z_2'-z_2| < H/2} d z_2'\, e^{-\frac{(z_2'-z_2)^2}{8 R^2}}\nonumber \\ \times \int_{|z_3'-z_3| \ge H_3/2} d z_3' \, e^{-\frac{\theta}{R_3} \left|z_3'-z_3 + \frac{|\bm{z}'|^2-|\bm{z}|^2}{2 L} \right|}. \label{eq:Pf12bb} \end{align} Because $|z_j'-z_j| < H/2$ for $j = 1,2$, and therefore $|\bm{z}'-\bm{z}| < H/\sqrt{2}$, we have \begin{align} \theta \left|\frac{|\bm{z}'|^2-|\bm{z}|^2}{2 L R_3}\right| \le \frac{\theta|\bm{z}'-\bm{z}| |\bm{z}'+\bm{z}|}{2 L R_3} \le \frac{\theta H D}{2 L R_3} = \frac{\theta \alpha D R}{2 L R_3} = \frac{3 D \alpha }{a} \lesssim {3 \alpha}. \end{align} Here we used that $|\bm{z}'+\bm{z}| \le D \sqrt{2}$ for all $\bm{z},\bm{z}' \in [-D/2,D/2] \times [-D/2,D/2]$, and substituted $H = R \alpha$ and definitions \eqref{eq:prop3.1} and \eqref{eq:defTheta}. The last inequality is by assumption \eqref{eq:as5}. With this bound we can estimate the integral over $z_3'$ as follows \begin{align} \int_{|z_3'-z_3| \ge H_3/2} d z_3' \, e^{-\frac{\theta}{ R_3} \left|z_3'-z_3 + \frac{|\bm{z}'|^2-|\bm{z}|^2}{2 L} \right|} &= \frac{R_3}{\theta} \left[\int_{\frac{\theta H_3}{2R_3}}^\infty dt \, e^{-\left|-t + \frac{\theta(|\bm{z}'|^2-|\bm{z}|^2)}{2 L R_3}\right|} + \int_{\frac{\theta H_3}{2 R_3}}^\infty dt \, e^{-\left|t + \frac{\theta(|\bm{z}'|^2-|\bm{z}|^2)}{2 LR_3}\right|}\right] \nonumber \\ & \le \frac{2 R_3}{\theta} \int_{\frac{\theta H_3}{2 R_3}}^\infty dt \, e^{-(t-3 \alpha)} \nonumber \\ &= \frac{2 R_3}{\theta} e^{- \left(\frac{\theta\alpha_3}{2 } - 3 \alpha \right)} \int_0^\infty dt e^{-t} \nonumber \\ &\le \frac{2 R_3}{\theta} e^{-\alpha}, \label{eq:Pf13bb} \end{align} where the last inequality is by the assumption $H_3/R_3 = \alpha_3 \ge 8 \alpha/\theta$. Substituting in \eqref{eq:Pf12bb} we get \begin{align} \frac{8}{H^2 H_3 } \int_{S_{\vec{\bm{z}},3}} d \vec{\bm{z}}' \, \mathscr{E}_{\vec{\bm{z}},\vec{\bm{z}}'} &\le \frac{16 R_3}{H^2 H_3 \theta}e^{-\alpha} \int_{|z_1'-z_1| < H/2} d z_1'\, e^{-\frac{(z_1'-z_1)^2}{8 R^2}} \int_{|z_2'-z_2| < H/2} d z_2'\, e^{-\frac{(z_2'-z_2)^2}{8 R^2}} \nonumber \\ &\le \frac{128 \pi R^2 R_3}{H^2 H_3 \theta} e^{-\alpha} = \frac{128 \pi}{\alpha^2 \alpha_3 \theta} \le \frac{16 \pi}{\alpha^3} e^{-\alpha}. \label{eq:Pf14bb} \end{align} Here we bounded each Gaussian integral by $2 \sqrt{2 \pi} R$, which is the integral over the real line, and used again the assumption on $\alpha_3$. The statement of Lemma \ref{lem.2} follows from \eqref{eq:Pf10}, with right hand side given by the sum of the integrals over $S_{\vec{\bm{z}},1}$ and $S_{\vec{\bm{z}},2}$ estimated in \eqref{eq:Pf11bb}, and over $S_{\vec{\bm{z}},3}$, estimated in \eqref{eq:Pf14bb}. $\Box$ \section{Numerical simulations} \label{sect:num} We present here numerical simulations obtained with the wave propagation model described in equation \eqref{eq:a6}, with random travel time fluctuations computed by the line integrals in \eqref{eq:a7p}, in one realization of the random process $\mu$. We generate it numerically using random Fourier series \cite{devroye2006nonuniform}, for the Gaussian autocorrelation \eqref{eq:a2}. All the length scales are normalized by $\ell$ in the simulations, and are chosen to satisfy marginally the assumptions in section \ref{sect:sc4}. Specifically, we take $\lambda_o = 1.75 \cdot 10^{-4} \,\ell $ and $L = 800\, \ell,$ so that \[ \sqrt{\lambda_o L} = 0.12 < \ell < 9.72 \, \ell = (\lambda_o L^3)^{1/4}, \] and the aperture is $a = 16 \ell$. This is slightly larger than the bound in \eqref{eq:as1}, but we also have the apodization \eqref{eq:R3}. We verify that \[ \frac{\lambda_o}{\sqrt{\ell L}} = 6.17 \cdot 10^{-7} \ll \frac{\sqrt{\lambda_o \ell}}{L} = 5.22 \cdot 10^{-6} \ll \left(\frac{\ell}{L}\right)^{3/2} = 4.42 \cdot 10^{-5}, \] and we take the strength of the fluctuations $\sigma = 1.5 \cdot 10^{-6}$. With this choice we obtain \[ \frac{\Omega_d}{\omega_o} = 0.083 \quad \mbox{and} \quad X_d = 0.14 \ell. \] We show results in two dimensions, for a linear array and a narrowband regime with bandwidth $B = 0.0032 \omega_o$. Since in this regime we can only expect improvements in the cross-range localization of the sources, we focus attention at a given range, and display cross-range sections of the images. The migration and CINT images are calculated as in equations \eqref{eq:f5} and \eqref{eq:f6} from the data contaminated with $5\%$ additive, uncorrelated, Gaussian noise. The thresholding parameters in the CINT image formation are $\Omega = B/2$ and $X = X_d/2$, and the sources are off the reconstruction mesh. The mesh size is $H = \lambda_o L/(6X)$, unless stated otherwise The optimization formulation is \begin{equation} \min_{\bm{u}} \|\bm{u}\|_1 \quad \mbox{such that } \quad \|\bm{\mathcal{M}}\, \bm{u} - \bm{d}\|_{2} \le \delta, \label{eq:OPTForm} \end{equation} with tolerance $\delta = 0.05 \|\bm{d}\|_2$. The entries of matrix $\bm{\mathcal{M}}$ are as defined in \eqref{eq:R8}, with constant $C$ given in appendix \ref{sect:apCINT}. For comparison, we also present the results of a direct application of $L_1$ optimization to the array data, without using the CINT image formation, as in \cite{borcea2015resolution}. We call this method "direct $l_1$ optimization" and refer to \cite[Appendix A]{borcea2015resolution} for details. We also refer to \eqref{eq:OPTForm} as "$l_1$ optimization" and solve it with the software package \cite{cvx}. \begin{figure}[t] \begin{centering} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_Slice_CINT}}} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_Slice_KM}}} \vspace{-0.15in} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_Slice_L1}}} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_Slice_L1D}}} \par\end{centering} \vspace{-0.2in} \caption{Results in one realization of the random medium. The abscissa is the cross-range in units $\lambda_o L/X$, and the source locations are shown with the stars. We display the CINT and migration images in the top row, and the $l_1$ and direct $l_1$ optimization results in the bottom row.} \label{fig:1} \end{figure} We begin with the results in Figure \ref{fig:1}, for two sources that are at about $2 \lambda_o L/X$ apart. The exact source locations are indicated on the abscissa in the plots, where the units are in $\lambda_o L/X$. We display in the top row the CINT and migration images, and in the bottom row the solutions of the $l_1$ optimization and the direct $l_1$ optimization. Both migration and direct $l_1$ optimization give spurious peaks, due to the random medium. To confirm this, we show in Figure \ref{fig:2} the results of the direct $l_1$ optimization for the same sources in the homogeneous medium, where the reconstruction is very good. The CINT image shown in the top left plot is blurry, and it cannot distinguish the two sources. The $l_1$ optimization improves the result, although there is a small shift in the estimate of the source locations. This shift changes from one realization to another, and it is due to the small random fluctuations of the CINT image. \begin{figure}[h] \begin{centering} \includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_WithOutRandom_L1D} \par\end{centering} \vspace{-0.2in} \caption{Result of direct $l_1$ optimization in the homogeneous medium. The abscissa is the cross-range in units $\lambda_o L/X$, and the source locations are shown with the stars. } \label{fig:2} \end{figure} To illustrate the robustness of the methods to different realizations of the random medium, we display in Figure \ref{fig:3} the histograms of the number of peaks found by each method at a given cross-range location. We define filtered peaks as local maxima whose values are above $33\%$ of the maximum of the image. The height of the histograms varies among the plots in Figure \ref{fig:3} because each method finds a different number of peaks. On the average, the migration images have $9.5$ peaks, the direct $l_1$ method finds $9.65$ peaks, the CINT image has $1.01$ peaks and the $l_1$ optimization finds $2.04$ peaks. While both migration and direct $l_1$ find many spurious peaks, that are far from the source locations, the $l_1$ optimization separates well the two sources and almost always peaks at their true locations. \begin{figure}[h] \begin{centering} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_hist_CINT}}} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_hist_KM}}} \vspace{-0.16in} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_hist_L1}}} {{\includegraphics[width=0.35\textwidth]{Figures/Example_RandomEffect_hist_L1D}}} \par\end{centering} \vspace{-0.2in} \caption{Histogram of the number of peaks obtained in $500$ realizations of the random medium. The abscissa is the cross-range in units $\lambda_o L/X$, and the source locations are shown with the stars. We display the results for the CINT and migration images in the top row, and the $l_1$ and direct $l_1$ optimization results in the bottom row. The heights of the histograms are different because each method finds a different number of peaks. On the average, the number of peaks found in simulations are\;\; Migration:\ 9.5,\;\;\; CINT: 1.01,\;\;\; $l_1$ optimization:\ 2.04,\;\;\; Direct $l_1$ optimization:\ 9.65. } \label{fig:3} \end{figure} \begin{figure}[h] \begin{centering} \includegraphics[width=0.35\textwidth]{Figures/Example_PixelSize_Resol_pix1ov1_rad1ov4_L1}\includegraphics[width=0.35\textwidth]{Figures/Example_PixelSize_Resol_pix1ov2_rad1ov4_L1} \vspace{-0.15in} \includegraphics[width=0.35\textwidth]{Figures/Example_PixelSize_Resol_pix1ov4_rad1ov4_L1}\includegraphics[width=0.35\textwidth]{Figures/Example_PixelSize_Resol_pix1ov8_rad1ov4_L1} \par\end{centering} \vspace{-0.2in} \caption{We display $l_1$ optimization results in one realization of the random medium, for $5$ sources with locations indicated by the stars in the abscissa. The units of the abscissa are in $\lambda_o L/X$. The dark thin bars show the reconstruction and the light gray bars give the aggregated values of the reconstruction in the intervals of length $r = L\lambda_{o}/(4X)$, centered at the source locations. The mesh size is, clockwise, starting from the top left, $H = 1, 1/2, 1/8$ and $1/4$ of $\lambda_o L/X$. The results improve as we refine the mesh.} \label{fig:4} \end{figure} We display in Figure \ref{fig:4} the effect of the mesh size $H$ on the $l_1$ optimization results. Because we have $5$ sources in this simulation, that are closer apart than $\lambda_o L/X$, we do not expect a nearly exact reconstruction with the $l_1$ optimization. Thus, we display in addition to the actual reconstructions the aggregated values recovered in the intervals of length $r = \lambda_o L/(4X),$ centered at the true source locations. We observe that the results improve as we decrease the mesh size from $H = \lambda_o L/X$ to $\lambda_o L/(8X)$. This is due to the fact that the sources are off the grid, and the discretization error decreases as we reduce $H$. \begin{figure}[h] \centering \includegraphics[scale=0.41]{Figures/Example_Resol_resol075_L1}\includegraphics[scale=0.41]{Figures/Example_Resol_resol100_L1} \includegraphics[scale=0.41]{Figures/Example_Resol_resol200_L1} \vspace{-0.2in} \caption{We display $l_1$ optimization results in one realization of the random medium, for $2$ sources separated by, from left to right, $3/4$, $1$ and $2$ $\lambda_o L/X$. The units of the abscissa are in $\lambda_o L/X$.} \label{fig:5} \end{figure} The last illustration, in Figure \ref{fig:5}, shows the effect of the source separation on the quality of the reconstructions. As expected from the results in section \ref{sect:res}, the reconstruction is better when the sources are further apart. \section{Summary} \label{sect:sum} We studied receiver array imaging of remote localized sources in random media, using convex optimization. The scattering regime is defined by precise scaling assumptions, and leads to large random wavefront distortions of the waves measured at the array. Conventional imaging methods like reverse time migration, also known as backprojection \cite{Biondi,munson1983tomographic}, or standard $l_1$ optimization \cite{chai2013robust,fannjiang2010compressed}, cannot deal with such distortions and produce poor and unreliable results. We base our imaging on the coherent interferometric (CINT) method \cite{borcea2006adaptive} which mitigates random media effects like wavefront distortions at the expense of image resolution. The goal of the convex optimization is to remove the blur in the CINT images and thus improve the source localization. We show with a detailed analysis that under generic conditions the CINT imaging function is approximately a convolution of a blurring kernel with the discretized unknown source intensity on the imaging mesh. The kernel has a generic expression, that depends on the known CINT resolution limits obtained in \cite{borcea2006adaptive,borcea2007asymptotics,borcea2011enhanced} for various wave propagation models. The optimization seeks to undo this convolution. The analysis and numerical simulations show that it gives very good estimates of the source locations when they are sufficiently far apart. This is in agreement with the results in \cite{candes2014towards}. We also show that when the sources are clustered together, the estimates are not close to the true locations pointwise, but they are supported in their vicinity. \section*{Acknowledgments} We gratefully acknowledge support from ONR Grant N00014-14-1-0077 and the AFOSR Grant FA9550-15-1-0118.
ee5bcf1d089f1adb0bc52970c0adc6c15ca43491
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In various papers \cite{BV}, \cite{MaM}, \cite{MP4} the Ray-Singer analytic torsion \cite{RS} has been used to study the growth of torsion in the cohomology of cocompact arithmetic groups. Since many important arithmetic groups are not cocompact, it is very desirable to extend these results to the noncompact case. There exist some results for hyperbolic 3-manifolds. In \cite{PR}, Pfaff and Raimbault obtained upper and lower bounds for the growth of torsion in the cohomology of congruence subgroups of Bianchi groups if the local system varies. In \cite{Ra1}, \cite{Ra2}, J. Raimbault has studied the case of sequences $(\Gamma_i)$ of congruence subgroups of Bianchi groups such that $\operatorname{vol}(\Gamma_i\bs\bH^3)\to\infty$ as $i\to\infty$. The approach in the cocompact case relies on the equality of analytic torsion and Reidemeister torsion of the corresponding locally symmetric manifolds. We briefly recall the definition of the Ray-Singer analytic torsion. Let $X$ be a compact Riemannian manifold of dimension $n$ and $\rho\colon \pi_1(X)\to\operatorname{GL}(V)$ a finite dimensional representation of its fundamental group. Let $E_\rho\to X$ be the flat vector bundle associated with $\rho$. Choose a Hermitian fiber metric in $E_\rho$. Let $\Delta_p(\rho)$ be the Laplace operator on $E_\rho$-valued $p$-forms with respect to the metrics on $X$ and in $E_\rho$. It is an elliptic differential operator, which is formally self-adjoint and non-negative. Let $h_p(\rho):=\dim\ker\Delta_p(\rho)$. Using the trace of the heat operator $e^{-t\Delta_p(\rho)}$, the zeta function $\zeta_p(s;\rho)$ of $\Delta_p(\rho)$ can be defined by \begin{equation}\label{zeta-fct} \zeta_p(s;\rho):= \frac{1}{\Gamma(s)}\int_0^\infty \left(\operatorname{Tr}\left(e^{-t\Delta_p(\rho)}\right)-h_p(\rho)\right)t^{s-1} \;dt. \end{equation} The integral converges for $\Re(s)>n/2$ and admits a meromorphic extension to the whole complex plane, which is holomorphic at $s=0$. Then the Ray-Singer analytic torsion $T_X(\rho)\in\R^+$ is defined by \begin{equation}\label{analtor0} \log T_X(\rho)=\frac{1}{2}\sum_{p=1}^d (-1)^p p \frac{d}{ds}\zeta_p(s;\rho)\big|_{s=0}. \end{equation} The analytic torsion has a topological counterpart. This is the Reidemeister torsion $\tau_X(\rho)$, which is defined in terms of a smooth triangulation of $X$ \cite{RS}, \cite{Mu5}. It is known that for unimodular representations $\rho$ (meaning that $|\det\rho(\gamma)|=1$ for all $\gamma\in\pi_1(X)$) one has the equality $T_X(\rho)=\tau_X(\rho)$ \cite{Ch}, \cite{Mu4}, \cite{Mu5}. In the general case of a non-unimodular representation the equality does not hold, but the defect can be described \cite{BZ}. This equality has the following interesting consequence. Assume that the space of the representation $\rho$ contains a lattice which is invariant under $\pi_1(X)$. Let $\cM$ be the associated local system of free $\Z$-modules. Let $H^p(X,\cM)_{\tors}$ be the torsion subgroup of $H^p(X,\cM)$. Then \begin{equation}\label{tor-coh} T_X(\rho)= R\cdot\prod_{p=0}^d |H^p(X,\cM)_{\tors}|^{(-1)^{p+1}}, \end{equation} where $R$ is the so called ``regulator'', defined in terms of the free part of the cohomology $H^p(X,\cM)$ (see \cite{BV}, \cite{MP4}). In particular, if $\rho$ is acyclic, i.e., $H^\ast(X,E_\rho)=0$, then $R=1$. The equality \eqref{tor-coh} is the starting point for the application of the analytic torsion to the study of the torsion in the cohomology of cocompact arithmetic groups. The definition of the analytic torsion \eqref{analtor0} obviously depends on the compactness of the underlying manifold. Without this assumption, the heat operator $e^{-t\Delta_p(\rho)}$ is, in general, not a trace class operator. If one attempts to generalize the above method to non-cocompact arithmetic groups, the first problem is to define an appropriate regularized trace of the heat operators. For hyperbolic manifolds of finite volume one can proceed as in Melrose \cite{Me} to define the regularized trace by means of the renormalized trace of the heat kernel. This method has been used in \cite{CV}, \cite{PR}, \cite{MP1}, \cite{MP3}, \cite{MP4}. One uses an appropriate height function to truncate the hyperbolic manifold $X$ at height $T>0$. This amounts to cut off the cusps at sufficiently high level $T>T_0$. Then one integrates the point wise trace of the heat kernel over the truncated manifold $X(T)$. This integral has an asymptotic expansion in $\log T$. The constant term is defined to be the renormalized trace of the heat operator. The purpose of the present paper is to start the investigation of the case of finite volume locally symmetric spaces of any rank by defining a regularized analytic torsion for arithmetic quotients associated to split forms of type $A_n$ over $\Q$. In the higher rank case we proceed in the same way as in the case of hyperbolic manifolds. The first problem is to define the truncation in the right way. For this we can build on Arthur's work. The definition of the truncation operator is an important issue in Arthur's trace formula \cite{Ar1}, which we will use for our purpose. To this end we need to switch to the adelic framework. Now we will describe the approach in more detail. For simplicity assume that $G$ is a connected semisimple algebraic group defined over $\Q$. Assume that $G(\R)$ is not compact. Let $K_\infty$ be a maximal compact subgroup of $G(\R)$. Put $\widetilde X=G(\R)/K_\infty$. Let $\A$ be the ring of adeles of $\Q$ and $\A_f$ the ring of finite adeles. Let $K_f\subset G(\A_f)$ be an open compact subgroup. We consider the adelic quotient \begin{equation} X(K_f)=G(\Q)\bs (\widetilde X\times G(\A_f))/K_f. \end{equation} This is the adelic version of a locally symmetric space. In fact, $X(K_f)$ is the disjoint union of finitely many locally symmetric spaces $\Gamma_i\bs\widetilde X$, $i=1,\dots,l$, (see section \ref{sec-arithm-mfd}). If $G$ is simply connected, then by strong approximation we \[ X(K_f)=\Gamma\bs \widetilde X, \] where $\Gamma=(G(\R)\times K_f)\cap G(\Q)$. We will assume that $K_f$ is neat, that is, the eigenvalues of any element in $\Gamma$ generate a torsion free subgroup in $\C^\times$, so that $X(K_f)$ is a manifold. Let $\nu\colon K_\infty\to \operatorname{GL}(V_\nu)$ be a finite dimensional unitary representation. It induces a homogeneous Hermitian vector bundle $\widetilde E_\nu$ over $\widetilde X$, which is equipped with the canonical connection $\nabla^\nu$. Being homogeneous, $\widetilde E_\nu$ can be pushed down to a locally homogeneous Hermitian vector bundle over each component $\Gamma_i\bs \widetilde X$ of $X(K_f)$. Their disjoint union is a Hermitian vector bundle $E_\nu$ over $X(K_f)$. Let $\widetilde\Delta_\nu$ (resp. $\Delta_\nu$) be the associated Bochner-Laplace operator acting in the space of smooth section of $\widetilde E_\nu$ (resp. $E_\nu$). Let $e^{-t\widetilde\Delta_\nu}$ (resp. $e^{-t\Delta_\nu}$), $t>0$, be the heat semigroup generated by $\widetilde\Delta_\nu$ (resp. $\Delta_\nu$). Since $\widetilde \Delta_\nu$ commutes with the action of $G(\R)$, it follows that $e^{-t\widetilde\Delta_\nu}$ is a convolution operator with kernel given by a smooth map $H_t^\nu\colon G(\R)\to \operatorname{End}(V_\nu)$. Let $h_t^\nu(g)=\operatorname{tr} H_t^\nu(g)$, $g\in G(\R)$. In fact, $h_t^\nu$ belongs to Harish-Chandra's Schwartz space $\Co(G(\R))$. Let $\chi_{K_f}$ be the characteristic function of $K_f$ in $G(\A_f)$. We define the function $\phi_t^\nu\in C^\infty(G(\A))$ by \[ \phi_t^\nu(g_\infty g_f)=h_t^\nu(g_\infty)\chi_{K_f}(g_f),\quad g_\infty\in G(\R),\; g_f\in G(\A_f). \] In fact, $\phi_t^\nu$ belongs to $\Co(G(\A);K_f)$, the adelic version of the Schwartz space (see section \ref{sec-trunc} for its definition). If $X(K_f)$ is compact, then one has \begin{equation}\label{trcocom} \operatorname{Tr}\left(e^{-t\Delta_\nu}\right)=\int_{G(\Q)\bs G(\A)}\sum_{\gamma\in G(\Q)} \phi_t^\nu(x^{-1}\gamma x)\;dx. \end{equation} This is our starting point for defining the renormalized trace in the noncompact case. We fix a minimal Levi subgroup $M_0$ of $G$. If $M\subseteq G$ is a Levi subgroup containing $M_0$, let $A_M$ be the split component of the center of $M$. Let $\af_0:=\af_{M_0}$ be the Lie algebra of $A_{M_0}(\R)$. Let $J_{\geo}$ be the geometric side of the Arthur trace formula introduced in \cite{Ar1}; see also \cite{Ar} for an introduction to the trace formula. For $f\in C_c^\infty(G(\A))$, Arthur defines $J_{\geo}(f)$ as the value at a point $T_0\in\af_0$, specified in \cite[Lemma 1.1]{Ar3}, of a polynomial $J^T(f)$ on $\af_0$. In fact, by \cite[Theorem 7.1]{FL1}, $J^T(f)$ is defined for all $f\in\Co(G(\A);K_f)$. Furthermore, we use an appropriate height function to truncate $G(\A)$. For $T\in\af_0$ let $G(\A)_{\le T}$ be obtained by truncating $G(\A)$ at level $T$ (see \eqref{trunc1}). This is a compact subset of $G(\A)$. By \cite[Theorem 7.1]{FL1} it follows that for sufficiently regular $T\in\af_0$ we have \begin{equation}\label{truncker} \int_{G(\Q)\bs G(\A)_{\le T}}\sum_{\gamma\in G(\Q)} \phi_t^\nu(x^{-1}\gamma x)\;dx = J^T(\phi_t^\nu)+O\left(e^{-c\|T\|}\right). \end{equation} Since $J^T(\phi_t^\nu)$ is a polynomial in $T$, we get an asymptotic expansion in $T$ of the truncated integral. Under additional assumption on $G$, which are satisfied for $\operatorname{GL}(n)$ and $\operatorname{SL}(n)$, the point $T_0\in \af_0$, determined by \cite[Lemma 1.1]{Ar1}, is equal to 0. Thus in this case $J_{\geo}(\phi_t^\nu)$ is the constant term of the polynomial $J^T(\phi_t^\nu)$. This leads to our definition of the regularized trace \begin{equation} \operatorname{Tr}_{\reg}\left(e^{-t\Delta_\nu}\right):=J_{\geo}(\phi_t^\nu). \end{equation} In general, $J_{\geo}(\phi_t^\nu)$ is not the constant term of the polynomial $J^T(\phi_t^\nu)$. Nevertheless, we prefer this definition, because of its independence on the choice of the minimal parabolic subgroup $P_0$. The next goal is to determine the asymptotic behavior of $\operatorname{Tr}_{\reg}\left(e^{-t\Delta_\nu}\right)$ as $t\to 0$ and $t\to\infty$, respectively. To this end we use the Arthur trace formula. Currently we are only able to deal with these problems for the groups $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$. For $N\in\N$ let $K(N)\subset G(\A_f)$ be the principal congruence subgroup of level $N$. Recall that $K(N)$ is neat for $N\ge 3$. Our first main result is the following proposition. \begin{theo}\label{prop-asymp3} Let $G=\operatorname{GL}(n)$ or $\operatorname{SL}(n)$. Let $K_f\subset G(\A_f)$ be an open compact subgroup. Assume that $K_f$ is contained in $K(N)$ for some $N\ge 3$. Let $\nu$ be finite dimensional unitary representation of $K_\infty$ and let $\Delta_\nu$ be the associated Bochner-Laplace operator. Let $d=\dim X(K_f)$. As $t\to +0$, there is an asymptotic expansion \begin{equation}\label{asex} \operatorname{Tr}_{\reg}\left(e^{-t\Delta_\nu}\right)\sim t^{-d/2}\sum_{j=0}^\infty a_j(\nu) t^j +t^{-(d-1)/2}\sum_{j=0}^\infty\sum_{i=0}^{r_j} b_{ij}(\nu)t^{j/2}(\log t)^i. \end{equation} Moreover $r_j\le n-1$ for all $j\in\N_0$. \end{theo} For hyperbolic manifolds a similar result was proved in \cite{Mu6}. To study the large time behavior we restrict attention to twisted Laplace operators, which are relevant for studying the analytic torsion with coefficients in local systems. Let $\tau\colon G(\R)\to \operatorname{GL}(V_\tau)$ be a finite dimensional complex representation. Let $\Gamma_i\bs \widetilde X$, $i=1,\dots,l$, be the components of $X(K_f)$. The restriction of $\tau$ to $\Gamma_i$ induces a flat vector bundle $E_{\tau,i}$ over $\Gamma_i\bs \widetilde X$. The disjoint union is a flat vector bundle $E_\tau$ over $X(K_f)$. By \cite{MM} it is isomorphic to the locally homogeneous vector bundle associated to $\tau|_{K_\infty}$. It can be equipped with a fiber metric induced from the homogeneous bundle. Let $\Delta_p(\tau)$ be the corresponding twisted Laplace operator on $p$-forms with values in $E_\tau$. Let $\operatorname{Ad}_{\pf}\colon K_\infty\to \operatorname{GL}(\pf)$ be the adjoint representation of $K_\infty$ on $\pf$, where $\pf=\kf^\perp$, and $\nu_p(\tau)=\Lambda^p\operatorname{Ad}_{\pf}^\ast\otimes\tau$. Up to a vector bundle endomorphism, $\Delta_p(\tau)$ equals the Bochner-Laplace operator $\Delta_{\nu_p(\tau)}$. So $\operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)$ is well defined. Let $\theta$ be the Cartan involution of $G(\R)$ with respect to $K_\infty$. Put $\tau_\theta:= \tau\circ\theta$. The large time behavior of the regularized trace is described by the following proposition. \begin{theo}\label{theo-lt} Let $G=\operatorname{GL}(n)$ or $\operatorname{SL}(n)$. Let $K_f\subset G(\A_f)$ be an open compact subgroup which is contained in $K(N)$ for some $N\ge 3$. Let $\tau$ be finite dimensional representation of $G(\R)$. Assume that $\tau\not\cong\tau_\theta$. Then we have \begin{equation}\label{largetime} \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)=O(e^{-ct}) \end{equation} as $t\to\infty$ for all $p=0,\dots,d$. \end{theo} The proof is an immediate consequence of Proposition \ref{asympinf} together with the trace formula. Without the assumption $\tau\not\cong\tau_\theta$ the behavior of $\operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)$ as $t\to\infty$ is more complicated and it is definitely not exponentially decreasing. This condition is also relevant in \cite{BV}. It implies that the representation $\tau$ is strongly acyclic \cite[Lemma 4.1]{BV}, which is a necessary condition to establish the main results of \cite{BV}. It is a very challenging problem to eliminate this condition. We also note that the condition $\tau\not\cong\tau_\theta$ implies the vanishing theorem of Borel-Wallach for the cohomology of a cocompact lattice in a semisimple Lie group \cite[Theorem 6.7, Ch. VII]{BW}. By Theorems \ref{prop-asymp3} and \ref{theo-lt} we can define the zeta function of $\Delta_p(\tau)$ as in \eqref{zetafct}, using the regularized trace of $e^{-t\Delta_p(\tau)}$ in place of the usual trace. The corresponding Mellin transform converges absolutely and uniformly on compact subsets of the half-plane $\Re(s)>d/2$ and admits a meromorphic extension to the whole complex plane. Because of the presence of the log-terms in the expansion \eqref{asex}, the zeta function may have a pole at $s=0$. Let $f(s)$ be a meromorphic function on $\C$. For $s_0\in\C$ let $f(s)=\sum_{k\ge k_0}a_k(s-s_0)^k$ be the Laurent expansion of $f$ at $s_0$. Put $\operatorname{FP}_{s=s_0}f(s):=a_0$. Now we define the analytic torsion $T_{X(K_f)}(\tau)\in\C\setminus\{0\}$ by \begin{equation}\label{analtor} \log T_{X(K_f)}(\tau)=\frac{1}{2}\sum_{p=0}^d (-1)^p p \left(\operatorname{FP}_{s=0}\frac{\zeta_p(s;\tau)}{s}\right). \end{equation} In the case of $G=\operatorname{GL}(3)$ we are able to determine the coefficients of the log-terms. This shows that the zeta functions definitely have a pole at $s=0$. However, the combination $\sum_{p=1}^5 (-1)^p p \zeta_p(s;\tau)$ turns out to be holomorphic at $s=0$ and we can define the logarithm of the analytic torsion by \[ \log T_{X(K_f)}(\tau)=\frac{d}{ds} \left(\frac{1}{2}\sum_{p=1}^5 (-1)^p p \zeta_p(s;\tau)\right)\bigg|_{s=0}. \] Let $\{K_f(N)\}_{N\in\N}$ be the family of principal congruence subgroups of $\operatorname{GL}(n,\A_f)$, and $X(N):=X(K_f(N))$, $N\in\N$. The next problem is to study the limiting behavior of $\log T_{X(N)}(\tau)/\operatorname{vol}(X(N))$ as $N\to\infty$ which we do in subsequent work. In consideration of the results for the cocompact case in \cite{BV}, one can expect a different behavior of $\log T_{X(N)}(\tau)/\operatorname{vol}(X(N))$ in the limit $N\to\infty$ for different $n$. More precisely, the fundamental rank $\operatorname{rank} G(\R)- \operatorname{rank} K_\infty$ determined in \cite{BV} whether the limit vanishes, which it does unless if the rank equals $1$. In our case of $\operatorname{SL}_n(\R)$, the fundamental rank is $1$ precisely when $n=3$ or $n=4$. An even more difficult problem is the question if there is a combinatorial counterpart of $T_{X(K_f)}(\tau)$ as there is in the compact case. Now we briefly explain our method to prove Theorems \ref{prop-asymp3} and \ref{theo-lt}. To determine the asymptotic behavior of the regularized trace as $t\to +0$, we use the geometric side of trace formula. The first step is to show that $\phi_t^\nu$ can be replaced by a compactly supported function $\widetilde \phi_t^\nu\in C_c^\infty(G(\A))$ without changing the asymptotic behavior. Next we use the coarse geometric expansion of the geometric side, which expresses $J_{\geo}(f)$, $f\in C_c^\infty(G(\A))$, as a sum of distributions $J_{\of}(f)$ associated to semisimple conjugacy classes of $G(\Q)$. Let $J_{\operatorname{unip}}(f)$ be the distribution associated to the class of $1$. If the support of $\widetilde \phi_t^\nu$ is a sufficiently small neighborhood of $1$, it follows that \begin{equation} \operatorname{Tr}_{\reg}\left(e^{-t\Delta_\nu}\right):=J_{\operatorname{unip}}(\widetilde\phi_t^\nu)+ O\left(e^{-c/t}\right) \end{equation} as $t\to +0$. To analyze $J_{\operatorname{unip}}(\widetilde\phi_t^\nu)$, we use the fine geometric expansion \cite{Ar4} which expresses $J_{\operatorname{unip}}(\widetilde\phi_t^\nu)$ in terms of weighted orbital integrals. If the real rank of $G(\R)$ is one, the weighted orbital integrals are rather simple and the weight factors are explicitly known (see \cite{Wa}). In order to deal with the weighted orbital integrals in the higher rank case, we need to restrict to the groups $\operatorname{GL}(n)$ or $\operatorname{SL}(n)$. In this case all unipotent orbits are Richardson, which simplifies the analysis considerably. We are only interested in the situation over the field $\R$. Let $M$ be a Levi subgroup of $G$. Let $\cU_M$ be the unipotent variety in $M$ and ${\mathcal V}\in (\cU_M)$ a conjugacy class. Let $U$ be an $M(\R)$ conjugacy class in ${\mathcal V}(\R)$. There exists a standard parabolic subgroup $Q=LN\in\cF$ and a constant $c>0$ such that for every $\operatorname{O}(n)$-conjugation invariant function $f\in C_c^\infty(G(\R))$ the weighted orbital integral $J_M(U,f)$ is given by \begin{equation}\label{orbint0} J_M(U,f)=c\int_{N(\R)}f(n) w_{M,{\mathcal V}}(n)\;dn, \end{equation} where $w_{M,{\mathcal V}}(n)$ is a certain weight function. The main problem is now to determine the structure of the weight function. For $G(\R)=\operatorname{SO}_0(n,1)$ the weighted orbital integral is of the same form with weight function $w(n)=\log\|\log n\|$, where the inner log is the isomorphism $\log \colon N\to\nf$. This fact has been exploited in \cite{Mu6} in order to establish the asymptotic expansion of the regularized trace in the case of hyperbolic manifolds of finite volume. It turns out that $w_{M,{\mathcal V}}$ has a similar behavior with respect to scaling. Note that the map $x\mapsto X=x-\operatorname{id}$ defines a bijection between the variety of unipotent elements in $G(\R)$ and the nilpotent cone in the Lie algebra $\mathfrak{g}(\R)$. For $s\in\R$ let $x_s:=\operatorname{id}+s(x-\operatorname{id})$. Let $x\in{\mathcal V}^G(\R)$ such that $w_{M,{\mathcal V}}(x)$ is defined. Then by Proposition \ref{prop}, $w_{M,{\mathcal V}}(x_s)$ is well-defined for every $s>0$ and $s\mapsto w_{M,{\mathcal V}}(x_s)$ is a polynomial in $\log s$ of degree at most $\dim\af_M^G$. Inserting a standard parametrix for the heat kernel into \eqref{orbint0} and using the structure of $w_{M,{\mathcal V}}$, we obtain Theorem \ref{prop-asymp3}. To eliminate the assumption that $K_f$ is contained in some $K(N)$ with $N\ge 3$, we would have to consider orbital integrals associated to classes of finite order. For $\operatorname{GL}(2)$ and $\operatorname{GL}(3)$ we discuss this issue in section \ref{sec-finite-ord}. To prove Theorem \ref{theo-lt}, we use the spectral side of the trace formula. Let $\phi_t^{\tau,p}$ be the function in $\Co(G(\A);K_f)$, which is defined in the same way as $\phi_t^\nu$ in terms of the kernel of the heat operator on the universal covering. Then by the trace formula \[ \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)=J_{\spec}(\phi^{\tau,p}_t). \] The key input to deal with the spectral side is the refinement of the spectral expansion of the Arthur trace formula established in \cite{FLM1} (see Theorem \ref{thm-specexpand}). For $f\in\Co(G(\A))$ we have \[ J_{\spec}(f)=\sum_{[M]}J_{\spec,M}(f), \] where $[M]$ runs over the conjugacy classes of Levi subgroups of $G$ and $J_{\spec,M}(f)$ is a distribution associated to $M$. The distribution associated to $G$ is $\operatorname{Tr} R_{\di}(f)$, where $R_{\di}$ denotes the restriction of the regular representation of $G(\A)$ in $L^2(G(\Q)\bs G(\A))$ to the discrete subspace. For a proper Levi subgroup $M$ of $G$, $J_{\spec,M}(f)$ is an integral whose main ingredient are logarithmic derivatives of intertwining operators. Using our assumption that $\tau\neq\tau_\theta$, we obtain $\dim\ker\Delta_p(\tau)=0$. Then it follows as in the compact case that there exists $c>0$ such that \[ \operatorname{Tr} R_{\di}(\phi^{\tau,p}_t)=O(e^{-ct}), \quad\text{as}\;\; t\to\infty. \] For a proper Levi subgroup $M$, the determination of the asymptotic behavior of $J_{\spec,M}(\phi^{\tau,p}_t)$ as $t\to\infty$ relies on two conjectural properties, one global and one local, of the intertwining operators. The global property is a uniform estimate on the winding number of the normalizing factors of the intertwining operators in the co-rank one case. For $\operatorname{GL}(n)$ and $\operatorname{SL}(n)$, this property follows from known, but delicate, properties of the Rankin-Selberg $L$-functions \cite{FLM2}. The local property is concerned with the estimation of logarithmic derivatives of normalized local intertwining operators, which are uniform in $\pi$. For $\operatorname{GL}(n)$ the pertinent estimates have been established in \cite[Proposition 0.2]{MS}. They are a consequence of a weak version of the Ramanujan conjecture. The case of $\operatorname{SL}(n)$ can be reduced to $\operatorname{GL}(n)$ in the same way as in the proof of \cite[Lemma 5.14]{FLM2}. Let $\theta\colon G\to G$ be the Cartan involution and let $\tau_\theta:=\tau\circ \theta$. Using these estimations, it follows that for $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$, a proper Levi subgroup $M$ of $G$ and a finite dimensional representation $\tau$ of $G(\R)$ such that $\tau\not\cong\tau_\theta$, one has $J_{\spec,M}(\phi^{\tau,p}_t)= O(e^{-ct})$ as $t\to\infty$. Putting everything together, we obtain Theorem \ref{theo-lt}. We end this introduction with some remarks on the possible extension of the our results to other groups $G$. First of all, Theorem \ref{theo-lt} depends on the estimations of logarithmic derivatives of global normalizing factors and normalized local intertwining operators. Using functoriality, T. Finis and E. Lapid \cite{FL2} have recently established similar estimates of the logarithmic derivatives of global normalizing factors associated to intertwining operators for the following reductive groups over number fields: inner forms of $\operatorname{GL}(n)$, quasi-split classical groups and their similitude groups, and the exceptional groups $G_2$. One can expect that the estimates of the logarithmic derivatives of the normalized local intertwining operators can be established by the same methods. This would lead to an extension of Theorem \ref{theo-lt} to these groups. It remains to deal with the unipotent orbital integrals for the groups above. The paper is organized as follows. In section \ref{sec-prelim} we fix notations and recall some basic facts. In section \ref{sec-arithm-mfd} we introduce the locally symmetric manifolds as adelic quotients. In section \ref{sec-trunc} we compare two different methods of truncation. One of them is based on the truncation of kernels of integral operators which leads to the geometric side of the trace formula. The other one consists in the truncation of the underlying manifold, which is the basis for the renormalization of the trace of the heat operator. In section \ref{sec-trform} we recall the spectral side of the Arthur trace formula. In section \ref{sec-unipotent} we are assuming that $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$. We discuss the unipotent contribution to the trace formula and derive a simplified formula for the weighted orbital integral. Section \ref{sec-wfct} is devoted to the study of the weight functions for the groups $\operatorname{GL}(n)$ and $\operatorname{SL}(n)$. The main result is Proposition \ref{prop}, which is the key result that enables us to determine the asymptotic behavior as $t\to +0$ of the corresponding orbital integrals. Examples of low rank are discussed in section \ref{sec-lowrank}. These are cases where the weight function is given explicitly. In section \ref{sec-bochlapl} we collect some basic facts concerning Bochner-Laplace operators. The regularized trace of the corresponding heat operators is introduced in section \ref{sec-regtr}. The definition is based on section \ref{sec-trunc}, which deals with truncation. In section \ref{sec-heatkernel} we establish some estimates of the heat kernel for Bochner-Laplace operators on the symmetric space $\widetilde X$. Combined with the analysis of the weight functions in section \eqref{sec-wfct}, the estimations are used in section \eqref{sec-asymp} to prove Theorem \ref{prop-asymp3}. In section \eqref{sec-analtor} we first use the spectral side of the Arthur trace formula to establish Theorem \ref{theo-lt}, which concerns the large time asymptotic behavior of the regularized trace of the heat operators. This finally enables us to define the regularized analytic torsion. In section \ref{sect-gl3} we assume that $G=\operatorname{GL}(3)$. Using the explicit form of the weight functions described in section \ref{sec-lowrank}, we determine the coefficients of the possible poles at $s=0$ of the zeta functions. It turns out that the combination of the zeta functions, which is used to define the analytic torsion, is holomorphic at $s=0$. In the final section \ref{sec-finite-ord} we consider for $G=\operatorname{GL}(2)$ or $G=\operatorname{GL}(3)$ an arbitrary subgroup $K_f$ of $G(\hat\Z)$ and study the additional weighted orbital integrals that arise in this case. \noindent {\bf Acknowledgment.} The authors would like to thank the referees for the careful reading of the manuscript and for their very helpful suggestions and comments. \section{Preliminaries}\label{sec-prelim} Let $G$ be a reductive algebraic group defined over $\Q$. We fix a minimal parabolic subgroup $P_0$ of $G$ defined over $\Q$ and a Levi decomposition $P_0=M_0\cdot N_0$, both defined over $\Q$. Let $\cF$ be the set of parabolic subgroups of $G$ which contain $M_0$ and are defined over $\Q$. Let $\cL$ be the set of subgroups of $G$ which contain $M_0$ and are Levi components of groups in $\cF$. For any $P\in\cF$ we write \[ P=M_PN_P, \] where $N_P$ is the unipotent radical of $P$ and $M_P$ belongs to $\cL$. Let $M\in\cL$. Denote by $A_M$ the $\Q$-split component of the center of $M$. Put $A_P=A_{M_P}$. Let $L\in\cL$ and assume that $L$ contains $M$. Then $L$ is a reductive group defined over $\Q$ and $M$ is a Levi subgroup of $L$. We shall denote the set of Levi subgroups of $L$ which contain $M$ by $\cL^L(M)$. We also write $\cF^L(M)$ for the set of parabolic subgroups of $L$, defined over $\Q$, which contain $M$, and $\cP^L(M)$ for the set of groups in $\cF^L(M)$ for which $M$ is a Levi component. Each of these three sets is finite. If $L=G$, we shall usually denote these sets by $\cL(M)$, $\cF(M)$ and $\cP(M)$. Let $X(M)_\Q$ be the group of characters of $M$ which are defined over $\Q$. Put \begin{equation}\label{liealg} \af_{M}:=\operatorname{Hom}(X(M)_\Q,\R). \end{equation} This is a real vector space whose dimension equals that of $A_M$. Its dual space is \[ \af_{M}^\ast=X(M)_\Q\otimes \R. \] We shall write, \begin{equation}\label{liealg1} \af_P=\af_{M_P},\;A_0=A_{M_0}\quad\text{and}\quad \af_0=\af_{M_0}. \end{equation} For $M\in\cL$ let $A_M(\R)^0$ be the connected component of the identity of the group $A_M(\R)$. Let $W_0=N_{\G(\Q)}(A_0)/M_0$ be the Weyl group of $(G,A_0)$, where $N_{G(\Q)}(H)$ is the normalizer of $H$ in $G(\Q)$. For any $s\in W_0$ we choose a representative $w_s\in G(\Q)$. Note that $W_0$ acts on ${\mathcal L}$ by $sM=w_s M w_s^{-1}$. For $M\in\cL$ let $W(M)=N_{\G(\Q)}(M)/M$, which can be identified with a subgroup of $W_0$. For any $L\in\cL(M)$ we identify $\af_L^\ast$ with a subspace of $\af_M^\ast$. We denote by $\af_M^L$ the annihilator of $\af_L^\ast$ in $\af_M$. We set \[ {\mathcal L}_1(M)=\{L\in{\mathcal L}(M):\dim\mathfrak{a}_M^L=1\} \] and \begin{equation}\label{f1} \cF_1(M)=\bigcup_{L\in{\mathcal L}_1(M)}\cP(L). \end{equation} We shall denote the simple roots of $(P,A_P)$ by $\Delta_P$. They are elements of $X(A_P)_\Q$ and are canonically embedded in $\af_P^\ast$. Let $\Sigma_P\subset \af_P^\ast$ be the set of reduced roots of $A_P$ on the Lie algebra of $G$. The set $\Delta_0=\Delta_{P_0}$ is a base for a root system. In particular, for every $\alpha\in\Delta_P$ we have a co-root $\alpha^\vee\in\af_{P_0}$. Let $P_1$ and $P_2$ be parabolic subgroups with $P_1\subset P_2$. Then $\af_{P_2}^\ast$ is embedded into $\af_{P_1}^\ast$, while $\af_{P_2}$ is a natural quotient vector space of $\af_{P_1}$. The group $M_{P_2}\cap P_1$ is a parabolic subgroup of $M_{P_2}$. Let $\Delta_{P_1}^{P_2}$ denote the set of simple roots of $(M_{P_2}\cap P_1,A_{P_1})$. It is a subset of $\Delta_{P_1}$. For a parabolic subgroup $P$ with $P_0\subset P$ we write $\Delta_0^P:=\Delta_{P_0}^P$. Let $\A$ (resp. $\A_f$) be the ring of adeles (resp. finite adeles) of $\Q$. We fix a maximal compact subgroup $\K=\prod_v K_v = K_\infty\cdot \K_{f}$ of $G(\A)=G(\R)\cdot G(\A_{f})$. We assume that the maximal compact subgroup $\K \subset G(\A)$ is admissible with respect to $M_0$ \cite[\S 1]{Ar5}. Let $H_M: M(\A)\rightarrow\mathfrak{a}_M$ be the homomorphism given by \begin{equation}\label{homo-M} e^{\sprod{\chi}{H_M(m)}}=\abs{\chi (m)}_\A = \prod_v\abs{\chi(m_v)}_v \end{equation} for any $\chi\in X(M)$ and denote by $M(\A)^1 \subset M(\A)$ the kernel of $H_M$. Then $M(\A)$ is the direct product of $M(\A)^1$ and $A_M(\R)^0$, the component of $1$ in $A_M(\R)$. By the conditions on $\K$, we have $G(\A)=P(\A) \K$. Hence any $x\in G(\A)$ can be written as \[ nmak,\quad n\in N(\A),\;m\in M(\A)^1,\; a\in A_M(\R)^0,\; k\in\K. \] Define $H_P\colon G(\A)\to\af_P$ by \begin{equation}\label{hp} H_P(x):=H_M(a), \end{equation} where $x=nmak$ as above. Let $\gf$ and $\kf$ denote the Lie algebras of $G(\R)$ and $K_\infty$, respectively. Let $\theta$ be the Cartan involution of $G(\R)$ with respect to $K_\infty$. It induces a Cartan decomposition $\mathfrak{g}= \mathfrak{p} \oplus \mathfrak{k}$. We fix an invariant bi-linear form $B$ on $\mathfrak{g}$ which is positive definite on $\mathfrak{p}$ and negative definite on $\mathfrak{k}$. This choice defines a Casimir operator $\Omega$ on $G(\R)$. Let $\Pi(G(\R))$ denote the set of equivalence classes of irreducible unitary representations of $G(\R)$. We denote the Casimir eigenvalue of any $\pi \in \Pi (G(\R))$ by $\lambda_\pi$. Similarly, we obtain a Casimir operator $\Omega_{K_\infty}$ on $K_\infty$ and write $\lambda_\tau$ for the Casimir eigenvalue of a representation $\tau \in \Pi (K_\infty)$ (cf.~\cite[\S 2.3]{BG}). The form $B$ induces a Euclidean scalar product $(X,Y) = - B (X,\theta(Y))$ on $\mathfrak{g}$ and all its subspaces. For $\tau \in \Pi (K_\infty)$ we define $\norm{\tau}$ as in \cite[\S 2.2]{CD}. Note that the restriction of the scalar product $(\cdot,\cdot)$ on $\gf$ to $\af_0$ gives $\af_0$ the structure of a Euclidean space. In particular, this fixes Haar measures on the spaces $\af_M^L$ and their duals $(\af_M^L)^\ast$. We follow Arthur in the corresponding normalization of Haar measures on the groups $M(A)$ (\cite[\S 1]{Ar1}). Let $L^2_{\operatorname{disc}}(A_M(\R)^0 M(\Q)\bs M(\A))$ be the discrete part of $L^2(A_M(\R)^0 M(\Q)\bs M(\A))$, i.e., the closure of the sum of all irreducible subrepresentations of the regular representation of $M(\A)$. We denote by $\Pi_{\operatorname{disc}}(M(\A))$ the countable set of equivalence classes of irreducible unitary representations of $M(\A)$ which occur in the decomposition of the discrete subspace $L^2_{\operatorname{disc}}(A_M(\R)^0 M(\Q)\bs M(\A))$ into irreducible representations. \section{Arithmetic manifolds}\label{sec-arithm-mfd} Let $G$ be a reductive algebraic group over $\Q$. Let $K_f\subset G(\A_f)$ be an open compact subgroup. The double coset space $A_G(\R)^0 G(\Q)\bs G(\A)/G(\R)K_f$ is known to be finite (see \cite[\S 5]{Bo1}). Let $x_1=1, x_2,\dots,x_l$ be a set of representatives in $G(\A_f)$ of the double cosets. Then the groups \[ \Gamma_i:=\left( G(\R)\times x_i K_f x_i^{-1}\right)\cap G(\Q),\quad 1\le i\le l, \] are arithmetic subgroups of $G(\R)$ and the action of $G(\R)$ on the space of double cosets $A_G(\R)^0 G(\Q)\bs G(\A)/K_f$ induces the following decomposition into $G(\R)$-orbits: \begin{equation}\label{adel-quot1} A_G(\R)^0 G(\Q)\bs G(\A)/K_f\cong \bigsqcup_{i=1}^l \left(\Gamma_i\bs G(\R)^1\right), \end{equation} where $G(\R)^1=G(\R)/A_G(\R)^0$. Thus we get an isomorphism of $G(\R)$-modules \begin{equation}\label{g-modules} L^2(A_G(\R)^0 G(\Q)\bs G(\A))^{K_f}\cong \bigoplus_{i=1}^lL^2(\Gamma_i\bs G(\R)^1). \end{equation} We note that, in general, $l>1$. However, if $G$ is semisimple, simply connected, and without any $\Q$-simple factors $H$ for which $H(\R)$ is compact, then by strong approximation we have \[ G(\Q)\bs G(\A)/K_f\cong \Gamma\bs G(\R), \] where $\Gamma=(G(\R)\times K_f)\cap G(\Q)$. In particular this is the case for $G=\operatorname{SL}(n)$. Let $K_\infty\subset G(\R)$ be a maximal compact subgroup. Let \begin{equation}\label{symspace} \widetilde X:=G(\R)^1/K_\infty \end{equation} be the associated global Riemannian symmetric space. Given an open compact subgroup $K_f\subset G(\A_f)$, we define the arithmetic manifold $X(K_f)$ by \begin{equation}\label{adel-quot2} X(K_f):= G(\Q)\bs (\widetilde X \times G(\A_f))/K_f. \end{equation} By \eqref{adel-quot1} we have \begin{equation} X(K_f)=\bigsqcup_{i=1}^l \left(\Gamma_i\bs \widetilde X\right), \end{equation} where each component $\Gamma_i\bs \widetilde X$ is a locally symmetric space. We will assume that $K_f$ is neat. Then $X(K_f)$ is a locally symmetric manifold of finite volume. Now consider $G=\operatorname{GL}(n)$ as algebraic group over $\Q$. Then $A_G(\R)^0$ is the group of scalar matrices with a positive real scalar and $K_\infty=\operatorname{O}(n)$. Let $N=\prod_pp^{r_p}$, $r_p\ge 0$. Put \[ K_p(N):=\{k\in G(\Z_p)\colon k\equiv 1\;\text{mod}\;p^{r_p}\Z_p\} \] and \[ K(N):=\prod_{p<\infty}K_p(N). \] Then $K(N)$ is an open compact subgroup of $G(\A_f)$ and \begin{equation}\label{gln-iso} A_G(\R)^0 G(\Q)\bs G(\A)/K(N)\cong \bigsqcup_{i=1}^{\varphi(N)}\Gamma(N)\bs SL(n,\R) \end{equation} where $\varphi(N)=\#[(\Z/N\Z)^\ast]$ (see \cite{Ar6}). Hence we have \begin{equation}\label{l2-gln} L^2(A_G(\R)^0 G(\Q)\bs G(\A))^{K(N)}\cong \bigoplus_{i=1}^{\varphi(N)} L^2(\Gamma(N)\bs \operatorname{SL}(n,\R)) \end{equation} as $\operatorname{SL}(n,\R)$-modules. We have \[ \widetilde X=\operatorname{SL}(n,\R)/\operatorname{SO}(n). \] Let \begin{equation} X(N):=G(\Q)\bs (\widetilde X\times G(\A_f))/K(N). \end{equation} Let $\nu\colon K_\infty\to\operatorname{GL}(V_\nu)$ be a finite dimensional unitary representation of $K_\infty$. Let $\widetilde E_\nu$ be the associated homogeneous Hermitian vector bundle over $\widetilde X$. Over each component of $X(K_f)$, $\widetilde E$ induces a locally homogeneous Hermitian vector bundle $E_{i,\nu}\to\Gamma_i\bs\widetilde X$. Let \begin{equation}\label{vectbdl} E_\nu:=\bigsqcup_{i=1}^l E_{i,\nu}. \end{equation} Then $E_\nu$ is a vector bundle over $X(K_f)$, which is locally homogeneous. \section{Truncation and the geometric side of the trace formula} \label{sec-trunc} The Arthur trace formula is obtained by truncating the kernels of integral operators associated to functions in $C^\infty_c(G(\A)^1)$. On the other hand, the regularization of the trace of heat operators is based on the truncation of the underlying locally symmetric space. In this section we compare the two methods. Let $P_0$ be the fixed minimal parabolic subgroup of $G$. For $f\in C^\infty_c(G(\A)^1)$ let \[ K_f(x,y)=\sum_{\gamma\in G(\Q)}f(x^{-1}\gamma y). \] This is the kernel of an integral operator. In general, $K_f(x,x)$ is not integrable over $G(\Q)\bs G(\A)^1$ and needs to be truncated to get an integrable function. To define the truncated kernel we need to introduce some notations. Let $P=M_PN_P$ be a standard parabolic subgroup and let $Q$ be a parabolic subgroup containing $P$. Let $\Delta^Q_P$ be the set of simple roots of $(M_Q\cap P,A_P)$. Similarly, we have the set of coroots $\Delta_0^\vee$ and, more generally and, the set $(\Delta_P^Q)^\vee$ which forms a basis of $\af_P^Q:=\af_P\cap \af_0^Q$. We denote the basis of $(\af_P^Q)^\ast$ (resp. $\af_P^Q$) dual to $(\Delta_P^Q)^\vee$ (resp. $\Delta_P^Q$) by $\hat\Delta_P^Q$ (resp. $(\hat\Delta_P^Q)^\vee$. Let $\tau_P^Q$ and $\widehat\tau_P^Q$ denote the characteristic functions of the set \[ \{X\in\af_0\colon \langle\alpha,X\rangle>0\;\text{for}\;\text{all}\; \alpha\in\Delta_P^Q\} \] and \[ \{X\in\af_0\colon \langle\varpi,X\rangle>0\;\text{for}\;\text{all} \;\varpi\in\hat\Delta_P^Q\}, \] respectively. If $Q=G$, we will suppress the superscript. Moreover we put $\tau_0:=\tau_0^G$ and $\hat\tau_0:=\widehat\tau_0^G$. Now we can define the truncated kernel. Put \[ K^P_f(x,y):=\int_{N_P(\Q)\bs N_P(\A)}\sum_{\gamma\in P(\Q)} f(x^{-1}\gamma ny)\;dn. \] Let $H_P\colon G(\A)\to\af_P$ be the map defined by \eqref{hp}. For any $T\in \af_0^+$ define \begin{equation}\label{truncker1} k^T(x,f):=\sum_P (-1)^{\dim(A_P/A_G)}\sum_{\delta\in P(\Q)\bs G(\Q)} K^P_f(\delta x,\delta x)\widehat\tau_P(H_P(\delta x)-T_P), \end{equation} where $T_P$ denotes the projection of $T$ on $\af_P$. Note that the term in \eqref{truncker1} which corresponds to $P=G$ is $K_f(x,x)$. If $G(\Q)\bs G(\A)^1$ is compact, there are no proper parabolic subgroups of $G$ over $\Q$. Thus, in this case we have $k^T(x,f)=K_f(x,x)$, and the truncation operation is trivial. By \cite[Theorem 6.1]{Ar4} the integral \begin{equation}\label{trf1} J^T(f):=\int_{G(\Q)\bs G(\A)^1} k^T(x,f)\;dx \end{equation} converges absolutely. This is the first step toward the trace formula. As shown by Hoffmann \cite{Ho}, $J^T(f)$ is defined for a larger class of functions $f$ and $J^T(f)$ is a polynomial in $T\in\af_0$ of degree at most $d_0=\dim\af_{P_0}^G$. There is a distinguished point $T_0\in\af_0$ specified by \cite[Lemma 1.1]{Ar3}, and Arthur defines the distribution $J$ on $G(\A)^1$ by \begin{equation}\label{j-distr} J(f):=J^{T_0}(f),\quad f\in C^\infty_c(G(\A)^1). \end{equation} This is the geometric side of the trace formula. To distinguish it from the spectral side, we will denote it by $J_{\geo}$. In \cite{Ar1} Arthur has introduced the coarse geometric expansion of $J^T(f)$. To define it, one has to introduce an equivalence relation in $G(\Q)$. Define two elements $\gamma$ and $\gamma^\prime$ in $G(\Q)$ to be equivalent, if the semisimple components $\gamma_s$ and $\gamma_s^\prime$ of their Jordan decompositions are $G(\Q)$-conjugate. Let $\cO$ be the set of equivalence classes. Note that the set $\cO$ is in obvious bijection with the semisimple conjugacy classes in $G(\Q)$. Furthermore, in case $G=\operatorname{GL}(n)$, the Jordan decomposition is given by the Jordan normal form. For $\of\in\cO$ and $f\in C^\infty_c(G(\A)^1)$ let \[ K_\of^P(x,y):=\int_{N_P(\Q)\bs N_P(\A)}\sum_{\gamma\in P(\Q)\cap \of} f(x^{-1}\gamma ny)\; dn. \] Given $T\in\af_0$ and $x\in G(\A)^1$, let \begin{equation}\label{truncker2} k^T_\of(x,f):=\sum_{P} (-1)^{\dim(A_P/A_G)}\sum_{\delta\in P(\Q)\bs G(\Q)} K_\of^P(\delta x,\delta x)\widehat\tau_P\left(H_P(\delta x)-T_P\right). \end{equation} Let \begin{equation}\label{distrib5} J^T_\of(f):=\int_{G(\Q)\bs G(\A)^1} k^T_\of(x,f)\;dx. \end{equation} The integral converges absolutely and one obtains an absolutely convergent expansion \begin{equation}\label{coarseexp} J^T(f)=\sum_{\of\in\cO} J^T_\of(f), \quad f\in C_c^\infty(G(\A)^1). \end{equation} This is the coarse geometric expansion, introduced in \cite{Ar1}. In \cite{FL1}, Finis and Lapid have shown that the coarse geometric expansion \eqref{coarseexp} extends continuously to the space of Schwartz functions $\Co(G(\A)^1)$ which is defined as follows. For any compact open subgroup $K_f$ of $G(\A_f)$ the space $G(\A)^1/K_f$ is the countable disjoint union of copies of $G(\R)^1=G(\R)\cap G(\A)^1$ and therefore, it is a differentiable manifold. Any element $X\in\mathcal{U}(\gf^1_\infty)$ of the universal enveloping algebra of the Lie algebra $\gf_\infty^1$ of $G(\R)^1$ defines a left invariant differential operator $f\mapsto f\ast X$ on $G(\A)^1/K_f$. Let $\Co(G(\A)^1;K_f)$ be the space of smooth right $K_f$-invariant functions on $G(\A)^1$ which belong, together with all their derivatives, to $L^1(G(\A)^1)$. The space $\Co(G(\A)^1;K_f)$ becomes a Fr\'echet space under the seminorms \[ \|f\ast X\|_{L^1(G(\A)^1)},\quad X\in\mathcal{U}(\gf^1_\infty). \] Denote by $\Co(G(\A)^1)$ the union of the spaces $\Co(G(\A)^1;K_f)$ as $K_f$ varies over the compact open subgroups of $G(\A_f)$ and endow $\Co(G(\A)^1)$ with the inductive limit topology. For $f\in\Co(G(\A)^1;K_f)$ and $\of\in\cO$ let $J^T(f)$ and $J^T_\of(f)$ be defined by \eqref{trf1} and \eqref{distrib5}, respectively. By \cite[Theorem 7.1]{FL1}, the integrals defining $J^T(f)$ and $J^T_\of(F)$ are absolutely convergent and we have \begin{equation}\label{coarseexp1} J^T(f)=\sum_{\of\in\cO}J_\of^T(f),\quad f\in\Co(G(\A)^1;K_f). \end{equation} We shall now discuss how $J^T(f)$ is related the integral of the kernel over the truncated manifold, where the truncated manifold is defined by a certain height function. For $T\in\af_0$ let \begin{equation}\label{trunc1} G(\A)^1_{\le T}=\{g\in G(\A)^1\colon \hat\tau_0(T-H_0(\gamma g))=1, \;\text{for}\;\text{all}\;\gamma\in G(\Q)\}. \end{equation} Note that by definition, $G(\A)^1_{\le T}$ is $G(\Q)$-invariant. Furthermore, for $T_1\in\af_0$ let \[ {\mathfrak S}_{T_1}=\{x\in G(\A)\colon \tau_0(H_0(x)-T_1)=1\} \] and more generally \[ {\mathfrak S}^P_{T_1}=\{x\in G(\A)\colon \tau_0^P(H_0(x)-T_1)=1\} \] for any $P\supset P_0$. Note that these sets are left $P_0(\A)^1$-invariant. By reduction theory, there exists $T_1\in\af_0$ such that \[ P(\Q){\mathfrak S}^P_{T_1}=G(\A) \] for all $P\supset P_0$, in particular for $P=G$. We fix such $T_1$. Let \begin{equation}\label{dT} d(T)=\min_{\alpha\in\Delta_0}\langle\alpha,T\rangle. \end{equation} There exists $d_0>0$, which depends only on $G$, $P_0$ and ${\bf K}$, such that for all $T\in\af_0$ with $d(T)>d_0$ one has \[ G(\A)^1_{\le T}\cap {\mathfrak S}_{T_1}=\{g\in G(A)^1\colon \tau_0(H_0(g)-T_1) \hat\tau_0(T-H_0(g))=1\}. \] For $f\in \Co(G(\A)^1)$ recall that \begin{equation}\label{kernel4} K_f(x,y)=\sum_{\gamma\in G(\Q)}f(x^{-1}\gamma y) \end{equation} The series converges absolutely and uniformly on compact subsets. Then the following theorem, which is an immediate consequence of \cite[Theorem 7.1]{FL1}, establishes the relation between $J^T(f)$ and naive truncation. \begin{theo}\label{theo-trunc} For every open compact subgroup $K_f$ of $G(\A_f)$ there exists $r\ge 0$ and a continuous seminorm $\mu$ on $\Co(G(\A)^1;K_f)$ such that \[ \left|\int_{G(\Q)\bs G(\A)^1_{\le T}}K_f(x,x)\;dx-J^T(f)\right|\le\mu(f) (1+\|T\|)^r e^{-d(T)} \] for all $f\in \Co(G(\A)^1;K_f)$ and $T\in\af_0$ such that $d(T)>d_0$. \end{theo} \begin{proof} For $\of\in\cO$ let \[ K_\of(x,y):=\sum_{\gamma\in\of} f(x^{-1}\gamma y). \] We have \begin{equation}\label{kernel-sum} K_f(x,y)=\sum_{\of\in\cO} K_\of(x,y), \end{equation} where the series converges absolutely. Using \eqref{coarseexp1}, we get \[ \left|\int_{G(\Q)\bs G(\A)^1_{\le T}}K_f(x,x)\;dx-J^T(f)\right|\le \sum_{\of\in\cO} \left|\int_{G(\Q)\bs G(\A)^1_{\le T}}K_\of(x,x)\;dx - J_\of^T(f)\right|. \] and the theorem follows from \cite[Theorem 7.1]{FL1}. We note that for the case of compactly supported functions $f$ this is due to Arthur (see \cite[\S 7]{Ar1}). \end{proof} \section{The non-invariant trace formula}\label{sec-trform} \setcounter{equation}{0} Arthur's (non-invariant) trace formula is the equality \begin{equation}\label{tracef1} J_{\geo}(f)=J_{\spec}(f),\quad f\in C_c^\infty(G(\A)^1), \end{equation} of the geometric side $J_{\geo}(f)$ and the spectral side $J_{\spec}(f)$ of the trace formula. The geome\-tric side has been described in the previous section. In this section we recall the definition of the spectral side, and in particular the refinement of the spectral expansion obtained in \cite{FLM1}. Combining \cite{FLM1} and \cite{FL1}, it follows that \eqref{tracef1} extends continuously to $f\in\Co(G(\A)^1)$. The main ingredient of the spectral side are logarithmic derivatives of intertwining operators. We briefly recall the structure of the intertwining operators. Let $P\in\cP(M)$. Let $U_P$ be the unipotent radical of $P$. Recall that we denote by $\Sigma_P\subset\af_P^*$ the set of reduced roots of $A_M$ of the Lie algebra $\mathfrak{u}_P$ of $U_P$. Let $\Delta_P$ be the subset of simple roots of $P$, which is a basis for $(\af_P^G)^*$. Write $\af_{P,+}^*$ for the closure of the Weyl chamber of $P$, i.e. \[ \mathfrak{a}_{P,+}^*=\{\lambda\in\mathfrak{a}_M^*:\sprod{\lambda}{\alpha^\vee}\ge0 \text{ for all }\alpha\in\Sigma_P\} =\{\lambda\in\mathfrak{a}_M^*:\sprod{\lambda}{\alpha^\vee}\ge0\text{ for all } \alpha\in\Delta_P\}. \] Denote by $\delta_P$ the modulus function of $P(\A)$. Let $\bar{\mathcal A}_2(P)$ be the Hilbert space completion of \[ \{\phi\in C^\infty(M(\Q)U_P(\A)\bs G(\A)):\delta_P^{-\frac12}\phi(\cdot x)\in L^2_{\operatorname{disc}}(A_M(\R)^0 M(\Q)\bs M(\A)),\ \forall x\in G(\A)\} \] with respect to the inner product \[ (\phi_1,\phi_2)=\int_{A_M(\R)^0 M(\Q)\bU_P(\A)\bs \G(\A)}\phi_1(g) \overline{\phi_2(g)}\ dg. \] Let $\alpha\in\Sigma_M$. We say that two parabolic subgroups $P,Q\in\cP(M)$ are \emph{adjacent} along $\alpha$, and write $P|^\alpha Q$, if $\Sigma_P\cap-\Sigma_Q=\{\alpha\}$. Alternatively, $P$ and $Q$ are adjacent if the group $\langle P,Q\rangle$ generated by $P$ and $Q$ belongs to $\cF_1(M)$ (see \eqref{f1} for its definition). Any $R\in\cF_1(\M)$ is of the form $\langle P,Q\rangle$, where $P,Q$ are the elements of $\cP(M)$ contained in $R$. We have $P|^\alpha Q$ with $\alpha^\vee\in\Sigma_P^\vee \cap\af^R_M$. Interchanging $P$ and $Q$ changes $\alpha$ to $-\alpha$. For any $P\in\cP(M)$ let $H_P\colon G(\A)\rightarrow\af_P$ be the extension of $H_M$ to a left $U_P(\A)$-and right $\K$-invariant map. Denote by $\cA^2(P)$ the dense subspace of $\bar\cA^2(P)$ consisting of its $\K$- and $\mathfrak{z}$-finite vectors, where $\mathfrak{z}$ is the center of the universal enveloping algebra of $\mathfrak{g} \otimes \C$. That is, $\cA^2(P)$ is the space of automorphic forms $\phi$ on $U_P(\A)M(\Q)\bs G(\A)$ such that $\delta_P^{-\frac12}\phi(\cdot k)$ is a square-integrable automorphic form on $A_M(\R)^0 M(\Q)\bs M(\A)$ for all $k\in\K$. Let $\rho(P,\lambda)$, $\lambda\in\af_{M,\C}^*$, be the induced representation of $G(\A)$ on $\bar\cA^2(P)$ given by \[ (\rho(P,\lambda,y)\phi)(x)=\phi(xy)e^{\sprod{\lambda}{H_P(xy)-H_P(x)}}. \] It is isomorphic to the induced representation \[ \operatorname{Ind}_{P(\A)}^{G(\A)}\left(L^2_{\operatorname{disc}}(A_M(\R)^0 M(\Q)\bs M(\A)) \otimes e^{\sprod{\lambda}{H_M(\cdot)}}\right). \] For alternative descriptions see \cite[\S 1]{Ar8}, \cite[I.2.17, I.2.18]{MW}. For $P,Q\in\cP(M)$ let \[ M_{Q|P}(\lambda):\cA^2(P)\to\cA^2(Q),\quad\lambda\in\af_{M,\C}^*, \] be the standard \emph{intertwining operator} \cite[\S 1]{Ar9}, which is the meromorphic continuation in $\lambda$ of the integral \[ [M_{Q|P}(\lambda)\phi](x)=\int_{U_Q(\A)\cap U_P(\A)\bs U_Q(\A)}\phi(nx) e^{\sprod{\lambda}{H_P(nx)-H_Q(x)}}\ dn, \quad \phi\in\cA^2(P), \ x\in G(\A). \] Given $\pi\in\Pi_{\di}(M(\A))$, let $\cA^2_\pi(P)$ be the space of all $\phi\in\cA^2(P)$ for which the function $M(\A)\ni x\mapsto \delta_P^{-\frac{1}{2}}\phi(xg)$, $g\in G(\A)$, belongs to the $\pi$-isotypic subspace of the space $L^2(A_M(\R)^0 M(\Q)\bs M(\A))$. For any $P\in\cP(\M)$ we have a canonical isomorphism of $G(\A_f)\times(\mathfrak{g}_{\C},K_\infty)$-modules \[ j_P:\operatorname{Hom}(\pi,L^2(A_M(\R)^0 M(\Q)\bs M(\A)))\otimes \operatorname{Ind}_{P(\A)}^{G(\A)}(\pi)\rightarrow\cA^2_\pi(P). \] If we fix a unitary structure on $\pi$ and endow $\operatorname{Hom}(\pi,L^2(A_M(\R)^0 M(\Q)\bs M(\A)))$ with the inner product $(A,B)=B^\ast A$ (which is a scalar operator on the space of $\pi$), the isomorphism $j_P$ becomes an isometry. Suppose that $P|^\alpha Q$. The operator $M_{Q|P}(\pi,s):=M_{Q|P}(s\varpi)|_{\cA^2_\pi(P)}$, where $\varpi\in \af^\ast_M$ is such that $\langle\varpi,\alpha^\vee\rangle=1$, admits a normalization by a global factor $n_\alpha(\pi,s)$ which is a meromorphic function in $s$. We may write \begin{equation} \label{normalization} M_{Q|P}(\pi,s)\circ j_P=n_\alpha(\pi,s)\cdot j_Q\circ(\operatorname{Id}\otimes R_{Q|P}(\pi,s)) \end{equation} where $R_{Q|P}(\pi,s)=\otimes_v R_{Q|P}(\pi_v,s)$ is the product of the locally defined normalized intertwining operators and $\pi=\otimes_v\pi_v$ \cite[\S 6]{Ar9}, (cf.~\cite[(2.17)]{Mu2}). In many cases, the normalizing factors can be expressed in terms automorphic $L$-functions \cite{Sha1}, \cite{Sha2}. For example, let $G=\operatorname{GL}(n)$. Then the global normalizing factors $n_\alpha$ can be expressed in terms of Rankin-Selberg $L$-functions. The known properties of these functions are collected and analyzed in \cite[\S\S 4,5]{Mu1}. Write $M \simeq \prod_{i=1}^r \operatorname{GL} (n_i)$, where the root $\alpha$ is trivial on $\prod_{i \ge 3} \operatorname{GL} (n_i)$, and let $\pi \simeq \otimes \pi_i$ with representations $\pi_i \in \Pi_{\operatorname{disc}}(\operatorname{GL}(n_i,\A))$. Let $L(s,\pi_1\times\tilde\pi_2)$ be the completed Rankin-Selberg $L$-function associated to $\pi_1$ and $\pi_2$. It satisfies the functional equation \begin{equation}\label{functequ} L(s,\pi_1\times\tilde\pi_2)=\epsilon(\frac12,\pi_1\times\tilde\pi_2) N(\pi_1\times\tilde\pi_2)^{\frac12-s}L(1-s,\tilde\pi_1\times\pi_2) \end{equation} where $\abs{\epsilon(\frac12,\pi_1\times\tilde\pi_2)}=1$ and $N(\pi_1\times\tilde \pi_2)\in\N$ is the conductor. Then we have \begin{equation}\label{rankin-selb} n_\alpha (\pi,s) = \frac{L(s,\pi_1\times\tilde\pi_2)}{\epsilon(\frac12,\pi_1\times \tilde\pi_2)N(\pi_1\times\tilde\pi_2)^{\frac12-s}L(s+1,\pi_1\times\tilde\pi_2)}. \end{equation} We now turn to the spectral side. Let $L \supset M$ be Levi subgroups in ${\mathcal L}$, $P \in \mathcal{P} (M)$, and let $m=\dim\mathfrak{a}_L^G$ be the co-rank of $L$ in $G$. Denote by $\mathfrak{B}_{P,L}$ the set of $m$-tuples $\underline{\beta}=(\beta_1^\vee,\dots, \beta_m^\vee)$ of elements of $\Sigma_P^\vee$ whose projections to $\af_L$ form a basis for $\af_L^G$. For any $\underline{\beta}=(\beta_1^\vee,\dots,\beta_m^\vee)\in\mathfrak{B}_{P,L}$ let $\operatorname{vol}(\underline{\beta})$ be the co-volume in $\af_L^G$ of the lattice spanned by $\underline{\beta}$ and let \begin{align*} \Xi_L(\underline{\beta})&=\{(Q_1,\dots,Q_m)\in\cF_1(M)^m: \ \ \beta_i^{\vee}\in\af_M^{Q_i}, \, i = 1, \dots, m\}\\&= \{\langle P_1,P_1'\rangle,\dots,\langle P_m,P_m'\rangle): \ \ P_i|^{\beta_i}P_i', \, i = 1, \dots, m\}. \end{align*} For any smooth function $f$ on $\af_M^*$ and $\mu\in\af_M^*$ denote by $D_\mu f$ the directional derivative of $f$ along $\mu\in\af_M^*$. For a pair $P_1|^\alpha P_2$ of adjacent parabolic subgroups in $\cP(M)$ write \begin{equation}\label{intertw2} \delta_{P_1|P_2}(\lambda)=M_{P_2|P_1}(\lambda)D_\varpi M_{P_1|P_2}(\lambda): \cA^2(P_2)\rightarrow\cA^2(P_2), \end{equation} where $\varpi\in\af_M^*$ is such that $\sprod{\varpi}{\alpha^\vee}=1$. \footnote{Note that this definition differs slightly from the definition of $\delta_{P_1|P_2}$ in \cite{FLM1}.} Equivalently, writing $M_{P_1|P_2}(\lambda)=\Phi(\sprod{\lambda}{\alpha^\vee})$ for a meromorphic function $\Phi$ of a single complex variable, we have \[ \delta_{P_1|P_2}(\lambda)=\Phi(\sprod{\lambda}{\alpha^\vee})^{-1} \Phi'(\sprod{\lambda}{\alpha^\vee}). \] For any $m$-tuple $\mathcal{X}=(Q_1,\dots,Q_m)\in\Xi_L(\underline{\beta})$ with $Q_i=\langle P_i,P_i'\rangle$, $P_i|^{\beta_i}P_i'$, denote by $\Delta_{\mathcal{X}}(P,\lambda)$ the expression \begin{equation}\label{intertw3} \frac{\operatorname{vol}(\underline{\beta})}{m!}M_{P_1'|P}(\lambda)^{-1}\delta_{P_1|P_1'}(\lambda)M_{P_1'|P_2'}(\lambda) \cdots \delta_{P_{m-1}|P_{m-1}'}(\lambda)M_{P_{m-1}'|P_m'}(\lambda)\delta_{P_m|P_m'}(\lambda)M_{P_m'|P}(\lambda). \end{equation} In \cite[pp. 179-180]{FLM1} the authors defined a (purely combinatorial) map $\mathcal{X}_L: \mathfrak{B}_{P,L} \to {\mathcal F}_1 (M)^m$ with the property that $\mathcal{X}_L(\underline{\beta}) \in \Xi_L (\underline{\beta})$ for all $\underline{\beta} \in \mathfrak{B}_{P,L}$.\footnote{The map $\mathcal{X}_L$ depends in fact on the additional choice of a vector $\underline{\mu} \in (\mathfrak{a}^*_M)^m$ which does not lie in an explicit finite set of hyperplanes. For our purposes, the precise definition of $\mathcal{X}_L$ is immaterial.} For any $s\in W_M$ let $L_s$ be the smallest Levi subgroup in ${\mathcal L}(M)$ containing $w_s$. We recall that $\mathfrak{a}_{L_s}=\{H\in\mathfrak{a}_M\mid sH=H\}$. Set \[ \iota_s=\abs{\det(s-1)_{\mathfrak{a}^{L_s}_M}}^{-1}. \] For $P\in{\mathcal F}(M_0)$ and $s\in W(M_P)$ let $M(P,s):{\mathcal A}^2(P)\to{\mathcal A}^2(P)$ be as in \cite[p.~1309]{Ar3}. $M(P,s)$ is a unitary operator which commutes with the operators $\rho(P,\lambda,h)$ for $\lambda\in{\mathrm i}\mathfrak{a}_{L_s}^*$. Finally, we can state the refined spectral expansion. \begin{theo}[\cite{FLM1}] \label{thm-specexpand} For any $h\in C_c^\infty(G(\A)^1)$ the spectral side of Arthur's trace formula is given by \begin{equation}\label{specside1} J_{\spec}(h) = \sum_{[M]} J_{\spec,M} (h), \end{equation} $M$ ranging over the conjugacy classes of Levi subgroups of $G$ (represented by members of $\mathcal{L}$), where \begin{equation}\label{specside2} J_{\spec,M} (h) = \frac1{\card{W(M)}}\sum_{s\in W(M)}\iota_s \sum_{\underline{\beta}\in\mathfrak{B}_{P,L_s}}\int_{{\mathrm i}(\mathfrak{a}^G_{L_s})^*} \operatorname{tr}(\Delta_{\mathcal{X}_{L_s}(\underline{\beta})}(P,\lambda)M(P,s)\rho(P,\lambda,h))\ d\lambda \end{equation} with $P \in \mathcal{P}(M)$ arbitrary. The operators are of trace class and the integrals are absolutely convergent with respect to the trace norm and define distributions on $\Co(G(\A)^1)$. \end{theo} Note that the term corresponding to $M=G$ is $J_{\spec,G} (h) = \operatorname{tr} R_{\operatorname{disc}}(h)$. \section{The unipotent contribution to the trace formula}\label{sec-unipotent} \setcounter{equation}{0} In this section we assume that $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$ as algebraic groups over $\Q$, and we specialize to one of these groups at some points. The purpose is to analyze the unipotent contribution to the geometric side of the trace formula. The point of departure is the coarse geometric expansion of $J_{\geo}$ as a sum of distributions \begin{equation}\label{geoside} J_{\geo}(f)=\sum_{{\mathfrak o}\in\cO}J_{{\mathfrak o}}(f),\quad f\in C^\infty_c(G(\A)^1), \end{equation} parametrized by the set $\cO$ of semisimple conjugacy classes of $G(\Q)$. The distribution $J_{{\mathfrak o}}(f)$ is the value at $T=0$ of the polynomial $J_{{\mathfrak o}}^T(f)$ defined in \cite{Ar1}. In particular, following Arthur, we write $J_{\operatorname{unip}}(f)$ for the contribution corresponding to the class of $\{1\}$. Let $K(N)$ be a principal congruence subgroup of level $N\ge 3$. By \cite[Corollary 5.2]{LM} there exists a bi-$K_\infty$-invariant compact neighborhood $\omega$ of $K(N)$ in $G(\A)^1$ such that \begin{equation}\label{unip-contr1} J_{\geo}(f)=J_{\operatorname{unip}}(f). \end{equation} for all $f\in C_c^\infty(G(\A)^1)$ supported in $\omega$. (\cite[Corollary 5.2]{LM} was only stated for $\operatorname{GL}(n)$, but its proof holds also for $\operatorname{SL}(n)$ without modification.) For our applications we can choose $f$ such that \eqref{unip-contr1} holds if we restrict to $K_f$ with $K_f\subseteq K(N)$ for some $N\ge3$. See \S~\ref{sec:examples} for computations for $n=2,3$ regarding orbits of finite order which appear in the case $K_f=G(\hat\Z)$. To analyze $J_{\operatorname{unip}}(f)$ we use Arthur's fundamental result (\cite[Corollaries 8.3 and 8.5]{Ar4}) to express $J_{\operatorname{unip}}(f)$ in terms of weighted orbital integrals. To state the result we recall some facts about weighted orbital integrals. Let $S$ be a finite set of places of $\Q$ containing $\infty$. Set \[ \Q_S=\prod_{v\in S}\Q_v,\quad \mathrm{and}\quad G(\Q_S)=\prod_{v\in S}G(\Q_v). \] Let \[ G(\Q_S)^1=G(\Q_S)\cap G(\A)^1 \] and write $C^\infty_c(G(\Q_S)^1)$ for the space of functions on $G(\Q_S)^1$ obtained by restriction of functions in $C^\infty_c(G(\Q_S))$ to $G(\Q_S)^1$. Further, let $\A^S=\prod_{v\not\in S}\Q_v$ be the restricted product over all places outside of $S$, and define $G(\A^S)$ similarly as above. \subsection{The fine geometric expansion} Let $M\in{\mathcal L}$ and $\gamma\in M(\Q_S)$. The general weighted orbital integrals $J_M(\gamma,f)$ defined in (\cite{Ar5}) are distributions on $G(\Q_S)$. Denote by $H_\gamma$ the centralizer of $\gamma$ in a subgroup $H$ of $G$. If $\gamma$ is such that $M_\gamma=G_\gamma$ then $J_M(\gamma,f)$ is given by an integral of the form \begin{equation}\label{orbint} J_M(\gamma,f)=\big|D(\gamma)\big|^{1/2}\int_{{G_\gamma}(\Q_S)\bs G(\Q_S)} f(x^{-1}\gamma x) v_M(x)\ dx, \end{equation} where $D(\gamma)$ is defined in \cite[p. 231]{Ar5} and $v_M(x)$ is the weight function associated to the $(G,M)$-family $\{v_P(\lambda,x)\colon P\in\mathcal{P}(M)\}$ defined in \cite[p.230]{Ar5}. It is a left $M(\A)$-invariant and right $\mathbf{K}$-invariant function on $G(\A)$. In particular, in the case $M=G$ (in which $v_M\equiv1$) we obtain the usual (invariant) orbital integral. Of course, implicit in $J_M(\gamma,f)$ is a choice of a Haar measure on $G_\gamma(\Q_S)$. When the condition $G_\gamma\subset M$ is not satisfied (for example, if $\gamma$ is unipotent and $M\neq G$), the definition of $J_M(\gamma,f)$ is more complicated. It is obtained as a limit of a linear combination of integrals as above. For more details we refer to \cite{Ar5}, see also the description below. If $\gamma$ belongs to the intersection of $M(\Q_S)$ with $G(\Q_S)^1$, one can obviously define the corresponding weighted orbital integral as a linear form on $C^\infty_c(G(\Q_S)^1)$. Note that $J_M(\gamma,f)$ depends only on the $M(\Q_S)$-conjugacy class of $\gamma$. To state the fine expansion of $J_{\operatorname{unip}}(f)$, we need to introduce a certain equivalence relation as defined in \cite{Ar4,Ar7}. Let $\cU_M$ denote the variety of unipotent elements in $M$ and $\cU_M(A)$ its $A$-points for any $\Z$-algebra $A$. We say that $u_1, u_2\in \cU_M(\Q)$ are \emph{$(M,S)$-equivalent} if they are conjugate by some element in $M(\Q_S)$, cf.~\cite[\S7]{Ar4}. We denote by $\left(\cU_M(\Q)\right)_{M,S}$ the set of $(M,S)$-equivalence classes in $\cU_M(\Q)$. We note that $\left(\cU_M(\Q)\right)_{M,S}$ is finite for any $S$ but may get larger as $S$ grows. We get an injective map \[ \left(\cU_M(\Q)\right)_{M,S}\longrightarrow \left(\cU_M(\Q_S)\right)_{M(\Q_S)} \] where $\left(\cU_M(\Q_S)\right)_{M(\Q_S)}$ denotes the set of $M(\Q_S)$-conjugacy classes in $\cU_M(\Q_S)$. Note that this last set is evidently the same as the direct product $\prod_{v\in S} \left(\cU_M(\Q_v)\right)_{M(\Q_v)}$ over the $M(\Q_v)$-conjugacy classes in $\cU_M(\Q_v)$, $v\in S$. In particular, we can identify an equivalence class $U\in\left(\cU_M(\Q)\right)_{M,S}$ with its image under the above map, that is, with a tuple $(U_v)_{v\in S}$ of $M(\Q_v)$-conjugacy classes in $\cU_M(\Q_v)$. By~\cite[Theorem 8.1]{Ar4} we have \begin{equation}\label{finegeoexp} J_{\operatorname{unip}}(f\otimes\mathbf 1_{K^S})=\operatorname{vol}(G(\Q)\backslash G(\A)^1)f(1) + \sum_{(M,U)}a^M(S,U) J_M(U,f), \end{equation} where $(M,U)$ runs over all pairs of Levi subgroups $M\in {\mathcal L}$ and $U\in \left(\cU_M(\Q)\right)_{M,S}$ with $(M,U)\neq (G, 1)$. Here $f\in C^\infty_c(G(\Q_S)^1)$, $\mathbf 1_{K^S}$ is the characteristic function of the standard maximal compact of $G(\A^S)$, and $a^M(S,U)$ are certain constants which depend on the normalization of measures (but are usually not known explicitly). The distributions $J_M(U,f)$ can be written as weighted orbital integrals \cite[p. 256]{Ar5}. In our case the integrals simplify as we are going to see later. \subsection{Unipotent conjugacy classes} Let $\left(\cU_M\right)$ denote the set of geometric (that is, over an algebraic closure $\bar\Q$ of $\Q$) $M$-conjugacy classes in $\cU_M$. Then $\left(\cU_M\right)$ is a finite set. Each ${\mathcal V}\in \left(\cU_M\right)$ is defined over $\Q$ and the set of $\Q$-points ${\mathcal V}(\Q)$ is non-empty. More precisely, each ${\mathcal V}\in\left(\cU_M\right)$ corresponds to a partition of $n$, and ${\mathcal V}(\Q)$ contains the matrix with Jordan normal form corresponding to that partition. For any ${\mathcal V}\in\left(\cU_M\right)$ the set ${\mathcal V}(\Q)$ is closed under the $(M,S)$-equivalence relation and we write $\left({\mathcal V}(\Q)\right)_{M,S}$ for the finite set of $(M,S)$-equivalence classes in ${\mathcal V}(\Q)$. \begin{remark} If $G=\operatorname{SL}(n)$, there might be infinitely many $M(\Q)$-conjugacy classes in ${\mathcal V}(\Q)$ depending on the type of ${\mathcal V}(\Q)$. This is in contrast to the case of $\operatorname{GL}(n)$, where the (finite) set of geometric unipotent conjugacy classes is in bijection with the set of rational unipotent conjugacy classes. \end{remark} Each class ${\mathcal V}\in\left(\cU_M\right)$ is a Richardson class, that is, there exists a standard parabolic subgroup $Q=LV\in{\mathcal L}(M)$ such that ${\mathcal V}$ is induced from the trivial orbit in $L$ to $M$, see \cite[\S 5.5 Proposition]{Hu}. Equivalently, the intersection of ${\mathcal V}$ with $V$ is an open and dense subset of $V$. Note that every $(M,S)$-equivalence class $U\subseteq {\mathcal V}(\Q)$ has a representative in $V(\Q)$. Now let ${\mathcal V}\in \left(\cU_M\right)$ and $U=(U_v)_{v\in S}\in \left({\mathcal V}(\Q_S)\right)_{M(\Q_S)}$. Here we write $\left({\mathcal V}(\Q_S)\right)_{M(\Q_S)}=\prod_{v\in S}\left({\mathcal V}(\Q_v)\right)_{M(\Q_v)}$ for the set of $M(\Q_S)$-conjugacy classes in ${\mathcal V}(\Q_S)$. To understand the $S$-adic integral $J_M(U,f)$, we decompose it into a sum of products of integrals at $\infty$ and at the finite places $S_{\fin}=S\backslash\{\infty\}$. More precisely, for every pair of Levi subgroups $L_1, L_2\in {\mathcal L}(M)$ there exists a coefficient $d_M^G(L_1, L_2)\in\C$ such that \begin{equation}\label{orbint11} J_M(U,f)=\sum_{L_1, L_2\in{\mathcal L}(M)} d_M^G(L_1, L_2) J_M^{L_1}(U_{\infty}, f_{\infty, Q_1}) J_M^{L_2}(U_{\fin}, f_{\fin, Q_2}) \end{equation} (see~\cite{Ar3}) where $U_{\fin}=(U_v)_{v\in S_{\fin}}\in \left({\mathcal V}(\Q_{S_{\fin}})\right)_{M(\Q_{S_{\fin}})}$, and $Q_1=L_1V_1\in \cP(L_1))$, $Q_2=L_2V_2\in\cP(L_2)$ are certain parabolic subgroups the exact choice of which does not matter to us. Moreover, \[ f_{\infty, Q_1}(m)=\delta_{Q_1}(m)^{1/2} \int_{K_{\infty}}\int_{V_1(\R)} f(k^{-1} mv k)\, dk\, dv, \; m\in L_1(\R), \] with $f_{\fin, Q_2}$ being defined analogously, and the coefficients $d_M^G(L_1, L_2)$ are independent of $S$ and they vanish unless the natural map $\mathfrak{a}_M^{L_1}\oplus\mathfrak{a}_M^{L_2}\longrightarrow \mathfrak{a}_M^G$ is an isomorphism. In case the coefficient does not vanish, it depends on the chosen measures on $\mathfrak{a}_M^{L_1}$, $\mathfrak{a}_M^{L_2}$ and $\mathfrak{a}_M^G$. The distributions $J_M^{L}(U, F)$, for $L\in{\mathcal L}(M)$, any finite set $S'$ of places of $\Q$, $F\in C^\infty(L(\Q_{S'}))$, and a unipotent conjugacy class $U\subseteq L(\Q_{S'})$, are defined similarly as the weighted orbital integrals $J_M^G=J_M$ but with $L$ in place of $G$. Our test function at the places in $v\in S_{\fin}$ is fixed once and for all so that the integrals at those places can be viewed as constant for our purposes. Hence we need to understand the integral at the Archimedean place. We therefore need to better understand the unipotent conjugacy classes over $\R$. If $G=\operatorname{GL}(n)$, the unipotent orbits in $\operatorname{GL}_n(\R)$ and all its Levi subgroups in ${\mathcal L}$ are easy to describe as they are in one-to-one correspondence with the geometric unipotent conjugacy classes and therefore classified by partitions of $n$. We assume for the moment that $G=\operatorname{SL}(n)$. Note that $\operatorname{SL}_n(\R)$ is normalized by $\operatorname{GL}_n(\R)$. Moreover, each $M\in{\mathcal L}$ is of the form $M=\bar M\cap\operatorname{SL}(n)$ for a unique Levi subgroup $\bar M\in {\mathcal L}^{\operatorname{GL}(n)}$. Then $\bar K^M_{\infty}:=\operatorname{O}(n)\cap \bar M(\R)$ is a maximal compact subgroup in $\bar M$, $\bar K^M_{\infty}\cap M(\R)=K^M_{\infty}$, and $\bar M(\R)$ normalizes $M(\R)$. In particular, it makes sense to speak of $\bar K_{\infty}^M$-conjugation invariant functions on $M(\R)$. \begin{lem} Let ${\mathcal V}\in\left(\cU_M\right)$. For any two equivalence classes $U_1, U_2\in\left({\mathcal V}(\R)\right)_{M(\R)}$ there exists $k\in\bar K^M_{\infty}$ with $k^{-1} U_1 k=U_2$ and $\left({\mathcal V}(\R)\right)_{M(\R)}$ consists of at most two equivalence classes. More precisely, if $U_1, U_2\in\left({\mathcal V}(\R)\right)_{M(\R)}$ are the two distinct classes, we have $k^{-1} U_1 k=U_2$ with $k=\operatorname{diag}(-1, 1, \ldots, 1)\in \bar K^M_{\infty}$. \end{lem} \begin{proof} Let $u_1\in U_1$ and $u_2\in U_2$. In $\bar M(\R)$, $u_1$ and $u_2$ are conjugate, that is, there exists some $g\in \bar M(\R)$ with $g^{-1} u_1 g=u_2$. Without loss of generality we can assume that $|\det g|=1$. If $\det g=1$, then $g\in \bar M(\R)\cap\operatorname{SL}_n(\R)=M(\R)$ and $U_1=U_2$. If $\det g=-1$, let $k=\operatorname{diag}(-1, 1,\ldots, 1)\in \bar K^M_{\infty}$ and $g_1=gk^{-1}$. Then $g_1\in \bar M(\R)\cap \operatorname{SL}_n(\R)=M(\R)$, and $U_2= g^{-1} U_1 g= k^{-1} g_1^{-1} U_1 g_1 k= k^{-1} U_1 k$ as asserted. \end{proof} \subsection{Measures on conjugacy classes} \begin{corollary}\label{cor:invariance:arch} If ${\mathcal V}\in\left(\cU_M\right)$ and $f_{\infty}\in C_c(G(\R))$ is conjugation invariant under $\bar K_{\infty}$, then we have $J_M(U_1,f_{\infty})=J_M(U_2, f_{\infty})$ for all $U_1, U_2\in\left({\mathcal V}(\R)\right)_{M(\R)}$. \end{corollary} \begin{proof} By the previous lemma, there are at most two distinct classes in $\left({\mathcal V}(\R)\right)_{M(\R)}$. If there is only one class in $\left({\mathcal V}(\R)\right)_{M(\R)}$, there is nothing to show. If there are two distinct classes $U_1, U_2$, they are conjugate to each other via the element $k=\operatorname{diag}(-1, 1, \ldots, 1)\in\bar K_{\infty}^M$. Let $u_1\in U_1$ and $u_2=k^{-1}u_1k\in U_2$. Then the centralizers of $u_1$ and $u_2$ in $G(\R)$ are the same, since they are the same in $\operatorname{GL}_n(\R)$ and $k$ normalizes $G(\R)=\operatorname{SL}_n(\R)$ in $\operatorname{GL}_n(\R)$. Hence the invariant measures on the $G(\R)$-conjugacy classes of $u_1$ and $u_2$ coincide, in particular, $J_G(U_1,f_{\infty})=J_G(U_2, f_{\infty})$ for every $\bar K_{\infty}$-conjugation invariant $f_{\infty}\in C_c(G(\R))$. The non-invariant measure defining $J_M(U_i,\cdot)$ can be written as the product of some weight function times the invariant measure. We shall see in the next section (see \eqref{eq:weight:function:invariance:P} and \eqref{eq:definition:weight}) that the weight function is invariant under the action of $k=\operatorname{diag}(-1, 1, \ldots, 1)$ so that the claim of the lemma follows for the weighted orbital integrals as well. \end{proof} The following corollary now is valid for $G=\operatorname{SL}(n)$ as well as $G=\operatorname{GL}(n)$. However, we shall only prove it for $\operatorname{SL}(n)$. For $G=\operatorname{GL}(n)$ the proof in fact is easier and was already given in \cite[Lemma 5.3]{LM} (see also \cite[Proposition 5]{Howe74}). \begin{corollary} For every ${\mathcal V}\in\left(\cU_G\right)$ there exists a standard parabolic subgroup $P=LV$ and a constant $c>0$ such that for every $\operatorname{O}(n)$-conjugation invariant function $f_{\infty}$ and every $U\in\left({\mathcal V}(\R)\right)_{G(\R)}$ the invariant orbital integral $J_G(U,f_{\infty})$ can be written as \[ J_G(U, f_{\infty})= c\int_{V(\R)} f_{\infty}(v)\, dv, \] where $dv$ denotes the Haar measure on $V(\R)$ normalized such that it coincides with the measure obtained from the Lebesgue measure on $\R^{\dim V}$ when $V(\R)$ is identified with $\R^{\dim V}$ via its matrix coordinates. \end{corollary} \begin{proof} Let $P=LV\in\mathcal{P}$ be a Richardson parabolic subgroup for ${\mathcal V}$ so that ${\mathcal V}(\R)\cap V(\R)$ is dense in $V(\R)$. We have ${\mathcal V}(\R)=\bigcup_{U\in\left({\mathcal V}(\R)\right)_{G(\R)}} U$ and this union is disjoint. Then \begin{equation}\label{eq:decomposition} {\mathcal V}(\R)\cap V(\R)=\bigcup_{U\in\left({\mathcal V}(\R)\right)_{G(\R)}} U\cap V(\R) \end{equation} is also a disjoint union which is dense in $V(\R)$. For each $U\in\left({\mathcal V}(\R)\right)_{G(\R)}$ we can pick a representative $u\in V(\R)$. Then the orbit of $u$ under $P(\R)$ equals $U\cap V(\R)$. Since $f_{\infty}$ is $\operatorname{O}(n)$-conjugation invariant, we have, using Iwasawa decomposition for $G(\R)$, that \[ J_G(U,f_{\infty})= \int_{P_u(\R)\backslash P(\R)}\delta_P(p)^{-1} f_{\infty} (p^{-1} u p)\, dp. \] It follows as in the proof of~\cite[Lemma 5.3]{LM} that \[ C_c^{\infty}(V(\R))\ni h\longrightarrow \int_{P_u(\R)\backslash P(\R)}\delta_P(p)^{-1} h (p^{-1} u p)\, dp \] is absolutely continuous with respect to the Haar measure on $V(\R)$. Hence \begin{equation}\label{eq:orb:int:sum} C_c^{\infty}(V(\R))\ni h\longrightarrow \sum_{u} \int_{P_u(\R)\backslash P(\R)}\delta_P(p)^{-1} h (p^{-1} u p)\, dp \end{equation} is also absolutely continuous with respect to the Haar measure on $V(\R)$. Here $u\in V(\R)$ runs over a set of representatives for the classes $U\in\left({\mathcal V}(\R)\right)_{G(\R)}$. Since the right hand side of \eqref{eq:decomposition} is a disjoint union and dense in $V(\R)$, the measure defined by the right hand side of~\eqref{eq:orb:int:sum} must be proportional to the Haar measure on $V(\R)$. Hence there exists a constant $C>0$ such that for every $f_{\infty}\in C_c^{\infty}(G(\R))$ we have \[ \sum_{U\in\left({\mathcal V}(\R)\right)_{G(\R)}} J_G(U,f_{\infty}) =C\int_{V(\R)} f_{\infty}(v)\, dv. \] By Corollary~\ref{cor:invariance:arch} and our assumption on $f_{\infty}$ we have $J_G(U_1, f_{\infty})=J_G(U_2, f_{\infty})$ for all $U_1, U_2\in \left({\mathcal V}(\R)\right)_{G(\R)}$. Hence \[ J_G(U,f_{\infty}) =\frac{C}{N}\int_{V(\R)} f_{\infty}(v)\, dv \] where $N$ is the number of classes in $\left({\mathcal V}(\R)\right)_{G(\R)}$. \end{proof} The weighted integral $J_M(U,f_{\infty})$ from \cite[p. 256]{Ar5} can be written as an integral over $\operatorname{Ind}_M^G U$ against the invariant measure on $\operatorname{Ind}_M^G U$ weighted by a certain function. Hence using the corollary above it follows that the real orbital integral $J_M(U,f_{\infty})$ simplifies for every $\operatorname{O}(n)$-conjugation invariant $f_{\infty}$ to \begin{equation}\label{weiorbint} J_M(U,f_{\infty})=\int_{N(\R)}f(x)w_{M,U}(x)\, dx \end{equation} where $Q=LN\in\mathcal{P}$ is a Richardson parabolic for $\operatorname{Ind}_M^G U$, the unipotent orbit induced from $M$ to $G$, and the weight function $w_{M,U}(x)$ is described in \cite[Lemma 5.4]{Ar5}. (See also \cite[\S 5]{LM}.) It is a finite linear combination of functions of the form $\prod_{i=1}^r\log{\norm{p_i(x)}}$ where $p_i$ are polynomials on $N_Q(\R)$ into an affine space, $i=1,\dots,r$ (not necessarily distinct) and \[ \norm{(y_1,\dots,y_m)}_v= \abs{y_1}_v^2+\dots+\abs{y_m}_v^2 \] (The fact that the product is over $r$ terms is implicit in \cite{Ar5} but follows from the proof.) For our purpose we need to describe the weight function in more detail which we shall do in the next section. \section{The weight function}\label{sec-wfct} \setcounter{equation}{0} In this section, again $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$. We are only interested in the situation over the field $\R$, but most of this section holds over $\Q_p$, $p<\infty$, as well. As before, $\cU_M$ denotes the unipotent variety in $M$ and ${\mathcal V}\in\left(\cU_M\right)$ a geometric conjugacy class. We write \[ \mathbf{K}^G=\begin{cases} \operatorname{SO}(n) &\text{if }G=\operatorname{SL}(n),\\ \operatorname{O}(n) &\text{if }G=\operatorname{GL}(n), \end{cases} \] and we write $\mathbf{K}$ for $\mathbf{K}^G$ if $G$ is clear from the context. We first recall the definition of the local weight functions from \cite{Ar5}. Let ${\mathcal V}\in\left(\cU_M\right)$ and $U\in \left({\mathcal V}(\R)\right)_{M(\R)}$. We denote by ${\mathcal V}^G$ the conjugacy class in $\left(\cU_G\right)$ induced from ${\mathcal V}$ along some parabolic subgroup in $\mathcal{P}(M)$ (they yield all the same induced class). Let $U^G\in \left({\mathcal V}^G(\R)\right)_{G(\R)}$ be such that $U^G\cap M(\R)=U$. (As explained above there are at most two different elements in $\left({\mathcal V}^G(\R)\right)_{G(\R)}$ which are conjugate by $\operatorname{diag}(-1, 1, \ldots, 1)$.) Arthur~\cite{Ar5} defines a weight function $w_{M,U}$ on a dense open subset of ${\mathcal V}^G(\R)$ such that the local weighted orbital integral $J_M(U,f)=J_M^G(U, f)$ can be defined as \begin{equation}\label{eq:integral} J_M^G(U,f) = \int_{U^G} f(x) w_{M,U}(x)\, dx \end{equation} for any $f\in C^{\infty}(G(\R))$ of almost compact support with $dx$ denoting the invariant measure on $U^G$, cf. \cite{Rao}. Let $Q=LN\in\mathcal{P}$ be a Richardson parabolic subgroup for ${\mathcal V}^G$. By the results of the previous section, if $f$ is conjugation invariant under the group $\operatorname{O}(n)$, we can write the above as \begin{equation}\label{realorbint} J_M^G(U,f) = c \int_{N(\R)} f(n) w_{M,U}(n)\, dn \end{equation} for some constant $c>0$. The function $w_{M, U}$ actually only depends on ${\mathcal V}$ and not on the specific $U\in \left({\mathcal V}(F)\right)_{M(\R)}$ so that we can also write $w_{M,{\mathcal V}}$ for this function. Note that the map $x\mapsto X=x-\operatorname{id}$ defines a bijection between the variety $\cU_G(\R)$ and the nilpotent cone in the Lie algebra $\mathfrak{g}(\R)$. Moreover, if $x\in U^G$, then for any $s\in \R$, $s\neq0$, the element \begin{equation}\label{eq:def:xs} x_s:=\operatorname{id}+ s(x-\operatorname{id}) \end{equation} is an element in ${\mathcal V}^G(\R)$ (but not necessarily in $U^G$). The goal is to show the following in this section: \begin{prop}\label{prop} Let $P=MV\in\mathcal{F}(M)$. Let $x\in {\mathcal V}^G(\R)$ be such that $w_{M,{\mathcal V}}(x)$ is defined. Then $w_{M,{\mathcal V}}(x_s)$ is also well-defined for every $s>0$, and as a function of $s$, it is a polynomial in $\log s$ of degree at most $\dim\mathfrak{a}_M^G$. Moreover, there are $k\in\N$, homogeneous polynomials $P_1, \ldots, P_t: N(\R)\longrightarrow \R^k$, and coefficients $\alpha_I\in\R$ for each subset $I\subseteq \{1,\ldots, t\}$ such that $w_{M,{\mathcal V}}(\operatorname{id}+X)=\sum_{I\subseteq \{1,\ldots, t\}} \alpha_I\prod_{i\in I} \log\|P_i(X)\|$ for $X$ in a dense subset of $N(\R)$. Here $\|\cdot\|$ denotes the vector norm on $\R^k$. \end{prop} To prove this proposition, we follow along the lines of Arthur's construction of the weight functions. Let $\Phi(A_M, G)$ be the set of roots of $(A_M, G)$, and let $\beta\in\Phi(A_M, G)$. Note that every root in $\Phi(A_M, G)$ is reduced in $\Phi(A_M, G)$, that is, if $\gamma\in\Phi(A_M, G)$, then $m\gamma\not\in\Phi(A_M, G)$ for any integer $m\neq\pm1$. Let $M_{\beta}\subseteq G$ be a Levi subgroup containing $M$ such that $\mathfrak{a}_{M_{\beta}}=\{X\in\mathfrak{a}_M\mid \beta(X)=0\}$. Let $P_{\beta}\in\mathcal{P}^{M_{\beta}}(M)$ be the unique parabolic subgroup of $M_{\beta}$ such that the unique root of $A_M$ on the unipotent radical of $P_\beta$ equals $\beta$. Suppose that $P$, $P_1\in\mathcal{P}(M)$, $P_1=MN_1$, are such that $P\cap M_{\beta}=P_{\beta}$ and $P_1\cap M_{\beta}=\overline{P_{\beta}}$ (the opposite parabolic), and write $P_{\beta}=L_{\beta}N_{\beta}$ for the Levi decomposition of $P_{\beta}$ with $M\subseteq L_{\beta}$. Suppose that $\pi=u\nu\in \mathcal{U}_M(\R) \overline{N_{\beta}}(\R)$. Then for any $a\in A_{M, \text{reg}}$ there is a unique $n_{\beta}\in \overline{N_{\beta}}(\R)$ such that \begin{equation}\label{eq:relation:unipotent} a\pi= n_{\beta}^{-1} a u n_{\beta}. \end{equation} Note that $n_{\beta}$ is independent of the $A_{M_{\beta}}$-part of $a$, that is, it only depends on $a^{\beta}$. Let $\operatorname{Wt}(\mathfrak{a}_M)\subseteq X(A_M)$ be the sublattice of all $\varpi$ which are extremal weights for some finite dimensional representation of $G(\R)$. Let $\varpi\in \operatorname{Wt}(\mathfrak{a}_M)\subseteq\mathfrak{a}_M^*$ be such that $\varpi(\beta^{\vee})>0$. Consider \[ v_P(\varpi, n_{\beta}):=e^{-\varpi(H_P(n_{\beta}))} =e^{-\varpi(H_{P_{\beta}}(n_{\beta}))}=v_{P_{\beta}}(\varpi,n_{\beta}) \] as a function of $a\in A_M/A_{M_{\beta}}\simeq A_M^{M_{\beta}}$, and $\pi$ as above. Write $\nu=\operatorname{id}+X$ and $n_{\beta}=\operatorname{id}+Y_{\beta}$ with $X,Y_\beta\in\overline{\mathfrak{n}_{\beta}}(\R)$. Note that $a^{-1}n_{\beta} a=\operatorname{id}+ a^{\beta} Y_{\beta}$. Further note that $Y_{\beta}^2=0$ since $2\beta\not\in\Phi(A_M, G)$. Hence \begin{align*} \pi= u+uX & = a^{-1} n_{\beta}^{-1} au n_{\beta} = (\operatorname{id} + a^{\beta} Y_{\beta})^{-1} u (\operatorname{id}+Y_{\beta})\\ & = u + (1-a^{\beta}) Y_{\beta} u + [u, Y_{\beta}] - a^{\beta} Y_{\beta} [u, Y_{\beta}] \end{align*} where $[u,Y_{\beta}]=uY_{\beta}-Y_{\beta} u$ is again nilpotent and contained in the Lie algebra of $\overline{N_{\beta}}(\R)$. Again, this implies that the term $Y_{\beta}[u,Y_{\beta}]$ vanishes since $2\beta$ is not a root of $(A_M, G)$. Hence \[ u+uX =u +(1-a^{\beta}) Y_{\beta} u + [u, Y_{\beta}]. \] Let $Q_0=M_0V_0$ be a semistandard minimal parabolic subgroup containing $N_1$, so in particular it contains $\overline{N_{\beta}}$ as well. Conjugating $u$ by some element in $\mathbf{K}^M:=\mathbf{K}\cap M(\R)$ if necessary, we can assume that $u\in V_0(\R)\cap M(\R)$. In particular, we can write $u=\operatorname{id}+X_0$ with $X_0$ a nilpotent matrix in the Lie algebra of $V_0(\R)\cap M(\R)$. Then $[u, Y_{\beta}]=[X_0, Y_{\beta}]$. Hence the above equality becomes \begin{equation}\label{eq:comparing:matrices} uX= X+X_0X = (1-a^{\beta}) Y_{\beta} (\operatorname{id}+X_0) + [X_0, Y_{\beta}] = (1-a^{\beta}) Y_{\beta} + X_0Y_{\beta} -a^{\beta}Y_{\beta}X_0 . \end{equation} $Q_0$ determines a choice of positive reduced roots $\Phi_{Q_0}^+:=\Phi(A_0,Q_0)$. Then there exists $\beta'\in\Phi_{Q_0}^+$ with $\beta'_{|A_M}=-\beta$, and we denote by $\Psi_{\beta}\subseteq \Phi_{Q_0}^+$ the subset of all such $\beta'$. Let $\Phi_{Q_0}^{M,+}\subseteq\Phi_{Q_0}^+$ denote the subset of positive roots on $M$. Then $\Psi_{\beta}\cap\Phi^{M,+}_{Q_0}=\emptyset$. We have a partial order $\prec_{\beta}$ on $\Psi_{\beta}$: If $\alpha_1, \alpha_2\in\Psi_{\beta}$, then $\alpha_1\prec_{\beta}\alpha_2$ if and only if there exists $\gamma\in\Phi_{Q_0}^+$ with $\alpha_2=\alpha_1+\gamma$. Note that then $\gamma\in\Phi_{Q_0}^{M,+} $. We define subsets of $\Psi_{\beta}$ according to how much the elements fail to be minimal with respect to $\prec_{\beta}$: Let $\Psi_{\beta}^{(0)}$ be the set of all $\alpha\in\Psi_{\beta}$ which are minimal with respect to $\prec_{\beta}$. If $k\geq0$ is a non-negative integer, let $\Psi_{\beta}^{(k+1)}$ be the set of all $\alpha\in\Psi_{\beta}\backslash\Psi_{\beta}^{(k)}$ which can be written as $\alpha=\alpha_1+\gamma$ with $\alpha_1\in\Psi_{\beta}^{(k)}$ and $\gamma\in\Phi^{M,+}_{Q_0}$. Note that $\Psi_{\beta}^{(k)}\neq\emptyset$ for only finitely many $k$. Moreover, $\alpha_1\in \Psi^{(l)}_\beta$ implies that $\alpha_2\prec_{\beta}\alpha_1$ for any $\alpha_2\in\Psi_{\beta}^{(k)}$ with $k<l$. To recover the matrix entries of $Y_{\beta}$ from~\eqref{eq:comparing:matrices} we now proceed inductively over $l$ by considering the matrix entries correspond to roots in $\Psi_{\beta}^{(l)}$. In the following we write \[ Z:=X+X_0X . \] Note that \[ \pi=\operatorname{id} +X_0+Z,\text{ so that } \pi_s=\operatorname{id} +sX_0+sZ. \] If $Y\in\mathfrak{g}(\R)$ is a matrix in the Lie algebra, and $\alpha\in\Phi(A_0,G)$ we denote by $Y^{(\alpha)}\in \R$ the matrix entry of $Y$ corresponding to $\alpha$. \begin{lem} Let $l\geq0$, and $\alpha\in\Psi_{\beta}^{(l)}$. There are rational polynomials $P_{\alpha, i}$, $0\leq i\leq l$ in the variables $X_0^{(\gamma)}$ ($\gamma\in\Phi_{Q_0}^{M,+}$), $Z^{(\gamma)}$ ($\gamma\in\Psi_{\beta}$), and $a^{\beta}$ such that for $a\neq1$ the following holds \begin{itemize} \item as a polynomial of $Z$ and $X_0$, $P_{\alpha,i}(Z, X_0, a^{\beta})$ is homogeneous of degree $(i+1)$ in the matrix entries of $Z+X_0$, so in particular, for every $s\in \R$ we have $P_{\alpha,i}(sZ, sX_0, a^{\beta})=s^{i+1}P_{\alpha,i}(Z, X_0, a^{\beta})$, \item and \[ Y_{\beta}^{(\alpha)}= \sum_{i=0}^{l} \frac{P_{\alpha,i}(Z, X_0, a^{\beta})}{(1-a^{\beta})^{i+1}}. \] \end{itemize} \end{lem} \begin{proof} We prove the lemma by induction on $l$. If $l=0$, then for any $\alpha\in\Psi_{\beta}^{(0)}$ we obtain from~\eqref{eq:comparing:matrices} that \[ Y_{\beta}^{(\alpha)} = \frac{ Z^{(\alpha)}}{(1-a^{\beta})} \] so the assertion of the lemma is true for $l=0$. Now suppose that for some non-negative integer $l\geq0$ we know that for every $0\leq k\leq l$ and every $\alpha\in \Psi_{\beta}^{(k)}$ we have \begin{equation}\label{eq:inductive:def:Y} Y_{\beta}^{(\alpha)} =\sum_{i=0}^{k}\frac{P_{\alpha,i}(Z,X_0, a^{\beta})}{(1-a^{\beta})^{i+1}} \end{equation} with $P_{\alpha, i}$ polynomials satisfying the assertions of the lemma. Then for $\alpha\in\Psi_{\beta}^{(k+1)}$ the equation~\eqref{eq:comparing:matrices} gives \[ Z^{(\alpha)} = (1-a^{\beta})Y_{\beta}^{(\alpha)} + (X_0Y_{\beta})^{(\alpha)} - a^{\beta} (Y_{\beta} X_0)^{(\alpha)}. \] By definition of $\Psi_{\beta}^{(k+1)}$ we have \[ (X_0Y_{\beta})^{(\alpha)} - a^{\beta} (Y_{\beta} X_0)^{(\alpha)} =\sum_{\substack{ \gamma\in \Psi_{\beta}^{(k)},~\delta\in\Phi_{Q_0}^{M,+}: \\ \gamma+\delta=\alpha}} e_{\alpha}^{\gamma,\delta} Y_{\beta}^{(\gamma)} X_{0}^{(\delta)} \] with \[ e_{\alpha}^{\gamma,\delta}= \begin{cases} 1 &\text{if }[E_{\delta},E_{\gamma}]=E_{\delta+\gamma},\\ -a^{\beta} &\text{if }[E_{\delta},E_{\gamma}]=-E_{\delta+\gamma} \end{cases} \] where $E_{\gamma}, E_{\delta}$ denote the elements of the standard Chevalley basis attached to our root system $\Phi_{Q_0}^+$. By the inductive assumption we can insert~\eqref{eq:inductive:def:Y} for $Y_{\beta}^{(\gamma)}$ for every $\gamma$ occurring in the sum. Dividing both sides of the so obtained equality by $(1-a^{\beta})$ then yields the assertion of the lemma. \end{proof} \subsection{Weight functions} We now consider the function $v_{P_{\beta}}(\varpi, n_{\beta})=e^{-\varpi(H_P(n_{\beta}))}$ for $\varpi\in\operatorname{Wt}(\mathfrak{a}_M)$. Note that $H_P$ is invariant under left and right multiplication with elements of $\mathbf{K}^M$. There is an irreducible representation $\Lambda_{\varpi}$ of $G$ on a finite dimensional vector space $V_{\varpi}$, defined over $\R$, together with an extremal vector $\phi_{\varpi}\in V_{\varpi}(\R)$ of weight $\varpi$ and a norm $\|\cdot\|$ on $V_{\varpi}(\R)$ such that $\|\phi_{\varpi}\|=1$ and \[ v_{P_{\beta}}(\varpi, n)= \|\Lambda_{\varpi}(n^{-1}) \phi_{\varpi}\|. \] We can identify $V_{\varpi}(\R)$ with $\R^{m_{\varpi}}$ for $m_{\varpi}=\dim V_{\varpi}$ and we can assume that the norm is of the form \[ \|(x_1,\ldots, x_{m_{\varpi}})\|= \left(x_1^2+\ldots+x_{m_{\varpi}}^2\right)^{1/2} \] for $(x_1,\ldots, x_{m_{\varpi}})\in \R^{m_{\varpi}}$. Recall that $\Lambda_\varpi$ is an algebraic representation. Thus $\Lambda_\varpi\colon G\to\operatorname{GL}(V_\varpi)$ is a morphism of algebraic varieties. Hence as a function of $n_{\beta}\in \overline{N_{\beta}}(\R)$ the function $v_{P_{\beta}}(\varpi, n_{\beta})$ is the vector norm applied to a polynomial function from $\overline{\mathfrak{n}_{\beta}}(\R)$ to $\R^{m_\varpi}$. Using the above lemma (and the notation therein), we therefore get that \begin{equation}\label{eq:weight} v_{P_{\beta}}(\varpi, n_{\beta})^2= \sum_{\kappa\in R} \frac{f_{\kappa}(Z, X_0, a^{\beta})}{(a^{\beta}-1)^{\kappa}} \end{equation} where $\kappa$ runs over a finite set of integers $R\subseteq\Z$, and $f_{\kappa}$ is the norm of some rational function that is homogeneous of degree $\kappa$ in the matrix entries of $X_0$ and $Z$, and has a finite value at $a=1$ so that we may write $f_{\kappa}(Z, X_0, 1)$. Here $n_\beta$, $Z$, $X_0$, and $a$ are related as in \eqref{eq:relation:unipotent}. The function $f_{\kappa}(Z, X_0, a^{\beta})$ in general depends on $\varpi$ and if we want to make this dependence explicit we write $f_{\kappa, \varpi}(Z, X_0, a^{\beta})$. Now let $u\in U$, $U\subseteq M(\R)$, and $U^G\subseteq G(\R)$ be as before, and write $u=\operatorname{id} +X_0$ with $X_0$ nilpotent. Let $\kappa_0(\beta,X_0)\in R$ be the largest $\kappa\in R$ such that $f_{\kappa}(\cdot, \cdot,a^{\beta})$ does not vanish identically on $\overline{\mathfrak{n}_{\beta}}(\R)\times \mathcal{N}$, where $\mathcal{N}\subseteq \mathfrak{m}(\R)$ is the nilpotent orbit defined by $U=\operatorname{id}+\mathcal{N}$. Let $\rho(\beta, X_0)$ be the product of $1/(2\varpi(\beta^{\vee}))$ with $\kappa_0(\beta, X_0)$. It follows from~\cite[p. 238]{Ar5} that $\kappa_0(\beta,X_0)\ge0$ and $\rho(\beta, X_0)\ge0$. The $\rho(\beta,X_0)$ is independent of $\varpi$ as explained in \cite[p. 238]{Ar5} but $\kappa_0(\beta, X_0)$ in general depends on $\varpi$. If we want to emphasize this dependence, we write $\kappa_0(\beta, X_0)=\kappa_0(\varpi, \beta, X_0)$. Recall the definition of the weight function $w_P(\lambda, a, \pi)$ from~\cite[(3.6)]{Ar5}: Fix a parabolic subgroup $P_1\in\mathcal{P}(M)$. Then for any other $P\in\mathcal{P}(M)$ and any $P$-dominant $\varpi\in \operatorname{Wt}(\mathfrak{a}_M)$ Arthur defines for $\pi=u\nu\in U N_1(\R)$ the function \begin{equation}\label{eq:weight:function:definition} w_P(\varpi, a, \pi) =\left(\prod_{\beta\in \Phi_P\cap\Phi_{\overline{P_1}}} r_{\beta}(\varpi,u,a)\right) v_P(\varpi,n) \end{equation} where $\Phi_P=\Phi(A_M, P)$ and $\Phi_{\overline{P_1}}=\Phi(A_M, \overline{P_1})$. Here $u$, $\nu$, $a$, and $n\in N_1(\R)$ are related by $a\pi=n^{-1} au n$. The function $r_{\beta}$ is given by (see~\cite[(3.4)]{Ar5}) \[ r_{\beta}(\varpi, u, a) =|a^{\beta}-a^{-\beta}|^{\rho(\beta,X_0)\varpi(\beta^{\vee})}. \] It follows from this definition that \begin{equation}\label{eq:weight:function:invariance:P} w_P(\varpi, a, k^{-1}\pi k) = w_P(\varpi, a, \pi) \end{equation} for $k=\operatorname{diag}(-1,1, \ldots, 1)\in \bar\mathbf{K}^M_\infty$. If $\varpi_1,\ldots, \varpi_r\in\operatorname{Wt}(\mathfrak{a}_M)$ is a basis of $\mathfrak{a}_M^*$ consisting of $P$-dominant weights, and $\lambda=\lambda_1\varpi_1+\ldots+\lambda_r\varpi_r\in\mathfrak{a}_{M,\C}^*$ with $\lambda_1,\ldots,\lambda_r\in\C$, then \begin{equation}\label{eq:product:weight:functions} w_P(\lambda, a, \pi)=\prod_{i=1}^r w_P(\varpi_i, a, \pi)^{\lambda_i} \end{equation} By~\cite[Lemma 4.1]{Ar5} the limit \[ w_P(\lambda, \pi):=\lim_{a\rightarrow 1}w_P(\lambda, a, \pi) \] exists and is non-zero for all $\pi$ in an open and dense subset of $U N_1(\R)$. \subsection{The adjacent case} For a unipotent element $\pi$ recall the definition of $\pi_s$ from~\eqref{eq:def:xs}. If $P$ and $P_1$ are adjacent, the function $w_P(\lambda, \pi)$ has the following behavior when $\pi$ is replaced by $\pi_s$: \begin{lem}\label{lemma:homog:weight:adjacent} Suppose $P$ and $P_1$ are adjacent via $\beta\in\Phi(A_M,P)$. Then for all $\pi$ in an open dense subset of $U N_1(\R)$ we have \[ w_P(\lambda, \pi_s) = s^{\rho(\beta, X_0) \lambda(\beta^\vee)}\cdot w_P(\lambda,\pi) \] for all $s>0$. \end{lem} \begin{proof} Note that $w_P(\lambda, a, \pi)=w_P(\lambda, a, k^{-1}\pi k)$ for any $k\in \mathbf{K}^M=\mathbf{K}\cap M(\R)$ so that we can assume $\pi$ to be of the form $\operatorname{id}+X_0+Z$ with $X_0\in\mathcal{N}\cap \mathfrak{v}_0(\R)$, and $Z\in \mathfrak{n}_1(\R)$ as before. Here $\mathfrak{v}_0$, resp.\ $\mathfrak{n}_1$, denotes the Lie algebra of $V_0$ (the unipotent radical of $Q_0$), resp.\ $N_1$ (the unipotent radical of $P_1$). Since $P$ and $P_1$ are assumed to be adjacent along $\beta$, we have $\Phi_P\cap \Phi_{\bar P_1}=\{\beta\}$. Hence it follows from \eqref{eq:weight:function:definition} that $ w_P(\lambda, a, \pi) = r_{\beta}(\lambda, X_0, a) v_{P}(\lambda, n)$ where \[ r_{\beta}(\lambda, X_0, a)=|a^{\beta}-a^{-\beta}|^{\rho(\beta,X_0)\lambda(\beta^{\vee})} \]. for $a$ with $a^{\beta}\neq1$. We can write $n\in N_1$ as $n=n_\beta \tilde n$ with $n_\beta$ in the unipotent radical of $\overline{P_\beta}$ and $\tilde n$ in the unipotent radical of $P_1\cap P$. Then $ v_{P}(\lambda, n) = v_{P_\beta}(\lambda, n_\beta)$ so that \[ w_P(\lambda, a, \pi) = r_{\beta}(\lambda, X_0, a) v_{P_{\beta}}(\lambda, n_\beta). \] Hence, using \eqref{eq:weight} and \eqref{eq:product:weight:functions}, \begin{align*} w_P(\lambda, \pi)^2 & = \lim_{a\rightarrow1}\prod_{i=1}^r\left(|a^{\beta}-a^{-\beta}|^{2\rho(\beta,X_0)\varpi_i(\beta^{\vee})} \sum_{\kappa\in R} \frac{f_{\kappa, \varpi_i}(Z, X_0, a^{\beta})}{(a^{\beta}-1)^{\kappa}}\right)^{\lambda_i} \\ & = \lim_{a\rightarrow1}\prod_{i=1}^r\left(|a^{\beta}-a^{-\beta}|^{\kappa_0(\varpi_i, \beta, X_0)} \sum_{\kappa\in R} \frac{f_{\kappa, \varpi_i}(Z, X_0, a^{\beta})}{(a^{\beta}-1)^{\kappa}}\right)^{\lambda_i} \\ & = \prod_{i=1}^ r 2^{\lambda_i} f_{\kappa_0(\varpi_i,\beta, X_0), \varpi_i}(Z, X_0, 1)^{\lambda_i}. \end{align*} Now $f_{\kappa_0(\varpi_i,\beta, X_0), \varpi_i}(Z, X_0, 1)$ is non-zero for generic $X_0$, $Z$, and satisfies \[ f_{\kappa_0(\varpi_i,\beta, X_0), \varpi_i} (sZ, sX_0, 1) = s^{\kappa_0(\varpi_i, \beta,X_0)}f_{\kappa_0(\varpi_i,\beta, X_0), \varpi_i} (Z, X_0, 1). \] Hence \[ w_P(\lambda, \pi_s)^2 =\prod_{i=1}^r s^{\lambda_i \kappa_0(\varpi_i, \beta,X_0)} \prod_{i=1}^r 2^{\lambda_i} f_{\kappa_0(\varpi_i,\beta, X_0), \varpi_i}(Z, X_0, 1)^{\lambda_i} = s^{2\rho(\beta, X_0) \lambda(\beta^\vee)} w_P(\lambda, \pi)^2 \] since $\sum_{i=1}^r \lambda_i \kappa_0(\varpi_i, \beta,X_0)= 2 \rho(\beta, X_0) \sum_{i=1}^r \lambda_i \varpi_i(\beta^\vee)= 2\rho(\beta, X_0) \lambda(\beta^\vee)$. Since $w_P(\lambda, \pi_s)$ is a real-valued function and continuous in $s$ with $w_P(\lambda, \pi_1)=w_P(\lambda, \pi)$, we can take the square-root on both sides of the equation and obtain the assertion of the lemma. \end{proof} \subsection{The general case} If $Q, Q'\in \mathcal{P}(M)$ are adjacent along some root $\beta\in\Phi(A_M,Q)$, we write $Q|_{\beta}Q'$. \begin{corollary} Suppose that $P$ and $P_1$ are not necessarily adjacent. Choose a minimal chain $P=Q_0|_{\beta_1}Q_1|_{\beta_2}\ldots|_{\beta_t} Q_t=P_1$ of adjacent parabolic subgroups $Q_1, \ldots, Q_t\in\mathcal{P}(M)$ from $P$ to $P_1$. Then there exist rational numbers $\rho_1, \ldots, \rho_t$ such that for all $\pi$ in an open dense subset of $U N_1(\R)$ we have \[ w_P(\lambda, \pi_s) = s^{\rho_1 \lambda(\beta_1^{\vee})+\ldots+\rho_t\lambda(\beta_t^{\vee})}\cdot w_P(\lambda,\pi) \] for all $s>0$. \end{corollary} \begin{proof} This follows by induction on $t$ together with Lemma~\ref{lemma:homog:weight:adjacent} and the proof of~\cite[Lemma 4.1]{Ar5}. The case $t=1$ is covered in the last lemma. Let $P_1':=Q_1=MN_1'$, and assume that the corollary is true for $P$ replaced by $P_1'$. Let $\pi= u\nu\in U N_1(\R)$, and $a\in A_{M, \text{reg}}$. Let $n\in N_1(\R)$ be the unique element with $\pi=a^{-1}n^{-1}aun$. Write $n=m'n'k'$ with $m'\in M(\R)$, $n'\in N_1'(\R)$, and $k'\in \mathbf{K}$, and put $u'=(m')^{-1} u m'$. Let $\pi'=a^{-1} (n')^{-1} a u' n'\in U N_1'(\R)$. Then by the proof of~\cite[Lemma 4.1, p. 241]{Ar5}, we have \[ w_P(\lambda,a,\pi)=w_P(\lambda,a,\pi')w_{P_1'}(\lambda,a,\pi). \] Suppose that $\pi=\operatorname{id}+ Y$. Then $\pi'=\operatorname{id} + k'Yk^{\prime-1}$ so that that the map $\pi\mapsto \pi'$ also maps $\pi_s$ to $\pi'_s $ for any $s$. We are further allowed to take the value at $a= 1$ on both sides because of~\cite[Lemma 4.1]{Ar5}. The assertion of the corollary therefore follows from Lemma~\ref{lemma:homog:weight:adjacent} and the induction hypothesis. \end{proof} Arthur defines polynomials $W_P(\varpi,a,\pi)\in V_{\varpi}$ for a $P$-dominant weight $\varpi\in \operatorname{Wt}(\mathfrak{a}_M)$ and $(a,\pi)\in A_{M,\text{reg}}\times U N_1(\R)$ by \[ W_P(\varpi,a,\pi) =\left(\prod_{\beta\in \Phi_P\cap\Phi_{\overline{P_1}}} r_{\beta}(\varpi,u,a)\right) \Lambda_{\varpi}(n^{-1}) \phi_{\varpi} \] so that $ w_P(\varpi, a, \pi) =\|W_P(\varpi,a,\pi)\|$ and \begin{equation}\label{eq:weight:as:prod} w_P(\lambda, a, \pi) =\prod_{i=1}^r \|W_P(\varpi_i,a,\pi)\|^{\lambda_i}. \end{equation} Here $\pi$, $u$, $a$, and $n$ are related as explained after \eqref{eq:weight:function:definition} \begin{corollary}\label{cor:homog:pol} If $\varpi\in\operatorname{Wt}(\mathfrak{a}_M)$ is a $P$-dominant weight, the polynomial $W_P(\varpi,a,\pi)$ is defined on all of $A_M\times U N_1$ and does not vanish at $a=1$. Moreover, there is a constant $r_{\varpi}$ depending only on $P$ and $\varpi$ such that for all $\pi\in U N_1(\R)$ and all $s>0$ we have \[ \|W_P(\varpi,1,\pi_s)\|=s^{r_{\varpi}} \|W_P(\varpi,1,\pi)\|. \] \end{corollary} \begin{proof} All assertions except the homogeneity are subject of~\cite[Corollary 4.3]{Ar5}. The homogeneity follows from the previous corollary and the definition of $W_P(\varpi,a,\pi)$. \end{proof} Let $\theta_P(\lambda)= v_P^{-1}\prod_{\alpha\in\Delta_P} \lambda(\alpha^{\vee})$ for $\lambda\in \mathfrak{a}_M^*$, where $v_P$ denotes the covolume of the lattice spanned by all $\alpha^{\vee}$, $\alpha\in\Delta_P$, in $\mathfrak{a}_M$. Then $\{w_P(\lambda,a,\pi)\}_{P\in\mathcal{P}(M)}$ defines a $(G,M)$-family, and one can attach a certain number to this family by defining \begin{equation}\label{eq:definition:weight} w_M(a,\pi)=\lim_{\lambda\rightarrow 0} \left(\sum_{P\in\mathcal{P}(M)} w_P(\lambda,a,\pi)\theta_P(\lambda)^{-1}\right) \end{equation} as in~\cite[\S 6]{Ar5}. By~\cite[(6.5)]{Ar3} this can be computed by \[ w_M(a,\pi) =\frac{1}{r!} \sum_{P\in\mathcal{P}(M)}\left(\lim_{t\rightarrow 0}\frac{d^r}{dt^r} w_P(t\Lambda,a,\pi) \right)\theta_P(\Lambda)^{-1} \] with $r=\dim\mathfrak{a}_M^G$ and $\Lambda\in\mathfrak{a}_M^*$ some fixed generic element. (Note that $w_M(a,\pi)$ is independent of the choice of $\Lambda$.) Using~\eqref{eq:weight:as:prod} we get (cf.~\cite[Lemma 5.4]{Ar5}) \[ w_M(a,\pi) = \sum_{\Omega} c_{\Omega} \prod_{(P,\varpi)\in\Omega} \log\|W_P(\varpi,a,\pi)\| \] where $\Omega$ runs over all finite multisets consisting of elements in $\mathcal{P}(M)\times \operatorname{Wt}(\mathfrak{a}_M)$, and the $c_{\Omega}\in\C$ are suitable coefficients which vanish for all but finitely many $\Omega$. Moreover, each $\Omega$ contains at most $r=\dim\mathfrak{a}_M^G$ many elements. Note that in a neighborhood of $a=1$ this expression is well-defined for all $\pi$ in an open dense subset of $U N_1$ because of Corollary~\ref{cor:homog:pol}. Hence we can evaluate $w_M(a,\pi)$ at $a=1$ by means of this expression. Moreover, Corollary~\ref{cor:homog:pol} implies the following result. \begin{corollary}\label{cor:homog:weight} For all $\pi$ in an open dense subset of $U N_1(\R)$ and all $s>0$ \[ w_M(1,\pi_s) = \sum_{\Omega} c_{\Omega} \prod_{(P,\varpi)\in\Omega}(r_{\varpi}\log s+ \log\|W_P(\varpi,1,\pi)\|). \] In particular, as a function of $s$, $w_M(1,\pi_s)$ is a polynomial in $\log s$ of degree at most $r=\dim\mathfrak{a}_M^G$. \end{corollary} If now $Q'\in \mathcal{F}(M)$ is an arbitrary subgroup, we can analogously define all the above functions with respect to the Levi component $M_{Q'}$ instead of $G$, in particular we can define the analogue of $w_M(1,\pi)$ which we denote by $w_M^{Q'}(1,\pi)$. The corollary then stays true for $w_M^{Q'}(1,\pi)$ with the necessary changes. Note that a priori $w_M^{Q'}(1,\pi)$ is defined for $\pi$ in a dense open subset of $U N_1^{M_{Q'}}$ for $P_1^{M_{Q'}}=P_1\cap M_{Q'}$. However, we can trivially extend $w_M^{Q'}(1,\pi)$ to a dense open subset of $U^G$. By the first equation on \cite[p. 256]{Ar5} the weight function $w_{M,{\mathcal V}}$ from~\eqref{eq:integral} can then be written as \[ w_{M,{\mathcal V}}(\pi) = \sum_{Q'\in \mathcal{F}(M)} w_M^{Q'}(1,\pi). \] This together with Corollary~\ref{cor:homog:weight} implies the Proposition~\ref{prop}. \subsection{Convergence} \begin{lem}\label{logconv} Let $p_i:\R^k\longrightarrow \R$, $i=1,...,l$, be homogeneous polynomials and put \begin{equation}\label{eq:log:homog:polynom} \lambda(x):=\prod_{i=1}^l\big|\log|p_i(x)|\big|,\quad x\in\R^k. \end{equation} Then for every $a>0$ \begin{equation}\label{eq:log:int2} \int_{\R^k}e^{-a\left(\log(1+\|x\|)\right)^2}\lambda(x)\;dx< \infty. \end{equation} \end{lem} \begin{proof} Since $\prod_{i=1}^l\big|\log|p_i(x)|\big|\le \sum_{i=1}^l \big|\log|p_i(x)|\big|^l$, it suffices to consider the case $\lambda(x)=\big|\log|p(x)|\big|^l$ with $p:\R^k\longrightarrow\R$ a homogeneous polynomial of degree $\kappa$. Further note that for every $M\in\N$ there exists $r_M>0$ such that $\left(\log(1+\|x\|)\right)^2\ge (M/a) \log \|x\|$ for all $x$ with $\|x\|\ge r_M$. In particular, \begin{equation}\label{eq:polynomial:decay} e^{-a\left(\log(1+\|x\|)\right)^2} \le \|x\|^{-M} \end{equation} for all $x$ with $\|x\|\ge r_M$. We decompose the integral in~\eqref{eq:log:int2} into a sum of integrals over certain subsets of $\R^k$ similarly as in~\cite[\S 7]{Ar5}. For each $m\in\N_0$ let \[ B_m=\{x\in\R^k\mid 2^m\le \|x\|\le 2^{m+1}\} \] and let $B=\{x\in\R^k\mid \|x\|\le 2\}$. For $\epsilon>0$ set \[ \Gamma_{\epsilon}=\{x\in B\mid |p(x)|<\epsilon\}, \text{ and } \Gamma_{m,\epsilon}=\{x\in B_m\mid |p(x)|<\epsilon\}. \] By \cite[(7.1)]{Ar5} there are constants $C, t>0$ such that for every $\epsilon>0$ we have \begin{equation}\label{eq:integral:log} \int_{\Gamma_\epsilon} \lambda(x)\, dx \le C \epsilon^t. \end{equation} For $m\ge0$ we can therefore compute, using that $p$ is homogeneous of degree $\kappa$, that \begin{align*} \int_{\Gamma_{m,\epsilon}} \lambda(x)\, dx &\le 2^{mk}\int_{\Gamma_{2^{-m\kappa}\epsilon}} \lambda(2^m y)\, dy \le c_1 2^{mk} m^l \int_{\Gamma_{2^{-m\kappa}\epsilon}} (1+\lambda( y))\, dy \le c_2 2^{mk} m^l 2^{-m\kappa t_1}\epsilon^{t_1} \\ & \le c_3 2^{m (k+l)} \epsilon^{t_1} \end{align*} where $c_1, c_2, c_3, t_1>0$ are suitable constants independent of $m$ and $\epsilon$ and the second last inequality follows from \eqref{eq:integral:log}. Fix $\epsilon>0$ and let $Z_\epsilon=\{x\in\R^k\mid |p(x)|<\epsilon\}$. Then, using \eqref{eq:polynomial:decay} for every $M\in\N$ there exist $c_4, r_M>0$ such that \begin{align*} & \int_{Z_\epsilon}e^{-a\left(\log(1+\|x\|)\right)^2}\lambda(x)\;dx \\ \le & \int_{\Gamma_\epsilon} \lambda(x)\, dx + \sum_{m\le \log_2 r_M} c_4\int_{\Gamma_{m,\epsilon}} \lambda(x)\, dx + \sum_{m\ge \log_2 r_M} 2^{-mM}\int_{\Gamma_{m,\epsilon}} \lambda(x)\, dx \\ \le & C\epsilon ^t + c_5 \epsilon^{t_1}\left( 1+ \sum_{m\ge\log_2 r_M} 2^{-mM} 2^{m(k+l)}\right). \end{align*} This is finite if we choose $M$ sufficiently large. Now on $\R^k\backslash Z_\epsilon$ the polynomial $p(x)$ is bounded away from $0$ so that $\log|p(x)|$ is bounded from below on $\R^k\backslash Z_\epsilon$. Since $p(x)$ is of degree $\kappa$, there is a constant $A>0$ such that $|p(x)|\le A (1+\|x\|)^\kappa$ for all $x\in \R^k$. Choose $M$ again sufficiently large and let $r_M$ be as in \eqref{eq:polynomial:decay}. Then for some suitable constant $A_1>0$ we get \[ \int_{\R^k\backslash Z_\epsilon} e^{-a\left(\log(1+\|x\|)\right)^2} \lambda(x)\, dx \le A_1 \int_{\R^k\backslash Z_\epsilon} \|x\|^{-M} \left(1+(\log(1+\|x\|))^l\right)\, dx, \] which is finite if $M$ was chosen sufficiently large. \end{proof} \section{Examples for weight functions in low rank}\label{sec-lowrank} \setcounter{equation}{0} \subsection{$G=\operatorname{GL}(2)$} There are two unipotent conjugacy classes in $\operatorname{GL}(2)$, the trivial class for which Richardson parabolic subgroup equals $G$, and the regular unipotent conjugacy class with Richardson parabolic equal to the minimal parabolic subgroup $P_0=M_0U_0$. The archimedean orbital integrals appearing in the fine expansion of $J_{\operatorname{unip}}$ are $J_G(1,f_{\infty})$, $J_G(u_0,f_{\infty})$, and $J_{M_0}(1, f_{\infty})$, where $u_0=\left(\begin{smallmatrix} 1& 1\\ 0&1\end{smallmatrix}\right)$ represents the regular class in $G(\Q)$. The first two integrals are unweighted, and the last integral $J_{M_0}(1, f_0)$ is up to a normalization of Haar measure equal to \[ \int_{U_0(\R)} f_{\infty}(\left(\begin{smallmatrix} 1& x\\ 0&1\end{smallmatrix}\right)) \log |x| \, dx, \] see, e.g., \cite{Gelbart}. \subsection{$G=\operatorname{GL}(3)$} There are three unipotent conjugacy class in $\operatorname{GL}(3)$: The trivial, the regular, and the subregular class. Let \[ u(x,y,z)=\left(\begin{smallmatrix} 1&x&y\\ \, & 1&z\\\,&\,& 1\end{smallmatrix}\right) \] for $x, y, z\in\R$. Then $u(0,1,0)$ is a representative for the subregular class, and $u(1,0,1)$ a representative for the regular class. Let $M_1$ be the Levi subgroup corresponding to the partition $(2,1)$ of $3$. Every other corank-$1$ Levi subgroup in ${\mathcal L}$ is conjugate to $M_1$ by some Weyl group element so that it suffices to consider $J_M(u,f_{\infty})$ for $M\in\{M_0, M_1, G\}$. The integrals $J_G(1,f_{\infty})$, $J_G(u(0,1,0), f_{\infty})$, and $J_G(u(1,0,1), f_{\infty})$ are all unweighted. For the other cases we get (up to normalization of the measures) \begin{equation}\label{unipot3} J_{M_1}(1, f_{\infty})=\int_{U_1(\R)} f_{\infty}(u(0,y,z)) \log(y^2+z^2) \,du(0,y,z), \end{equation} \begin{equation}\label{unipot4} J_{M_1}(u(1,0,0), f_{\infty}) =\int_{U_0(\R)} f_{\infty}(u(x,y,z)) \log|xz| \, du(x,y,z), \end{equation} and \begin{equation}\label{unipot5} J_{M_0}(1, f_{\infty}) =\int_{U_0(\R)} f_{\infty}(u(x,y,z)) \left(\log|x|\log|z|+ (\log|x|)^2+ (\log|z|)^2\right) \, du(x,y,z), \end{equation} cf.\ \cite[p. 67, Lemma 4]{Flicker}. \section{Bochner Laplace operators}\label{sec-bochlapl} \setcounter{equation}{0} In this section we summarize some basic facts about Bochner-Laplace operators on global Riemannian symmetric spaces. For simplicity we assume that $G$ is semisimple and $G(\R)$ is of noncompact type. Then $G(\R)$ is a semisimple real Lie group of noncompact type. Let $K_\infty\subset G(\R)$ be a maximal compact subgroup and \[ \widetilde X=G(\R)/K_\infty \] the associated Riemannian symmetric space. Let $\Gamma\subset G(\R)$ be a torsion free lattice and let $X=\Gamma\bs\widetilde X$. Let $\nu$ be a finite-dimensional unitary representation of $K_\infty$ on $(V_{\nu},\left<\cdot,\cdot\right>_{\nu})$. Let \begin{align*} \widetilde{E}_{\nu}:=G(\R)\times_{\nu}V_{\nu} \end{align*} be the associated homogeneous vector bundle over $\tilde{X}$. Then $\left<\cdot,\cdot\right>_{\nu}$ induces a $G(\R)$-invariant metric $\tilde{h}_{\nu}$ on $\tilde{E}_{\nu}$. Let $\widetilde{\nabla}^{\nu}$ be the connection on $\tilde{E}_{\nu}$ induced by the canonical connection on the principal $K_\infty$-fibre bundle $G(\R)\to G(\R)/K_\infty$. Then $\widetilde{\nabla}^{\nu}$ is $G(\R)$-invariant. Let \begin{align*} E_{\nu}:=\Gamma\bs \widetilde E_\nu \end{align*} be the associated locally homogeneous vector bundle over $X$. Since $\tilde{h}_{\nu}$ and $\widetilde{\nabla}^{\nu}$ are $G(\R)$-invariant, they push down to a metric $h_{\nu}$ and a connection $\nabla^{\nu}$ on $E_{\nu}$. Let $C^\infty(\widetilde X,\widetilde E_\nu)$ resp. $C^\infty(X,E_\nu)$ denote the space of smooth sections of $\widetilde E_\nu$, resp. $E_\nu$. Let \begin{align}\label{globsect} \begin{split} C^{\infty}(G(\R),\nu):=\{f:G(\R)\rightarrow V_{\nu}\colon f\in C^\infty,\: f(gk)=&\nu(k^{-1})f(g),\\ &\forall g\in G(\R), \,\forall k\in K_\infty\}, \end{split} \end{align} Let $L^2(G(\R),\nu)$ be the corresponding $L^2$-space. There is a canonical isomorphism \begin{equation}\label{iso-glsect} \widetilde A\colon C^\infty(\widetilde X,\widetilde E_\nu)\cong C^\infty(G(\R),\nu) \end{equation} (see \cite[p. 4]{Mia}). $\widetilde A$ extends to an isometry of the corresponding $L^2$-spaces. Let \begin{align}\label{globsect1} C^{\infty}(\Gamma\backslash G(\R),\nu):=\left\{f\in C^{\infty}(G(\R),\nu)\colon f(\gamma g)=f(g)\:\forall g\in G(\R), \forall \gamma\in\Gamma\right\} \end{align} and let $L^2(\Gamma\bs G(\R),\nu)$ be the corresponding $L^2$-space. The isomorphism \eqref{iso-glsect} descends to isomorphisms \begin{equation}\label{iso-glsect1} A\colon C^{\infty}(X,E_{\nu})\cong C^{\infty}(\Gamma\backslash G(\R),\nu),\quad L^2(X,E_\nu)\cong L^2(\Gamma\bs G(\R),\nu). \end{equation} Let $\widetilde{\Delta}_{\nu}={\widetilde{\nabla^\nu}}^{*}{\widetilde{\nabla}}^{\nu}$ be the Bochner-Laplace operator of $\widetilde{E}_{\nu}$. This is a $G(\R)$-invariant second order elliptic differential operator whose principal symbol is given by \[ \sigma_{\widetilde\Delta_\nu}(x,\xi)=\|\xi\|^2_x\cdot\operatorname{Id}_{E_{\nu,x}},\quad x\in\widetilde X, \;\xi\in T^\ast_x(\widetilde X). \] Since $\widetilde{X}$ is complete, $\widetilde{\Delta}_{\nu}$ with domain the smooth compactly supported sections is essentially self-adjoint \cite[p. 155]{LM}. Its self-adjoint extension will be denoted by $\widetilde{\Delta}_\nu$ too. Let $\Omega\in\cZ(\gf_\C)$ and $\Omega_{K_\infty}\in\cZ(\kf)$ be the Casimir operators of $\gf$ and $\kf$, respectively, where the latter is defined with respect to the restriction of the normalized Killing form of $\mathfrak{g}$ to $\mathfrak{k}$. Let $C^\infty(G(\R),V_\nu)$ be the space of smooth $V_\nu$-valued functions on $G(\R)$ and $R$ the right regular representation of $G(\R)$ in in $C^\infty(G(\R),V_\nu)$. Let $R(\Omega)$ be the differential operator induced by $\Omega$. Since $\operatorname{Ad}(g)\Omega=\Omega$, $g\in G(\R)$, it follows that $R(\Omega)$ preserves the subspace $C^\infty(G(\R),\nu)$. Then with respect to the isomorphism \eqref{iso-glsect} we have \begin{align}\label{BLO} \widetilde{\Delta}_{\nu}=-R(\Omega)+\nu(\Omega_{K_\infty}), \end{align} (see \cite[Proposition 1.1]{Mia}). Let $e^{-t\widetilde\Delta_{\nu}}$, $t>0$, be the heat semigroup generated by $\widetilde\Delta_\nu$. It commutes with the action of $G(\R)$. With respect to the isomorphism \eqref{iso-glsect} we may regard $e^{-t\widetilde\Delta_{\nu}}$ as bounded operator in $L^2(G(\R),\nu)$, which commutes with the action of $G(\R)$. Hence it is a convolution operator, i.e., there exists a smooth map \begin{equation}\label{heatker1} H_t^\nu\colon G(\R)\to \operatorname{End}(V_\nu) \end{equation} such that \[ (e^{-t\widetilde\Delta_{\nu}}\phi)(g)=\int_{G(\R)} H_t^\nu(g^{-1}g^\prime)(\phi(g^\prime))\; dg^\prime,\quad \phi\in L^2(G(\R),\nu). \] The kernel $H_t^\nu$ satisfies \begin{align}\label{propH} {H}^{\nu}_{t}(k^{-1}gk')=\nu(k)^{-1}\circ {H}^{\nu}_{t}(g)\circ\nu(k'), \:\forall k,k'\in K, \forall g\in G. \end{align} For $q>0$ let $\ccC^q(G(\R))$ be Harish-Chandra's $L^q$-Schwartz space. We briefly recall its definition. Let $\Xi$ and $\|\cdot\|$ be the functions on $G(\R)$ used to define Harish-Chandra's Schwartz space $\Co(G(\R))$ (see \cite[7.1.2]{Wal}). Furthermore, for $Y\in\cU(\gf_\C)$ denote by $L(Y)$ (resp. $R(Y)$) the associated left (resp. right) invariant differential operator on $G(\R)$. Then $\ccC^q(G(\R))$ consists of all $f\in C^\infty(G(\R))$ such that \[ \sup_{x\in G(\R)}(1+\|x\|)^m\Xi(x)^{-2/q}|L(Y_1)R(Y_2)f(x)|<\infty \] for all $m\ge 0$ and $Y_1,Y_2\in\cU(\gf_\C)$. Note that $\ccC^2(G(\R))$ equals Harish-Chandra's Schwartz space $\Co(G(\R))$. Proceeding as in the proof of \cite[Proposition 2.4]{BM} it follows that $H^\nu_t$ belongs to $(\ccC^{q}(G(\R))\otimes\operatorname{End}(V_{\nu}))^{K_\infty\times K_\infty}$ for all $q>0$. Let $\pi$ be a unitary representation of $G(\R)$ on a Hilbert space $\cH_\pi$. Define a bounded operator on $\cH_{\pi}\otimes V_{\nu}$ by \begin{align}\label{Definiton von pi(tilde(H))} \tilde{\pi}(H^{\nu}_{t}(g)):=\int_{G(\R)}\pi(g)\otimes H^{\nu}_{t}(g)\;dg. \end{align} Then relative to the splitting \begin{align*} \mathcal{H}_{\pi}\otimes V_{\nu}=\left(\mathcal{H}_{\pi}\otimes V_{\nu}\right)^{K_\infty}\oplus\left(\left(\mathcal{H}_{\pi}\otimes V_{\nu}\right)^{K_\infty}\right)^{\bot}, \end{align*} $\tilde{\pi}(H^{\nu}_{t})$ has the form \begin{align*} \begin{pmatrix} \pi({H}^{\nu}_{t})&0\\0&0 \end{pmatrix}, \end{align*} where $\pi(H^{\nu}_{t})$ acts on $\left(\mathcal{H}_{\pi}\otimes V_{\nu}\right)^{K_\infty}$. Assume that $\pi$ is irreducible. Let $\pi(\Omega)$ be the Casimir eigenvalue of $\pi$. Then as in \cite[Corollary 2.2]{BM} it follows from \eqref{BLO} that \begin{equation}\label{integop} \pi(H_t^\nu)=e^{t(\pi(\Omega)-\nu(\Omega_{K_\infty}))}\operatorname{Id}, \end{equation} where $\operatorname{Id}$ is the identity on $\left(\mathcal{H}_{\pi}\otimes V_{\nu}\right)^{K_\infty}$. Put \begin{equation}\label{loctrace} h_t^\nu(g):=\operatorname{tr} H_t^\nu(g),\quad g\in G(\R). \end{equation} Then $h_t^\nu\in\ccC^q(G(\R))$ for all $q>0$. In particular, $h_t^\nu$ belongs to $\ccC^2(G(\R))$, which equals Harish-Chandra's Schwartz space $\Co(G(\R))$. Let $\pi$ be a unitary representation of $G(\R)$. Put \[ \pi(h_t^\nu)=\int_{G(\R)}h_t^\nu(g)\pi(g)\; dg. \] Assume that $\pi(H_t^\nu)$ is a trace class operator. Then it follows as in \cite[Lemma 3.3]{BM} that $\pi(h_t^\nu)$ is a trace class operator and \begin{equation}\label{equ-tr} \operatorname{Tr}\pi(h_t^\nu)=\operatorname{Tr}\pi(H_t^\nu). \end{equation} Now assume that $\pi$ is a unitary admissible representation. Let $A:\mathcal{H}_\pi\rightarrow \mathcal{H}_\pi$ be a bounded operator which is an intertwining operator for $\pi|_{K}$. Then $A\circ\pi(h_t^\nu)$ is again a finite rank operator. Define an operator $\tilde{A}$ on $\mathcal{H}_\pi\otimes V_\nu$ by $\tilde{A}:=A\otimes\operatorname{Id}$. Then by the same argument as in \cite[Lemma 5.1]{BM} one has \begin{equation}\label{equtrace} \operatorname{Tr} \left(\tilde{A}\circ \tilde{\pi}(H_t^\nu)\right)=\operatorname{Tr}\left(A\circ\pi(h_t^\nu)\right). \end{equation} Together with \eqref{integop} we obtain \begin{equation}\label{TrFT} \operatorname{Tr}\left(A\circ\pi(h_t^\nu)\right)=e^{t(\pi(\Omega)-\nu(\Omega_{K_\infty}))} \operatorname{Tr}\left(\tilde{A}|_{(\mathcal{H}_{\pi}\otimes V_{\nu})^{K}}\right). \end{equation} \section{Heat kernel estimates}\label{sec-heatkernel} \setcounter{equation}{0} Let the notation be as in the previous section. In this section we prove some estimations for the function $h_t^\nu$ defined by \eqref{loctrace}. Let $\widetilde K^\nu(t,x,y)$ be the kernel of $e^{-t\widetilde\Delta_\nu}$. Observe that $\widetilde K^\nu(t,x,y)\in\operatorname{Hom}((\widetilde E_\nu)_y,(\widetilde E_\nu)_x)$. Denote by $|\widetilde K^\nu(t,x,y)|$ the norm of this homomorphism. Furthermore, let $r(x,y)$ denote the geodesic distance of $x,y\in\widetilde X$. \begin{prop}\label{prop-estim} Let $d=\dim\widetilde X$. For every $T>0$ there exists $C>0$ such that we have \[ |\widetilde K^\nu(t,x,y)|\le C t^{-d/2} exp\left(-\frac{r^2(x,y)}{4t}\right) \] for all $0<t\le T$ and $x,y\in\widetilde X$. \end{prop} \begin{proof} If $\nu$ is irreducible, this is proved in \cite[Proposition 3.2]{Mu1}. However, the proof does not make any use of the irreducibility of $\nu$. So it extends without any change to the case of finite-dimensional representations. \end{proof} Let $x_0:=eK_\infty\in\widetilde X$ be the base point. For $g\in G(\R)$ and $x\in\widetilde X$ let $L_g\colon\widetilde E_x\to\widetilde E_{gx}$ be the isomorphism induced by the left translation. The kernel $\widetilde K^\nu$ is related to the convolution kernel $H_t^\nu\colon G(\R)\to\operatorname{End}(V_\nu)$ by \begin{equation}\label{heat-conv} H_t^\nu(g_1^{-1}g_2)=L_{g_1}^{-1}\circ \widetilde K^\nu(t,g_1x_0,g_2x_0)\circ L_{g_2}, \quad g_1,g_2\in G(\R). \end{equation} Thus we get \begin{equation}\label{heat-conv1} h_t^\nu(g):=\operatorname{tr} H_t^\nu(g)=\operatorname{tr}(\widetilde K^\nu(t,x_0,gx_0)\circ L_g), \quad g\in G(\R). \end{equation} Using Proposition \ref{prop-estim} and the fact that $L_g$ is an isometry, we obtain the following corollary. \begin{corollary}\label{estim3} Let $d=\dim\widetilde X$. For all $T>0$ there exists $C>0$ such that we have \[ |h_t^\nu(g)|\le C t^{-d/2} \exp\left(-\frac{r^2(gx_0,x_0)}{4t}\right) \] for all $0<t\le T$ and $g\in G(\R)$. \end{corollary} Next we turn to the asymptotic expansion of the heat kernel. For $x_0\in\widetilde X$ and $\bx\in T_{x_0}\widetilde X$ let $d_{\bx}\exp_{x_0}$ be the differential of the exponential map $\exp_{x_0}\colon T_{x_0}\widetilde X\to\widetilde X$ at the point $\bx$. It is a map from $T_{x_0}\widetilde X$ to $T_x\widetilde X$, where $x=\exp_{x_0}(\bx)$. Let \begin{equation}\label{jacob1} j_{x_0}(\bx):=|\det(d_{\bx}\exp_{x_0})| \end{equation} be the Jacobian, taken with respect to the inner products in the tangent spaces. We use $\exp_{x_0}$ to introduce normal coordinates centered at $x_0$. Let $g_{ij}(\bx)$ denote the components of the metric tensor in these coordinates. Then on has \begin{equation}\label{jacob2} j_{x_0}(\bx)=|\det(g_{ij}(\bx))|^{1/2} \end{equation} (see \cite[(1.22)]{BGV}). Given $y\in\widetilde X$ and $\bx\in T_y\widetilde X$, let $x=\exp_y(\bx)$. Put \begin{equation}\label{jacob3} j(x;y):=j_y(\bx). \end{equation} Let $\varepsilon>0$ be sufficiently small. Let $\psi\in C^\infty(\R)$ with $\psi(u)=1$ for $u<\varepsilon$ and $\psi(u)=0$ for $u>2\varepsilon$. \begin{prop}\label{prop-asympexp} Let $d=\dim\widetilde X$. Let $(\nu,V_\nu)$ be a finite-dimensional unitary representation of $K_\infty$. There exist smooth sections $\Phi_i^\nu\in C^\infty(\widetilde X\times\tilde X, \widetilde E_\nu\boxtimes \widetilde E^*_\nu)$, $i\in\N_0$, such that for every $N\in\N$ \begin{equation}\label{asympexp} \begin{split} \widetilde K^\nu(t,x,y)=(4\pi t)^{-d/2}\psi(d(x,y)) \exp\left(-\frac{r^2(x,y)}{4t}\right)\sum_{i=0}^N\Phi_i^\nu(x,y&) j(x;y)^{-1/2}t^i\\ &+O(t^{N+1-d/2}), \end{split} \end{equation} uniformly for $0< t\le 1$. Moreover the leading term $\Phi_0^\nu(x,y)$ is equal to the parallel transport $\tau(x,y)\colon (\widetilde E_\nu)_y\to (\widetilde E_\nu)_x$ with respect to the connection $\nabla^\nu$ along the unique geodesic joining $x$ and $y$. \end{prop} \begin{proof} Let $\Gamma\subset G$ be a co-compact torsion free lattice. It exists by \cite{Bo}. Let $X=\Gamma\bs \widetilde X$ and $E_\nu=\Gamma\bs \widetilde E_\nu$. As in \cite[Sect 3]{Do}, the proof can be reduced to the compact case, which follows from \cite[Theorem 2.30]{BGV}. \end{proof} Let $\gf=\kf\oplus \pf$ be the Iwasawa decomposition. We recall that the mapping \[ \varphi\colon \pf\times K_\infty\to G(\R), \] defined by $\varphi(Y,k)=\exp(Y)\cdot k$ is a diffeomorphism \cite[Ch. VI, Theorem 1.1]{He}. Thus each $g\in G(\R)$ can be uniquely written as \begin{equation}\label{cartan2} g=\exp(Y(g))\cdot k(g),\quad Y(g)\in\pf,\; k(g)\in K_\infty. \end{equation} Using \eqref{heat-conv1} and Proposition \ref{prop-asympexp}, we obtain the following corollary. \begin{corollary} There exist $a^\nu_i\in C^\infty(G(\R))$, $i\in\N_0$, such that for every $N\in\N$ we have \begin{equation}\label{asympexp1} h^\nu_t(g)=(4\pi t)^{-d/2}\psi(d(gx_0,x_0)) \exp\left(-\frac{r^2(gx_0,x_0)}{4t}\right) \sum_{i=0}^N a_i^\nu(g) t^i+O(t^{N+1-d/2}) \end{equation} which holds for $0<t\le1$. Moreover the leading coefficient $a_0^\nu$ is given by \begin{equation}\label{leadcoeff} a_0^\nu(g)=\operatorname{tr}(\nu(k(g)))\cdot j(x_0,gx_0)^{-1/2}. \end{equation} \end{corollary} \begin{proof} By \eqref{heat-conv} we have \[ H^\nu_t(g)=\widetilde K^\nu(t,x_0,gx_0)\circ L_g,\quad g\in G(\R). \] Put \begin{equation}\label{trace5} a_i^\nu(g):=\operatorname{tr}(\Phi_i^\nu(x_0,gx_0)\circ L_g)\cdot j(x_0,gx_0)^{-1/2}, \quad g\in G(\R). \end{equation} Then \eqref{asympexp1} follows immediately from \eqref{asympexp} and the definition of $h_t^\nu$. To prove the second statement, we recall that $\Phi_0^\nu(x,y)$ is the parallel transport $\tau(x,y)$ with respect to the canonical connection of $\widetilde E_\nu$ along the geodesic connecting $x$ and $y$. Let $g=\exp(Y)\cdot k$, $Y\in\pf$, $k\in K_\infty$. Then the geodesic connecting $x_0$ and $gx_0$ is the curve $\gamma(t)=\exp(tY)x_0$, $t\in [0,1]$ (see \cite[Ch. IV, Theorem 3.3]{He}). The parallel transport along $\gamma(t)$ equals $L_{\exp(Y)}$. Thus $\Phi_0^\nu(x_0,gx_0)=L_{\exp(Y)}^{-1}$. Hence we get. \[ \Phi_0^\nu(x_0,gx_0)\circ L_g=L_k=\nu(k). \] Together with \eqref{trace5} the claim follows. \end{proof} \section{Regularized traces}\label{sec-regtr} \setcounter{equation}{0} Let $G$ be a reductive algebraic group over $\Q$. For simplicity, we assume that the center $Z_G$ splits over $\Q$, i.e., we have $Z_G=A_G$. Let \[ G(\R)^1=G(\A)^1\cap G(\R). \] Then $G(\R)^1$ is semisimple and \begin{equation}\label{iso4} G(\R)=G(\R)^1\cdotA_G(\R)^0. \end{equation} Let $K_\infty\subset G(\R)^1$ be a maximal compact subgroup and let $K_\infty^0$ be the connected component of the identity. Let \[ \widetilde X=G(\R)^1/K_\infty^0 \] be the associated Riemannian symmetric space. Let $\Gamma\subset G(\Q)$ be an arithmetic subgroup. For simplicity we assume that $\Gamma$ is torsion free. Note that $\Gamma\subset G(\R)^1$. Let $X=\Gamma\bs \widetilde X$ be the associated locally symmetric manifold. Let $\nu\colon K_\infty\to \operatorname{GL}(V_\nu)$ an irreducible unitary representation of $K_\infty$ and let $E_\nu\to X$ be the associated locally homogeneous vector bundle over $X$. Let $\Delta_\nu$ be the corresponding Bochner-Laplace operator acting in $C^\infty(X,E_\nu)$. Our goal is to define a regularized trace of the heat operator $e^{-t\Delta_\nu}$. Recall that $e^{-t\Delta_\nu}$ is an integral operator with a smooth kernel $K_\nu(t,x,y)$. By definition, $K_\nu(t,x,y)\in\operatorname{Hom}((E_\sigma)_y, (E_\sigma)_x)$. Especially, $K(t,x,x)$ is an endomorphism of $(E_\sigma)_x$. Let $\operatorname{tr} K_\nu(t,x,x)$ be the trace of this endomorphism. If $X$ is compact, then we have \begin{equation}\label{regtrace0} \operatorname{Tr}\left(e^{-t\Delta_\nu}\right)=\int_X\operatorname{tr} K_\nu(t,x,x)\,dx. \end{equation} To begin with we rewrite \eqref{regtrace0}. Let $\widetilde\Delta_\nu$ be the Bochner-Laplace operator on the universal covering $\widetilde X$. Let $H_t^\nu\colon G(\R)^1\to \operatorname{End}(V_\nu)$ be the convolution kernel of $e^{-t\widetilde\Delta_\nu}$ and \begin{equation}\label{ptw-trace} h_t^\nu(g):=\operatorname{tr} H_t^\nu(g). \end{equation} Assume that $\Gamma\bs G(\R)^1$ is compact. Then one has \begin{equation}\label{trace4} \operatorname{Tr}\left(e^{-t\Delta_\nu}\right)=\int\limits_{\Gamma\bs G(\R)^1}\sum_{\gamma\in\Gamma} h_t^\nu(g^{-1}\gamma g)\,dg. \end{equation} For the proof see \cite[(3.13)]{MP2}. Let $R_\Gamma$ denote the right regular representation of $G(\R)^1$ on $L^2(\Gamma\bs G(\R)^1)$. Recall that for any $f\in C^\infty_c(G(\R)^1)$, $R_\Gamma(f)$ is the integral operator with kernel $\sum_{\gamma\in\Gamma}f(g_1^{-1}\gamma g_2)$. Thus the right hand side of \eqref{trace4} equals $\operatorname{Tr} R_\Gamma(h_t^\nu)$ and \eqref{regtrace0} can be rewritten as \begin{equation}\label{trace4a} \operatorname{Tr}\left(e^{-t\Delta_\nu}\right)=\operatorname{Tr} R_\Gamma(h_t^\nu). \end{equation} If $X$ is not compact, we choose an appropriate height function $h$ on $X$, that is, a function which measures how far out in the cusp a point is. For $Y>0$ let $X_Y=\{x\in X\colon h(x)\le Y\}$. If $X=\Gamma\backslash \tilde X$ with $\Gamma\subseteq G(\Q)$ a congruence subgroup, there is a canonical choice of height function: Recall the function $H_0=H_{P_0}: G(\R)^1\longrightarrow\mathfrak{a}_0$ and fix a norm $\|\cdot\|$ on $\mathfrak{a}_0$. Then $X\ni x\mapsto \max_{\gamma\in \Gamma}\|H_0(\gamma x)\|$ defines a height function on $X$. Assume that $X_Y$ is compact. Then the integral \begin{equation}\label{regtrace1} \int_{X_Y}\operatorname{tr} K_\nu(t,x,x)\,dx \end{equation} is well defined. Suppose that this integral has an asymptotic expansion in $Y$. Then it is natural to define the regularized trace $\operatorname{Tr}_{\reg}(e^{-t\Delta_\nu})$ as the finite part of the integral as $Y\to\infty$. For hyperbolic manifolds this has been carried out in \cite{MP1}. The regularized trace defined in this way depends of course on the choice of the height function. To choose the height function, we pass to the adelic setting. In fact, we will not define it explicitly. Instead we use directly the truncated manifold. Let $K_f\subset G(\A_f)$ be a neat open compact subgroup. Let $X(K_f)$ be the arithmetic manifold defined by \eqref{adel-quot2} and let $E_\nu\to X(K_f)$ be the locally homogeneous vector bundle defined by \eqref{vectbdl}. By \eqref{g-modules} we have \[ L^2(X(K_f),E_\nu)=\bigoplus_{i=1}^l L^2(\Gamma_i\bs\widetilde X,E_{i,\nu}). \] Using \eqref{iso4} we extend $h_t^\nu$ to a $C^\infty$-function on $G(\R)$ by \begin{equation}\label{extension1} \widetilde h_t^\nu(g_\infty z)=h_t^\nu(g_\infty),\quad g_\infty\in G(\R)^1,\; z\in A_G(\R)^0. \end{equation} Let ${\bf 1}_{K_f}$ denote the characteristic function of $K_f$ in $G(\A_f)$. We normalize the characteristic function by \begin{equation}\label{charact1} \chi_{K_f}:=\frac{{\bf 1}_{K_f}}{\operatorname{vol}(K_f)}. \end{equation} Define $\phi_t^\nu\in C^\infty(G(\A))$ by \begin{equation}\label{extension2} \phi_t^\nu(g_\infty g_f)=\widetilde h_t^\nu(g_\infty)\chi_{K_f}(g_f) \end{equation} for $ g_\infty\in G(\R)$, $g_f\in G(\A_f)$. Let $R$ denote the right regular representation of $G(\A)$ on $L^2(A_G(\R)^0 G(\Q)\bs G(\A))$, and let $\Pi_{K_f}$ denote the orthogonal projection of $L^2(A_G(\R)^0 G(\Q)\bs G(\A))$ onto $L^2(A_G(\R)^0 G(\Q)\bs G(\A))^{K_f}$. Using the isomorphism \eqref{g-modules} of $G(\R)$-modules, it follows that \begin{equation}\label{trace6} R(\phi^\nu_t)=\left[\bigoplus_{i=1}^lR_{\Gamma_i}(h_t^\nu)\right]\circ \Pi_{K_f}. \end{equation} Let $\Co(G(\R)^1)$ be Harish-Chandra's Schwartz space (see \cite[7.1.2]{Wal}). As explained in section \ref{sec-bochlapl}, we have $h_t^\nu\in\Co(G(\R)^1)$ for all $t>0$. This implies $\phi_t^\nu\in\Co(G(\A),K_f)$. Thus by \cite[Theorem 7.1]{FL1}, $J^T(\phi_t^\nu)$ is defined for all $T\in\af_0$. Let $G(\A)^1_{\le T}$ be defined by \eqref{trunc1} and for $T\in\af_0$ let $d(T)$ be defined by \eqref{dT}. Let $C\subset\af_0^+$ be a positive cone for which there exists $c>0$ such that \[ d(T)\ge c\|T\|,\quad \text{for}\;\text{all}\;T\in C. \] Then by Theorem \ref{theo-trunc} it follows that for every $t>0$ we have \begin{equation}\label{trunc3} \int\limits_{G(\Q)\bs G(\A)^1_{\le T} }\sum_{\gamma\in G(\Q)} \phi_t^\nu(x^{-1}\gamma x)\,dx=J^T(\phi_t^\nu)+O\left(e^{-\|T\|/2}\right) \end{equation} for all $T\in C$ with $\|T\|>d_0/c$, where $d_0>0$ is as in Theorem \ref{theo-trunc} and the implied constant in the remainder term depends on $t$. Now by \cite[Theorem 7.1]{FL1}, $J^T(\phi_t^\nu)$ is a polynomial in $T$. Therefore \eqref{trunc3} suggests to define the regularized trace of $e^{-t\Delta_\nu}$ as the constant term of this polynomial. In order to relate it to the trace formula, we need an additional assumption. Let $T_0\in\af_0$ be the unique point determined by \cite[Lemma 1.1]{Ar3}. Then the distribution $J_{\geo}$ is defined by \[ J_{\geo}(f):=J^{T_0}(f),\quad f\in C^\infty_c(G(\A)^1). \] and by \cite[Theorem 7.1]{FL1}, $J_{\geo}$ extends continuously to $\Co(G(\A)^1)$. As proved in \cite[p. 19]{Ar3}, $J_{\geo}$ depends only on the choice of a maximal compact subgroup $K$ of $G(\A)$ and $M_0$, but is independent of the choice of the minimal parabolic subgroup $P_0$ with Levi component $M_0$. Let $W_0$ be the Weyl group of $(G,A_0)$. For $s\in W_0$ let $w_s\in G(\Q)$ be a representative of $s$. As shown in \cite[p. 10]{Ar3}, $w_s$ belongs to $KM_0(\A)$ for all $s\in W_0$. Now assume that each $s\in W_0$ has a representative in $G(\Q)\cap K$. Then it follows from \cite[Lemma 1.1]{Ar3} that $sT_0=T_0$ for all $s\in W_0$, and therefore $T_0=0$. Thus in this case, the constant term of $J^T(f)$ equals $J_{\geo}(f)$. Let $G=\operatorname{GL}(n)$. Then $A_0$ consists of diagonal matrices and $W_0$ is equal to the symmetric group $S_n$ which acts by permutations. A permutation matrix $P_\pi$ is a $n\times n$-matrix where in row $i$ the entry $\pi(i)$ is equal to $1$ and all other entries are equal to $0$. Such a matrix belongs to $\operatorname{GL}(n,\Q)\cap K$. The case $G=\operatorname{SL}(n)$ is similar. Thus for $G=\operatorname{GL}(n)$ and $G=\operatorname{SL}(n)$ the constant term of $J^T(f)$ equals $J_{\geo}(f)$ and the above discussion suggests to define the regularized trace to be $J_{\geo}(\phi^\nu_t)$. In general $T_0\neq 0$ and therefore, $J_{\geo}(\phi^\nu_t)$ is not the constant term of the polynomial $J_{\geo}^T(\phi^\nu_t)$. Nevertheless we choose $J_{\geo}(\phi^\nu_t)$ as the definition of the regularized trace in general, because of its independence on the choice of the minimal parabolic subgroup $P_0$. \begin{definition}\label{def-regtrace} Let $G$ be a reductive algebraic group over $\Q$ with $\Q$-split center. The regularized trace of $e^{-t\Delta_\nu}$ is defined to be \[ \operatorname{Tr}_{\reg}\left(e^{-t\Delta_\nu}\right):=J_{\geo}(\phi^\nu_t). \] \end{definition} \section{The asymptotic expansion of the regularized trace for $\operatorname{GL}(n)$ and $\operatorname{SL}(n)$} \label{sec-asymp} \setcounter{equation}{0} It is well known that on a compact manifold, the trace of the heat operator $\operatorname{Tr}\left(e^{-t\Delta}\right)$ of a generalized Laplacian $\Delta$ admits an asymptotic expansion as $t\to 0$ (see \cite{BGV}). We wish to establish a similar result for the regularized trace $\operatorname{Tr}_{\reg}(e^{-\Delta_\nu})$ introduced in section \ref{sec-regtr}. For technical reasons we have to restrict to the group $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$. \subsection{Auxiliary results} We fix an open compact subgroup $K_f\subset G(\A_f)$. Let $\phi_t^\nu\in C^\infty(G(\A)^1)$ be defined by \eqref{extension2}. To begin with we replace $\phi_t^\nu$ by a compactly supported function. Let $0<a<b$. Let $h\in C^\infty(\R)$ such that $h(u)=1$ if $|u|\le a$, and $h(u)=0$, if $|u|\ge b$. Let $d(x,y)$ denote the geodesic distance of $x,y\in\widetilde X$. Put \[ r(g_\infty):=d(g_\infty K_\infty,K_\infty). \] Let $\varphi\in C^\infty_c(G(\R))$ be defined by \[ \varphi(g_\infty):=h(r(g_\infty)). \] Define $\tilde\phi_t^\nu\in C^\infty(G(\A))$ by \begin{equation}\label{extension3} \tilde\phi_t^\nu(g_\infty g_f):=\varphi(g_\infty)h_t^\nu(g_\infty)\chi_{K_f}(g_f). \end{equation} for $g_\infty\in G(\R)$ and $g_f\in G(\A_f)$. Then the restriction of $\tilde\phi_t^\nu$ to $G(\A)^1$ belongs to $C^\infty_c(G(\A)^1)$. \begin{prop}\label{prop-asympexp5} There exist $C,c>0$ such that \[ |J_{\geo}(\phi_t^\nu)-J_{\geo}(\tilde\phi_t^\nu)|\le C e^{-c/t} \] for $0<t\le 1$. \end{prop} \begin{proof} Let $J_{\spec}(\Phi)$, $\Phi\in C_c^\infty(G(\A)^1)$, be the spectral side of the trace formula. By \cite{FLM1}, $J_{\spec}(\Phi)$ converges absolutely for $\Phi\in \Co(G(\A)^1;K_f)$ and by the trace formula we have \begin{equation}\label{traceform3} J_{\geo}(\Phi)=J_{\spec}(\Phi),\quad \Phi\in \Co(G(\A)^1;K_f). \end{equation} Put $\psi_t^\nu:=\phi_t^\nu-\tilde\phi_t^\nu$ and $f:=1-\varphi$. Let $\Omega$ (resp. $\Omega_{K_\infty}$) denote the Casimir operator of $G(\R)$ (resp. $K_\infty$). Let \[ \Delta_{G}=-\Omega+2\Omega_{K_\infty}. \] By the proof of Theorem 3 of \cite{FLM1} (see \cite[Sect. 5]{FLM1}), it follows that there exists $k\in \N$ such that \[ |J_{\spec}(\phi_t^\nu)-J_{\spec}(\tilde\phi_t^\nu)|=|J_{\spec}(\psi_t^\nu)|\le C \|(\operatorname{Id}+\Delta_{G})^k(\psi_t^\nu)\|_{L^1(G(\A)^1)} \] for some $C>0$. Now note that by definition \[ \psi_t^\nu(g_\infty g_f)=f(g_\infty)h_t^\nu(g_\infty)\chi_{K_f}(g_f). \] Hence \[ \|(\operatorname{Id}+\Delta_{G})^k(\psi_t^\nu)\|_{L^1(G(\A)^1)} =\operatorname{vol}(K_f)\|(\operatorname{Id}+\Delta_{G})^k(fh_t^\nu)\|_{L^1(G(\R)^1)}. \] Let $\gf$ be the Lie algebra of $G(\R)$ and let $Y_1,\dots,Y_r$ be an orthonormal basis of $\gf$. Then $\Delta_G=-\sum_i Y_i^2$. Denote by $\nabla$ the canonical connection on $G(\R)$. Then it follows that there exists $C_1>0$ such that \[ |(\operatorname{Id}+\Delta_G)^kF(g)|\le C\sum_{l=0}^k\|\nabla^lF(g)\|, \quad g\in G(\R), \] for all $F\in C^\infty(G(\R))$. Let $m=\dim G(\R)$. and $j\in\N$. By \cite[Proposition 2.1]{Mu1}, for every $j\in\N$ there exist $C_2,c>0$ such that \begin{equation}\label{connect} \|\nabla^j h_t^\nu(g)\|\le C_2 t^{-(m+j)/2} e^{-cr^2(g)/t},\quad g\in G(\R), \end{equation} for all $0<t\le 1$. Since $f$ vanishes in neighborhood of $1\in G(\R)$, it follows that there exist $C_4,c_4,c_5>0$ such that \[ \sum_{l=0}^k\|\nabla^l(fh_t^\nu)(g)\|\le C_4 e^{-c_4/t}e^{-c_5r^2(g)} \] for all $g\in G(\R)$ and $0<t\le 1$. Since $G(\R)\ni g\mapsto e^{-c_5r^2(g)}$ is integrable on $G(\R)$, we obtain \[ \|(\operatorname{Id}+\Delta_{G})^k(fh_t^\nu)\|_{L^1(G(\R)^1)}\le C_5 e^{-c_4/t} \] for some constant $C_5>0$, which completes the proof. \end{proof} Let $K(N)\subset G(\A_f)$ be the principal congruence subgroup of level $N\in\N$. From now on we assume that $K_f$ is contained in $K(N)$ for some $N\ge 3$. By Proposition \ref{prop-asympexp5} it suffices to show that $J_{\geo}(\widetilde\phi_t^\nu)$ admits an asymptotic expansion as $t\to 0$. Now by our assumption on $K_f$ it follows that if the support of $f$ is a sufficiently small neighborhood of $0$, then by \eqref{unip-contr1} we have \[ J_{\geo}(\widetilde\phi_t^\nu)=J_{\operatorname{unip}}(\widetilde\phi_t^\nu). \] Now we use the fine geometric expansion \eqref{finegeoexp} by which we express $J_{\operatorname{unip}}(\widetilde\phi_t^\nu)$ in terms of a finite linear combination of weighted orbital integrals. Next we use \eqref{orbint11}. Since at the finite places our test function is fixed, we are reduced to the consideration of real weighted orbital integrals in some Levi subgroup of $G$. Since the weighted orbital integrals $J_M^L$ for $L\subseteq G$ a semistandard Levi subgroup can be treated analogously to the case of $L=G$ (they can be reduced to weighted orbital integrals of the form $J_{M'}^{\operatorname{GL}(m)}$ or $J_{M'}^{\operatorname{SL}(m)}$ for suitable $m$ and $M'$), it suffices to consider the weighted orbital integrals for $L=G$. Now observe that the kernel $H_t^\nu\colon G(\R)^1\to \operatorname{End}(V_\sigma)$ of $e^{-t\widetilde \Delta_\nu}$ satisfies \[ H_t^\nu(k^{-1}gk^\prime)=\nu(k)^{-1}\circ H_t^\nu(g)\circ \nu(k^\prime),\quad \forall k,k^\prime\in K,\forall\; g\in G(\R)^1. \] (see \cite[\S 3]{MP4}). Therefore the function $h_t^\nu=\operatorname{tr} H_t^\nu$ is invariant under conjugation by $k_\infty\in K_\infty$. Define $F_t^\nu$ by \[ F_t^\nu(g)=\varphi(g)h_t^\nu(g),\quad g\in G(\R). \] Then by \eqref{realorbint} the orbital integral that we need to consider is given by \begin{equation}\label{orbint1} J_M^G(U, F_t^\nu)=c\int_{N(\R)}F_t^\nu(n)w_{M}(n)\; dn, \end{equation} where $M\in\cL$, $U\in\left(\cU_M(\R)\right)_{M(\R)}$, $w_{M}(n):=w_{M,U}(n)$ is the weight function described in section \ref{sec-wfct} and $N$ is the unipotent radical of some parabolic subgroup $Q\in\cF$. Our goal is to determine the asymptotic behavior of the integral on the right hand side as $t\to 0^+$. To study this integral, we identify $N(\R)$ with its Lie algebra $\nf$ via the map $n\in N(\R)\mapsto n-\operatorname{Id}$. Furthermore $\nf\cong\R^k$ for some $k\in\N$. Let \[ x\in\R^k\mapsto n(x)\in N(\R) \] be the inverse map. With respect to the isomorphism the invariant measure $dn$ is identified with Lebesgue measure $dx$ in $\R^k$. Thus \eqref{orbint1} equals \begin{equation}\label{orbint2} \int_{\R^k} F_t^\nu(n(x))w_{M}(n(x))\; dx. \end{equation} To determine the asymptotic behavior of this integral as $t\to 0^+$, we will use the asymptotic expansion \eqref{asympexp1} of $h_t^\nu$. To this end we need to estimate the function \begin{equation} r(x):=r(n(x)x_0,x_0),\quad x\in\R^k, \end{equation} where $x_0=eK_\infty\in\widetilde X$. Note that $\widetilde X$ is a Hadamard manifold of nonpositive curvature and the orbit $N(\R)x_0$ is a horosphere in $\widetilde X$. Then it follows by \cite[Theorem 4.6]{HH} that there exist constants $C,c>0$ such that \begin{equation}\label{horobd} r(x)\ge C\operatorname{arcsinh}(c\|x\|),\quad x\in\R^k. \end{equation} Now note that \[ \operatorname{arcsinh}(x)=\ln\left(x+\sqrt{x^2+1}\right). \] Thus we get \begin{equation}\label{bound2} r(x)\ge C\ln\left(1+\|x\|\right),\quad x\in\R^k. \end{equation} We also need the Taylor expansion of $r(x)^2$ at $x=0$. This is described by the following lemma. \begin{lem}\label{lem-taylor} We have \[ r(x)^2=\frac{1}{4}\|x\|^2+O(\|x\|^3) \] as $x\to 0$. \end{lem} \begin{proof} Let $H= (H_1,\ldots, H_n)\in\mathfrak{a}$, $H_1+\ldots, H_n=0$, with $n(x)\in K_\infty e^H K_\infty$. Now note that $r(e^Hx_0,x_0)=\|H\|$ (see \cite[Corollary 10.42]{BH}). Thus it follows that $r(x)^2= \|H\|^2$. Moreover, \[ n + \|x\|^2=\operatorname{tr} \left(n(x)^t n(x)\right) = \operatorname{tr} e^{2H} = n + 4 \|H\|^2 + O(\|H\|^3). \] If $x\rightarrow 0$, then also $H\rightarrow 0$ so that this equation implies $\|H\|^3 = O(\|x\|^3)$ for small $x$. Hence $r(x)^2=\|H\|^2 = \frac{1}{4}\|x\|^2 + O(\|x\|^3)$ as $x\rightarrow0$. \end{proof} \subsection{Asymptotics for $t\to0$} We can now turn to the estimation of the weighted orbital integral \eqref{orbint2}. For $\varepsilon>0$ let $B(\varepsilon)\subset \R^k$ denote the ball of radius $\varepsilon$ centered at the origin and let $U(\varepsilon)=\R^k\setminus B(\varepsilon)$. Let $\psi$ be the function occurring in \eqref{asympexp1}. Choose $\varepsilon>0$ so small such that $\varphi(n(x))=1$ for $x\in B(\varepsilon)$ and $\operatorname{supp} \psi(n(\cdot))\subset B(\varepsilon)$. Let $0<t\le 1$. By Corollary~\ref{estim3} we have \[ \bigg|\int_{U(\varepsilon)}\varphi(n(x))h_t^\nu(n(x))w_M(n(x))\;dx\bigg|\le C t^{-d/2}\int_{U(\varepsilon)}\exp\left(-\frac{r^2(x)}{4t}\right)\left|w_M(n(x))\right|\;dx \] for some absolute constant $C>0$. Using the lower bound~\ref{bound2} for $r(x)$, and the result on the weight function from Proposition~\ref{prop} we can find $c, C_1, C_2>0$ such that \begin{multline*} t^{-d/2}\int_{U(\varepsilon)}\exp\left(-\frac{r^2(x)}{4t}\right)\left|w_M(n(x))\right|\;dx \\ \le C_1 \exp\left(-\frac{c(\varepsilon)}{t}\right) \int_{\R^k} \exp\left(- c\left(\log(1+ \|x\|)\right)^2\right)\lambda(x)\;dx \end{multline*} where $c(\varepsilon)=C_2\log(1+\varepsilon)$ and $\lambda:\R^k\longrightarrow\C$ is a function of the form~\eqref{eq:log:homog:polynom}. By~\eqref{eq:log:int2} of Lemma~\ref{logconv} the last integral is bounded by a constant so that we finally obtain \begin{equation}\label{bound3} \bigg|\int_{U(\varepsilon)}\varphi(n(x))h_t^\nu(n(x))w_M(n(x))\;dx\bigg| \le C_3\exp\left(-\frac{c(\varepsilon)}{t}\right) \end{equation} for some absolute constant $C_3>0$. To deal with the integral over $B(\varepsilon)$, we use \eqref{asympexp1}, which gives \begin{equation}\label{asympexp3} \begin{split} \int_{B(\varepsilon)}h_t^\nu(n(x))w_M(n(x))\;dx= &t^{-d/2}\sum_{i=0}^N t^i \int_{B(\varepsilon)}\exp\left(-\frac{r^2(x)}{4t}\right) a_i^\nu(x)w_M(n(x))\;dx\\ &+O(t^{N+1-d/2}) \end{split} \end{equation} for $0<t\le 1$, where $a_i^\nu\in C_c^\infty(B(\varepsilon))$. Each of the integrals is of the form \begin{equation}\label{integr1} \int_{B(\varepsilon)}\exp\left(-\frac{r^2(x)}{4t}\right)f(x)w_M(\operatorname{Id}+x)\;dx, \end{equation} where $f\in C_c^\infty(B(\varepsilon))$. We expand $r^2(x)$ and $f(x)$ in their Taylor series at $0$. Let $N\in \N$, $N\ge 3$. By Lemma \ref{lem-taylor} we have \begin{equation}\label{taylor5} r^2(x)=a\|x\|^2+\psi_N(x),\quad \psi_N(x) =\sum_{3\le|\alpha|\le N}a_\alpha x^\alpha+R_N(x), \end{equation} where \begin{equation}\label{remainder} R_N(x)=\sum_{|\alpha|=N+1}\frac{D^\alpha r^2(\theta x)}{\alpha !}x^\alpha \end{equation} for $x\in B(\varepsilon)$ and some $0\le\theta\le 1$. Now we change variables by $x\mapsto \sqrt{t}x$. Then \eqref{integr1} equals \begin{equation}\label{integr2} t^{k/2}\int_{B(\varepsilon t^{-1/2})}\exp(-a\|x\|^2) \exp\left(-\frac{\psi_N(t^{1/2}x)}{t}\right)f(t^{1/2}x)w_M(1,\operatorname{Id}+t^{1/2}x) \end{equation} Now observe that by \eqref{taylor5} we have \[ t^{-1}\psi_N(t^{1/2}x)=\sum_{3\le|\alpha|\le N}a_\alpha t^{|\alpha|/2-1} x^\alpha+ t^{-1}R_N(t^{1/2}x) \] and \[ t^{-1}R_N(t^{1/2}x)=t^{(N-1)/2}\sum_{|\alpha|=N+1} \frac{D^\alpha r^2(t^{1/2}\theta x)}{\alpha !}x^\alpha. \] There is $C>0$ such that \[ |D^\alpha r^2(t^{1/2}\theta x)|\le C \] for all $x\in B(t^{-1/2}\varepsilon)$, $0<t\le 1$, and $0\le\theta\le1$. Hence it follows that there is $C_1>0$ such that for all $0<t\le1$ \[ |t^{-1}R_N(t^{1/2}x)|\le C_1 t^{(N-1)/2}\|x\|^{N+1}, \quad x\in B(t^{-1/2}\varepsilon). \] Using the Taylor expansion of $\exp(u)$, we get for $n\ge 3$ \begin{equation}\label{taylor6} \exp\left(-\frac{\psi_N(t^{1/2}x)}{t}\right)=\sum_{j=0}^N t^{j/2}p_j(x)+R_N(t,x), \end{equation} where $p_j(x)$ is a polynomial of degree $\le N^2$ and the remainder term satisfies \begin{equation}\label{remaind3} |R_N(t,x)|\le C_2 t^{(N-1)/2} (1+\|x\|)^{N^2} \end{equation} for some constant $C_2>0$, $0<t\le 1$ and $x\in B(t^{-1/2}\varepsilon)$. Similarly, using the Taylor expansion of $f(x)$ we get \begin{equation}\label{taylor7} f(t^{1/2}x)=\sum_{|\alpha|\le N}b_\alpha t^{|\alpha|/2}x^\alpha+Q_N(t,x) \end{equation} with \[ |Q_N(t,x)|\le C_3 t^{(N+1)/2}(1+\|x\|)^{N+1} \] for $0<t\le 1$ and $x\in B(t^{-1/2}\varepsilon)$. Using \eqref{integr2}, \eqref{taylor6}, \eqref{taylor7}, and Proposition \ref{prop}, it follows that \begin{equation}\label{int-exp} \begin{split} \int_{B(\varepsilon)}\exp\left(-\frac{r^2(x)}{4t}\right)f(x)w_M(\operatorname{Id}+x)\;dx= t^{k/2}\sum_{j=0}^N\sum_{i=0}^{r_j}a_{ij}(t)(\log t)^i t^{j/2}+\phi_N(t), \end{split} \end{equation} where each $a_{ij}(t)$ is of the form \[ \int_{B(t^{-1/2}\varepsilon)} e^{-a\|x\|^2}p(x)\prod_{l=1}^h\big|\log|p_l(x)|\big|\;dx, \] if $i<r_j$, or \[ \int_{B(t^{-1/2}\varepsilon)} e^{-a\|x\|^2}p(x)\;dx, \] if $i=r_j$, with homogeneous polynomials $p(x)$, $p_1(x),...,p_h(x)$. The fact that for $i=r_j$ no logarithm appears in the integral for $a_{ij}(t)$ can easily be seen by changing variables from $x$ to $t^{1/2} x$ in the integral on the left hand side of \eqref{int-exp} and collecting all $\log t$ terms coming from the weight function. Finally, $\phi_N(t)$ satisfies \[ |\phi_N(t)|\le C t^{(N+k+1)/2}\int_{B(t^{-1/2}\varepsilon)}e^{-a\|x\|^2}(1+\|x\|)^{N^2} \prod_{i=1}^m\big\|\log|p_i(x)|\big\|\;dx. \] Let $U(r)=\R^k\setminus B(r)$. Since $\log(1+\|x\|)\le\|x\|$ for all $x$, it follows from Lemma \ref{logconv} that \[ \Big|\int_{U(t^{-1/2}\varepsilon)} e^{-a\|x\|^2}p(x) \prod_{l=1}^h\big|\log|p_l(x)|\big|\;dx\Big|\le C e^{-a\varepsilon^2/(2t)}, \] for $0<t\le 1$. Thus there are constants $c_{ij}\in\R$ and $c>0$ such that \[ a_{ij}(t)=c_{ij}+O(e^{-c/t}) \] for $0<t\le 1$. By the considerations above, for each pair $(i,j)$, $0\le i\le r_j$, there exist homogeneous polynomials $p,p_1,\dots,p_h$, such that \begin{equation}\label{coeffi} c_{ij}=\begin{cases} \int_{\R^k} e^{-a\|x\|^2}p(x)\prod_{l=1}^h\big|\log|p_l(x)|\big|\;dx,& \text{if}\; i<r_j,\\ {}&{}\\ \int_{\R^k} e^{-a\|x\|^2}p(x)\;dx,&\text{if}\; i=r_j. \end{cases} \end{equation} In the same way we get \[ |\phi_N(t)|\le C t^{(N+k+1)/2},\quad 0<t\le1. \] Putting everything together, we get \begin{prop}\label{prop-asympexp1} Let $M\in\cL$, $M\neq G$. For every $N\in\N$, $N\ge 3$, there is an expansion \begin{equation}\label{unipo-asympexp} J_M(u,\widetilde\phi_t^\nu)=t^{-(d-k)/2}\sum_{j=0}^N\sum_{i=0}^{r_j}c_{ij}(\nu)t^{j/2} (\log t)^i+O(t^{(N-d+k+1)/2}) \end{equation} as $t\to 0^+$. \end{prop} Now we come to the first term in \eqref{finegeoexp}, where $f=\widetilde\phi_t^\nu$. Then we have to determine the asymptotic behavior of $h_t^\nu(1)$ as $t\to +0$. Let $\Gamma^\prime\subset G(\R)$ be a cocompact torsion free lattice. Such a lattice exists by \cite{Bo1}. Let $X^\prime=\Gamma^\prime\bs \widetilde X$ and $E_\nu^\prime\to X^\prime$ the locally homogeneous vector bundle associated to $\nu$. Let $\Delta_{X^\prime,\nu}$ be the corresponding Bochner-Laplace operator. The kernel of $e^{-t\Delta_{X^\prime,\nu}}$, regarded as operator in $L^2(\Gamma\bs G(\R),\nu)$, is given by \[ K^\nu(t,g_1,g_2):=\sum_{\gamma\in\Gamma^\prime} H_t^{\nu}(g_1^{-1}\gamma g_2). \] Hence \[ \begin{split} \operatorname{Tr}\left(e^{-t\Delta_{X^\prime,\nu}}\right)&=\int_{\Gamma\bs G(\R)}\operatorname{tr} K^\nu(t,g,g)\;dg =\int_{\Gamma\bs G(\R)}\sum_{\gamma\in\Gamma^\prime} h_t^\nu(g^{-1}\gamma g)\; dg\\ &=\operatorname{vol}(\Gamma^\prime\bs G(\R))h_t^\nu(1) + \int_{\Gamma^\prime\bs G(\R)} \sum_{\gamma\in\Gamma^\prime\setminus\{1\}} h_t^\nu(g^{-1}\gamma g)\;dg. \end{split} \] As in \cite[(5.10)]{MP2}, the last term on the right can be estimated by $C_1e^{-c_1/t}$ for $0<t\le 1$ and some constants $C_1,c_1>0$. Thus we get \[ h_t^{\nu}(1)=\frac{1}{\operatorname{vol}(\Gamma^\prime\bs G(\R)}\operatorname{Tr}\left(e^{-t\Delta_{X^\prime,\nu}} \right)+O(e^{-c_1/t}) \] for $0<t\le 1$. Now the trace of the heat operator on a compact manifold has an asymptotic expansion as $t\to +0$ (see \cite{Gi}). Hence, it follows that there is an asymptotic expansion \[ h_t^\nu(1)\sim \sum_{j=0}^\infty a_j t^{-d/2+j} \] as $t\to +0$. Combined with Proposition \ref{prop-asympexp1} we obtain Theorem \ref{prop-asymp3}. \section{The analytic torsion}\label{sec-analtor} \setcounter{equation}{0} In this section we assume that $G=\operatorname{GL}(n)$ or $G=\operatorname{SL}(n)$. We consider the case $G=\operatorname{SL}(n)$. The case $G=\operatorname{GL}(n)$ is similar. We choose $K_\infty=\operatorname{SO}(n)$ as maximal compact subgroup of $G(\R)=\operatorname{SL}(n,\R)$. Then $\widetilde X=\operatorname{SL}(n,\R)/\operatorname{SO}(n)$ and $X(K_f)=\Gamma\bs \widetilde X$, where $\Gamma=(G(\R)\times K_f)\cap G(\Q)$. \subsection{The Hodge-Laplace operator and heat kernels} Let $\tau$ be an irreducible finite-dimensional representation of $G(\R)$ on $V_{\tau}$. Let $E_{\tau}$ be the flat vector bundle over $X$ associated to the restriction of $\tau$ to $\Gamma$. Let $\widetilde E^\tau$ be the homogeneous vector bundle associated to $\tau|_{K_\infty}$ and let $E^\tau:=\Gamma\bs \widetilde E^\tau$. There is a canonical isomorphism \begin{equation}\label{iso-vb} E^\tau\cong E_\tau \end{equation} \cite[Proposition 3.1]{MM}. By \cite[Lemma 3.1]{MM}, there exists a positive definite inner product $\left<\cdot,\cdot\right>$ on $V_{\tau}$ such that \begin{enumerate} \item $\left<\tau(Y)u,v\right>=-\left<u,\tau(Y)v\right>$ for all $Y\in\mathfrak{k}$, $u,v\in V_{\tau}$ \item $\left<\tau(Y)u,v\right>=\left<u,\tau(Y)v\right>$ for all $Y\in\mathfrak{p}$, $u,v\in V_{\tau}$. \end{enumerate} Such an inner product is called admissible. It is unique up to scaling. Fix an admissible inner product. Since $\tau|_{K_\infty}$ is unitary with respect to this inner product, it induces a metric on $E^{\tau}$, and by \eqref{iso-vb} on $E_\tau$, which we also call admissible. Let $\Lambda^{p}(E_{\tau})=\Lambda^pT^*(X)\otimes E_\tau$. Let \begin{align}\label{repr4} \nu_{p}(\tau):=\Lambda^{p}\operatorname{Ad}^{*}\otimes\tau:\:K_\infty\rightarrow\operatorname{GL} (\Lambda^{p}\mathfrak{p}^{*}\otimes V_{\tau}). \end{align} Then by \eqref{iso-vb} there is a canonical isomorphism \begin{align}\label{pforms} \Lambda^{p}(E_{\tau})\cong\Gamma\backslash(G(\R)\times_{\nu_{p}(\tau)} (\Lambda^{p}\mathfrak{p}^{*}\otimes V_{\tau})) \end{align} of locally homogeneous vector bundles. Let $\Lambda^{p}(X,E_{\tau})$ be the space the smooth $E_{\tau}$-valued $p$-forms on $X$. The isomorphism \eqref{pforms} induces an isomorphism \begin{align}\label{isoschnitte} \Lambda^{p}(X,E_{\tau})\cong C^{\infty}(\Gamma\backslash G(\R),\nu_{p}(\tau)), \end{align} where the latter space is defined as in \eqref{globsect1}. A corresponding isomorphism also holds for the spaces of $L^{2}$-sections. Let $\Delta_{p}(\tau)$ be the Hodge-Laplacian on $\Lambda^{p}(X,E_{\tau})$ with respect to the admissible metric in $E_\tau$. Let $R_\Gamma$ denote the right regular representation of $G(\R)$ in $L^2(\Gamma\bs G(\R))$. By \cite[(6.9)]{MM} it follows that with respect to the isomorphism \eqref{isoschnitte} one has \begin{equation}\label{laplace1} \Delta_{p}(\tau)=-R_\Gamma(\Omega)+\tau(\Omega)\operatorname{Id}. \end{equation} Let $\widetilde E_\tau\to \widetilde X$ be the lift of $E_\tau$ to $\widetilde X$. There is a canonical isomorphism \begin{equation}\label{iso-vbcov} \Lambda^p(\widetilde X,\widetilde E_\tau)\cong C^\infty(G(\R),\nu_p(\tau)). \end{equation} Let $\widetilde\Delta_p(\tau)$ be the lift of $\Delta_p(\tau)$ to $\widetilde X$. Then again it follows from \cite[(6.9)]{MM} that with respect to the isomorphism \eqref{iso-vbcov} we have \begin{equation}\label{kuga} \widetilde \Delta_p(\tau)=-R(\Omega)+\tau(\Omega)\operatorname{Id}. \end{equation} Let $e^{-t\widetilde\Delta_p(\tau)}$ be the corresponding heat semigroup. Regarded as an operator in $L^2(G(\R),\nu_p(\tau))$, it is a convolution operator with kernel \begin{align}\label{DefHH} H^{\tau,p}_t\colon G(\R)\to\operatorname{End}(\Lambda^p\mathfrak p^*\otimes V_\tau) \end{align} which belongs to $C^\infty\cap L^2$ and satisfies the covariance property \begin{equation}\label{covar} H^{\tau,p}_t(k^{-1}gk')=\nu_p(\tau)(k)^{-1} H^{\tau,p}_t(g)\nu_p(\tau)(k') \end{equation} with respect to the representation \eqref{repr4}. Moreover, for all $q>0$ we have \begin{equation}\label{schwartz1} H^{\tau,p}_t \in (\mathcal{C}^q(G(\R))\otimes \operatorname{End}(\Lambda^p\pf^*\otimes V_\tau))^{K_\infty\times K_\infty}, \end{equation} where $\mathcal{C}^q(G(\R))$ denotes Harish-Chandra's $L^q$-Schwartz space (see \cite[Sect. 4]{MP2}). We note that the kernel $H^{\tau,p}_t$ can be expressed in terms of the kernel $H_t^{\nu_p(\tau)}$ of the heat semigroup $e^{-t\widetilde \Delta_{\nu_p(\tau)}}$ associated to the Bochner-Laplace operator $\widetilde\Delta_{\nu_p(\tau)}$ acting in $C^\infty(\widetilde X,\widetilde E_{\nu_p(\tau)})$. For $p=0,\dots,n$ put \[ E_p(\tau):=\nu_p(\tau)(\Omega_{K_\infty}), \] which we regard as an endomorphism of $\Lambda^p\pg^*\otimes V_\tau$. It defines an endomorphism of $\Lambda^pT^\ast(X)\otimes E_\tau$. By \eqref{BLO} and \eqref{kuga} we have \[ \widetilde\Delta_{p}(\tau)=\widetilde\Delta_{\nu_p(\tau)}+ \tau(\Omega)\operatorname{Id}-E_p(\tau). \] Let $\nu_p(\tau)=\oplus_{\sigma\in\Pi(K_\infty)} m(\sigma)\sigma$ be the decomposition of $\nu_p(\tau)$ into irreducible representations. This induces a corresponding decomposition of the homogeneous vector bundle \begin{equation}\label{vb-decomp} \widetilde E_{\nu_p(\tau)}=\bigoplus_{\sigma\in\Pi(K_\infty)}m(\sigma)\widetilde E_\sigma. \end{equation} With respect to this decomposition we have \[ E_p(\tau)=\bigoplus_{\sigma\in\Pi(K_\infty)}m(\sigma) \sigma(\Omega_{K_\infty})\operatorname{Id}_{V_\sigma}, \] where $\sigma(\Omega_{K_\infty})$ is the Casimir eigenvalue of $\sigma$ and $V_\sigma$ the corresponding representation space. Let $\widetilde\Delta_\sigma$ be the Bochner-Laplace operator associated to $\sigma$. By \eqref{vb-decomp} we get a corresponding decomposition of $C^\infty(\widetilde X,\widetilde E_{\nu_p(\tau)})$ and with respect to this decomposition we have \[ \widetilde\Delta_{\nu_p(\tau)}=\bigoplus_{\sigma\in\Pi(K_\infty)} m(\sigma) \widetilde\Delta_\sigma. \] This shows that $\widetilde\Delta_{\nu_p(\tau)}$ commutes with $E_p(\tau)$. Hence we get \begin{equation}\label{equ-kernel} H^{\tau,p}_t=e^{-t(\tau(\Omega)-E_p(\tau))}\circ H^{\nu_p(\tau)}_t. \end{equation} Let $h^{\tau,p}_t\in C^\infty(G(\R))$ be defined by \begin{equation}\label{tr-kern} h^{\tau,p}_t(g)=\operatorname{tr} H^{\tau,p}_t(g),\quad g\in G(\R). \end{equation} Then by \eqref{equ-kernel} we get \begin{equation}\label{equ-kernel2} h^{\tau,p}_t=e^{t(\tau(\Omega)-\operatorname{tr} E_p(\tau))}h^{\nu_p(\tau)}_t, \end{equation} where $h^{\nu_p(\tau)}_t=\operatorname{tr} H^{\nu_p(\tau)}_t$. As in \eqref{extension2} we define $\phi^{\tau,p}_t\in C^\infty(G(\A))$ by \begin{equation}\label{extension5} \phi^{\tau,p}_t(g_\infty g_f):=h^{\tau,p}_t(g_\infty)\chi_{K_f}(g_f) \end{equation} for $g_\infty\in G(\R)$ and $g_f\in G(\A_f)$. Following Definition \ref{def-regtrace}, we define the regularized trace of $e^{-t\Delta_p(\tau)}$ by \begin{equation}\label{regtr-heat} \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right):=J_{\geo}(\phi^{\tau,p}_t). \end{equation} \subsection{Decay for the continuous spectrum} The next goal is to determine the asymptotic behavior of $\operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)$ as $t\to\infty$ and $t\to 0^+$. To study the asymptotic behavior as $t\to\infty$ we use the trace formula \eqref{tracef1}. By Theorem \ref{thm-specexpand}, $J_{\spec}$ is a distribution on $\Co(G(\A);K_f)$ and by \cite[Theorem 7.1]{FL1}, $J_{\geo}$ is continuous on $\Co(G(\A);K_f)$. This implies that \eqref{tracef1} holds for $\phi_t^{\tau,p}$ and we have \begin{equation}\label{regtrace1a} \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)=J_{\spec}(\phi_t^{\tau,p}). \end{equation} Now we apply Theorem \ref{thm-specexpand} to study the asymptotic behavior as $t\to\infty$ of the right hand side. Let $M\in\cL$ and $P\in\cP(M)$. Recall that $L^2_{\di}(A_M(\R)^0 M(\Q)\bs M(\A))$ splits as the completed direct sum of its $\pi$-isotypic components for $\pi\in\Pi_{\di}(M(\A))$. We have a corresponding decomposition of $\bar{\cA}^2(P)$ as a direct sum of Hilbert spaces $\hat\oplus_{\pi\in\Pi_{\di}(M(\A))}\bar{\cA}^2_\pi(P)$. Similarly, we have the algebraic direct sum decomposition \[ \cA^2(P)=\bigoplus_{\pi\in\Pi_{\di}(M(\A))}\cA^2_\pi(P), \] where $\cA^2_\pi(P)$ is the ${\bf K}$-finite part of $\bar{\cA}^2_\pi(P)$. For $\sigma\in\widehat{K_\infty}$ let $\cA^2_\pi(P)^\sigma$ be the $\sigma$-isotypic subspace. Then $\cA^2_\pi(P)$ decomposes as \[ \cA^2_\pi(P)=\bigoplus_{\sigma\in\widehat{K_\infty}}\cA^2_\pi(P)^\sigma. \] Let $\cA^2_\pi(P)^{K_f}$ be the subspace of $K_f$-invariant functions in $\cA^2_\pi(P)$, and for any $\sigma\in\widehat{K_\infty}$ let $\cA^2_\pi(P)^{K_f,\sigma}$ be the $\sigma$-isotypic subspace of $\cA^2_\pi(P)^{K_f}$. Recall that $\cA^2_\pi(P)^{K_f,\sigma}$ is finite dimensional. Let $M_{Q|P}(\pi,\lambda)$ denote the restriction of $M_{Q|P}(\lambda)$ to $\cA^2_\pi(P)$. Recall that the operator $\Delta_{\mathcal{X}}(P,\lambda)$, which appears in the formula \eqref{specside2}, is defined by \eqref{intertw3}. Its definition involves the intertwining operators $M_{Q|P}(\lambda)$. If we replace $M_{Q|P}(\lambda)$ by its restriction $M_{Q|P}(\pi,\lambda)$ to $\cA^2_\pi(P)$, we obtain the restriction $\Delta_{\mathcal{X}}(P,\pi,\lambda)$ of $\Delta_{\mathcal{X}}(P,\lambda)$ to $\cA^2_\pi(P)$. Similarly, let $\rho_\pi(P,\lambda)$ be the induced representation in $\bar{\cA}^2_\pi(P)$. Fix $\beta\in\mathfrak{B}_{P,L_s}$ and $s\in W(M)$. Then for the integral on the right of \eqref{specside2} with $h=\phi_t^{\tau,p}$ we get \begin{equation}\label{specside3} \sum_{\pi\in\Pi_{\di}(M(\A))}\int_{i(\af^G_{L_s})^*}\operatorname{Tr}\left( \Delta_{\mathcal{X}_{L_s}(\underline{\beta})}(P,\pi,\lambda)M(P,\pi,s)\rho_\pi(P,\lambda,\phi^{\tau,p}_t) \right)\;d\lambda. \end{equation} Let $P,Q\in\cP(M)$ and $\nu\in\Pi(K_\infty)$. Denote by $\widetilde M_{Q|P}(\pi,\nu,\lambda)$ the restriction of \[ M_{Q|P}(\pi,\lambda)\otimes\operatorname{Id}\colon \cA^2_\pi(P)\otimes V_{\nu}\to \cA^2_\pi(P)\otimes V_{\nu} \] to $(\cA^2_\pi(P)^{K_f}\otimes V_{\nu})^{K_\infty}$. Denote by $\widetilde\Delta_{\mathcal{X}_{L_s}(\underline{\beta})}(P,\pi,\nu,\lambda)$ and $\widetilde M(P,\pi,\nu,s)$ the corresponding restrictions. Let $m(\pi)$ denote the multiplicity with which $\pi$ occurs in the regular representation of $M(\A)$ in $L^2_{\di}(A_M(\R)^0 M(\Q)\bs M(\A))$. Then \begin{equation}\label{iso-ind} \rho_\pi(P,\lambda)\cong \oplus_{i=1}^{m(\pi)}\operatorname{Ind}_{P(\A)}^{G(\A)}(\pi,\lambda). \end{equation} Fix positive restricted roots of $\af_P$ and let $\rho_{\af_P}$ denote the corresponding half-sum of these roots. For $\xi\in \Pi(M(\R))$ and $\lambda\in\af^\ast_P$ let \[ \pi_{\xi,\lambda}:=\operatorname{Ind}_{P(\R)}^{G(\R)}(\xi\otimes e^{i\lambda}) \] be the unitary induced representation. Let $\xi(\Omega_M)$ be the Casimir eigenvalue of $\xi$. Define a constant $c(\xi)$ by \begin{equation}\label{casimir4} c(\xi):=-\langle\rho_{\af_P},\rho_{\af_P}\rangle+\xi(\Omega_M). \end{equation} Then for $\lambda\in\af^\ast_P$ one has \begin{equation}\label{casimir5} \pi_{\xi,\lambda}(\Omega)=-\|\lambda\|^2+c(\xi) \end{equation} (see \cite[Theorem 8.22]{Kn}). Let \begin{equation}\label{def-F} \cT:=\{\nu\in\Pi(K_\infty)\colon [\nu_p(\tau)\colon\nu]\neq 0\}. \end{equation} Using \eqref{equ-kernel}, \eqref{iso-ind} and \eqref{TrFT}, it follows that \eqref{specside3} is equal to \begin{equation}\label{specside4} \begin{split} \sum_{\pi\in\Pi_{\di}(M(\A))}\sum_{\nu\in\cT}e^{-t(\tau(\Omega)-c(\pi_\infty))} \int_{i(\af^G_{L_s})^*}e^{-t\|\lambda\|^2}\operatorname{Tr}\left( \widetilde\Delta_{\mathcal{X}_{L_s}(\underline{\beta})}(P,\pi,\nu,\lambda) \widetilde M(P,\pi,\nu,s)\right)\;d\lambda. \end{split} \end{equation} In order to estimate \eqref{specside4} from above, we need the following two preparatory results. \begin{lem}\label{est-casim} Let $(\tau,V_\tau)\in\operatorname{Rep}(G(\R))$. Assume that $\tau\not\cong\tau_\theta$. Let $P=MAN$ be a proper parabolic subgroup of $G$ and let $K_\infty^M=M(\R)\cap K_\infty$. Let $\xi\in\widehat{M(\R)}$ and assume that $\dim(W_\xi\otimes\Lambda^p\pg^\ast\otimes V_\tau)^{K^M_\infty}\neq 0$. Then one has \[ \tau(\Omega)-c(\xi)>0. \] \end{lem} \begin{proof} Let $\xi\in\widehat{M(\R)}$ with $\dim(W_{\xi}\otimes\Lambda^p\pg^\ast \otimes V_\tau)^{K^M_\infty}\neq 0$. Assume that $\tau(\Omega)-c(\xi)\le 0$. Then by \eqref{casimir5} there exists $\lambda_0\in\af^\ast$ such that \[ \pi_{\xi,\lambda_0}(\Omega)=\tau(\Omega). \] By Frobenius reciprocity we have \[ \dim\left(W_\xi\otimes\Lambda^p\pg^\ast\otimes V_\tau\right)^{K^M_\infty}= \dim\left(\H_{\xi,\lambda_0}\otimes\Lambda^p\pg^\ast\otimes V_\tau\right)^{K_{\infty}}. \] Combined with our assumption and \cite[Proposition II,3.1]{BW} it follows that \[ \dim H^p(\gf,K_\infty;\H_{\xi,\lambda_0,K_{\infty}}\otimes V_\tau)\neq 0, \] where $\H_{\xi,\lambda_0,K_{\infty}}$ denotes the subspace of $K_\infty$-finite vectors of $\H_{\xi,\lambda_0}$. Since $\tau\neq\tau_\theta$, this is a contradiction to the first statement of \cite[Proposition II. 6.12]{BW}. Thus it follows that \[ \tau(\Omega)-c(\xi)>0 \] for all $\xi\in \widehat{M(\R)}$ satisfying $\dim(W_\xi\otimes\Lambda^p\pg^\ast\otimes V_\tau)^{K^M_\infty}\neq 0$. \end{proof} \begin{lem}\label{lem-finite} For every $R\ge 0$, the number of $\pi\in \Pi_{\di}(M(\A))$ with $\lambda_{\pi_\infty}\ge -R$ and $\cA_\pi^2(P)^{K_f,\nu}\neq 0$ for some $\nu\in\cT$ is finite. \end{lem} \begin{proof} To prove the lemma, it suffices to show that for every $R\ge 0$ we have \begin{equation}\label{estim11} \sum_{\substack{\pi\in\Pi_{\di}(M(\A))\\-\lambda_{\pi_\infty}\le R}} \dim(\cA_\pi^2(P)^{K_f,\nu})<\infty. \end{equation} By passing to a subgroup of finite index, we may assume that $K_f=\prod_{p<\infty}K_p$. Let $K_{M,f}=K_f\cap M(\A_f)$ and $K_{M.\infty}=K_\infty\cap M(\R)$. For $\pi\in\Pi(M(\A))$ and $\tau\in\Pi(K_{M,\infty})$ let $\H_{\pi_\infty}(\tau)$ denote the $\tau$-isotypical subspace of the representation space $\cH_{\pi_\infty}$. Arguing as in the proof of Proposition 3.5 in \cite{Mu1}, it follows that in order to establish \eqref{estim11}, it suffices to show that for every $\tau\in\Pi(K_{M,\infty})$ \[ \sum_{\substack{\pi\in\Pi_{\di}(M(\A))\\-\lambda_{\pi_\infty}\le R}} \dim(\cH_{\pi_f}^{K_{M,f}})\cdot\dim(\cH_{\pi_\infty}(\tau))<\infty. \] Let $\Gamma_M\subset M(\R)$ be an arithmetic subgroup. Let $\Omega_{M(\R)^1}$ be the Casimir element of $M(\R)^1$ and let $A_\tau$ be the differential operator in $C^\infty(\Gamma_M\bs M(\R)^1;\tau)$ which is induced by $-\Omega_{M(\R)^1}$. Let $\bar A_\tau$ be its self-adjoint extension of $A_\tau$ in $L^2$. Proceeding as in the proof of Lemma 3.2 of \cite{Mu1}, it follows that it suffices to show that for every $R\ge 0$, the number of eigenvalues $\lambda_i$ of $\bar A_\tau$ (counted with multiplicities), satisfying $\lambda_i\le R$ is finite. Let $\Delta_\tau$ be the Bochner-Laplace operator and let $\Lambda_\tau$ be the Casimir eigenvalue of $\tau$. Then $\Delta_\tau=A_\tau+\Lambda_\tau\operatorname{Id}$. Since $\Delta_\tau\ge 0$ and by \cite{Mu3}, the counting function of the eigenvalues has a polynomial bound, the lemma follows. \end{proof} Now we can begin with the estimation of \eqref{specside4}. Using that $M(P,\pi,s)$ is unitary, it follows that \eqref{specside4} can be estimated by \begin{equation}\label{est-specside} \begin{split} \sum_{\pi\in\Pi_{\di}(M(\A))} \sum_{\nu\in\cT}& \dim\left(\cA^2_\pi(P)^{K_f,\nu}\right)\\ &\cdot e^{-t(\tau(\Omega)-c(\pi_\infty))}\int_{i(\af^G_{L_s})^*}e^{-t\|\lambda\|^2}\| \widetilde\Delta_{\mathcal{X}_{L_s}(\underline{\beta})}(P,\pi,\nu,\lambda)\|\;d\lambda. \end{split} \end{equation} For $\pi\in \Pi(M(\A))$ denote by $\lambda_{\pi_\infty}$ the Casimir eigenvalue of the restriction of $\pi_\infty$ to $M(\R)^1$. Given $\lambda>0$, let \[ \Pi_{\di}(M(\A);\lambda):=\left\{\pi\in\Pi_{\di}(M(\A))\colon |\lambda_{\pi_\infty}|\le\lambda\right\}. \] Let $d=\dim M(\R)^1/K_\infty^M$. As in \cite[Proposition 3.5]{Mu1} it follows that for every $\nu\in\Pi(K_\infty)$ there exists $C>0$ such that \begin{equation}\label{estim10} \sum_{\pi\in\Pi_{\di}(M(\A);\lambda)}\dim\cA^2_\pi(P)^{K_f,\nu}\le C(1+\lambda^{d/2}) \end{equation} for all $\lambda\ge 0$. Next we estimate the integral in \eqref{est-specside}. Let $\underline{\beta}=(\beta_1^\vee,\dots,\beta_m^\vee)$ and $\mathcal{X}_{L_s}(\underline{\beta})= (Q_1,\dots,Q_m)\in\Xi_{L_s}(\underline{\beta})$ with with $Q_i=\langle P_i,P_i'\rangle$, $P_i|^{\beta_i}P_i'$, $i=1,\dots,m$. Using the definition \eqref{intertw3} of $\Delta_{\mathcal{X}_{L_s}(\underline{\beta})}(P,\pi,\nu,\lambda)$, it follows that we can bound the integral by a constant multiple of \begin{equation}\label{est-integral} \dim(\nu)\int_{i(\af^G_{L_s})^*}e^{-t\|\lambda\|^2}\prod_{i=1}^m\left\| \delta_{P_i|P_i^\prime}(\lambda)\Big|_{\cA^2_\pi(P_i^\prime)^{K_f,\nu}}\right\| \;d\lambda. \end{equation} We introduce new coordinates $s_i:=\langle\lambda,\beta_i^\vee\rangle$, $i=1,\dots,m$, on $(\af^G_{L_s,\C})^\ast$. Using \eqref{normalization}, we can write \begin{equation}\label{delta} \delta_{P_i|P_i^\prime}(\lambda)=\frac{n^\prime_{\beta_i}(\pi,s_i)}{n_{\beta_i}(\pi,s_i)} +j_{P_i^\prime}\circ(\operatorname{Id}\otimes R_{P_i|P_i^\prime}(\pi,s_i)^{-1} R^\prime_{P_i|P_i^\prime}(\pi,s_i))\circ j_{P_i^\prime}^{-1}. \end{equation} Put \[ \cA^2_\pi(P)^{K_f,\cT}=\bigoplus_{\nu\in\cT}\cA^2_\pi(P)^{K_f,\nu}, \] where $\cT$ is defined by \eqref{def-F}. It follows from \cite[Theorem 5.3]{Mu2} that there exist $N,k\in\N$ and $C>0$ such that \begin{equation}\label{log-deriv} \int_{i\R}\left|\frac{n^\prime_{\beta_i}(\pi,s)}{n_{\beta_i}(\pi,s)}\right| (1+|s|^2)^{-k}\;ds\le C(1+\lambda_{\pi_\infty}^2)^N,\;i=1,\dots,m, \end{equation} for all $\pi\in \Pi_{\di}(M(\A))$ with $\cA^2_\pi(P)^{K_f,\cT}\neq 0$. Furthermore, for $G=\operatorname{GL}(n)$ it follows from \cite[Proposition 0.2]{MS} that there exist $k, C>0$ such that \begin{equation}\label{log-deriv1} \int_{i\R}\left\|R_{P_i|P_i^\prime}(\pi,s)^{-1} R^\prime_{P_i|P_i^\prime}(\pi,s)\Big|_{\cA^2_\pi(P_i^\prime)^{K_f,\nu}}\right\|(1+|s|^2)^{-k}\, ds \le C,\; i=1,\dots,m, \end{equation} for all $\nu\in\cT$ and $\pi\in \Pi_{\di}(M(\A))$ with $\cA^2_\pi(P)^{K_f,\nu}\neq 0$. To show that \eqref{log-deriv1} also holds for $G=\operatorname{SL}(n)$, we proceed as in the proof of \cite[Lemma 5.14]{FLM2}. Combining \eqref{delta}, \eqref{log-deriv} and \eqref{log-deriv1}, it follows that for $t\ge 1$ we have \[ \int_{i(\af^G_{L_s})^*}e^{-t\|\lambda\|^2}\prod_{i=1}^m\left\| \delta_{P_i|P_i^\prime}(\lambda)\Big|_{\cA^2_\pi(P_i^\prime)^{K_f,\nu}}\right\| \;d\lambda\ll (1+\lambda_{\pi_\infty}^2)^{mN} \] for all $\pi\in \Pi_{\di}(M(\A))$ with $\cA^2_\pi(P)^{K_f,\cT}\neq 0$. Thus \eqref{est-specside} can be estimated by a constant multiple of \begin{equation}\label{est-specside1} \sum_{\pi\in\Pi_{\di}(M(\A))}\sum_{\nu\in\cT}\dim\left(\cA^2_\pi(P)^{K_f,\nu}\right) (1+\lambda_{\pi_\infty}^2)^{mN} e^{-t(\tau(\Omega)-c(\pi_\infty))}. \end{equation} First assume that $M$ is a proper Levi subgroup. Note that by \eqref{casimir4} one has \begin{equation}\label{eigenv} \tau(\Omega)-c(\pi_\infty)=\tau(\Omega)+\|\rho_\af\|^2-\lambda_{\pi_\infty}. \end{equation} Together with Lemma \ref{lem-finite}, it follows that there exists $\lambda_0> 0$ such that \[ \tau(\Omega)-c(\pi_\infty)\ge |\lambda_{\pi_\infty}|/2 \] for all $\pi\in \Pi_{\di}(M(\A))$ with $\cA^2_\pi(P)^{K_f,\cT}\neq 0$ and $|\lambda_{\pi_\infty}|\ge \lambda_0$. Decompose the sum over $\pi$ in \eqref{est-specside1} in two summands $\Sigma_1(t)$ and $\Sigma_2(t)$, where in $\Sigma_1(t)$ the summation runs over all $\pi$ with $|\lambda_{\pi_\infty}|\le \lambda_0$. Using \eqref{estim10}, it follows that for $t\ge 1$ \[ \Sigma_2(t)\ll e^{-t|\lambda_0|/2}. \] Since $\Sigma_1(t)$ is a finite sum by Lemma \ref{lem-finite}, both in $\pi$ and $\nu$, it follows from Lemma \ref{est-casim} that there exists $c>0$ such that \[ \Sigma_1(t)\ll e^{-ct} \] for $t\ge 1$. Putting everything together it follows that for every $\tau\in \operatorname{Rep}(G(\R))$ such that $\tau\not\cong\tau_\theta$ and every proper Levi subgroup $M$ of $G$ there exists $c>0$ such that \begin{equation}\label{trspec} J_{\spec,M}(\phi_t^{\tau,p})=O(e^{-ct}) \end{equation} for $t\ge 1$. Now consider the case $M=G$. Then $c(\pi_\infty)=\pi_\infty(\Omega)$ and we need to show that \begin{equation}\label{positiv1} \tau(\Omega)-\pi_\infty(\Omega)>0 \end{equation} for all $\pi\in\Pi_{\di}(G(\A))$ with $\dim\cH_\pi^{K_f,\cT}\neq 0$. This follows from \cite[Lemma 4.1]{BV}, and we can proceed as in the case $M\neq G$ to prove that \[ J_{\spec,G}(\phi_t^{\tau,p})=O(e^{-ct}) \] for $t\ge 1$. Combined with \eqref{trspec} we obtain \begin{prop}\label{asympinf} Let $\tau\in\operatorname{Rep}(G(\R))$. Assume that $\tau\not\cong\tau_\theta$. Then there exists $c>0$ such that \[ J_{\spec}(\phi_t^{\tau,p})=O\left(e^{-ct}\right) \] for all $t\ge 1$ and $p=0,\dots,n$. \end{prop} \subsection{Definition of analytic torsion} Applying the trace formula \eqref{tracef1}, we get \[ \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)=O(e^{-ct}),\quad\text{as}\; t\to\infty, \] which is the proof of Theorem \ref{theo-lt}. Using \eqref{regtr-heat}, \eqref{equ-kernel2} and Theorem \ref{prop-asymp3}, it follows that as $t\to +0$, there is an asymptotic expansion of the form \[ \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right)\sim t^{-d/2}\sum_{j=0}^\infty a_jt^{j}+ t^{-(d-1)/2}\sum_{j=0}^\infty\sum_{i=0}^{r_j} b_{ij}t^{j/2} (\log t)^i. \] Thus the corresponding zeta function $\zeta_p(s;\tau)$, defined by the Mellin transform \begin{equation}\label{zetafct} \zeta_p(s;\tau):=\frac{1}{\Gamma(s)}\int_0^\infty \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right) t^{s-1}\; dt. \end{equation} is holomorphic in the half-plane $\Re(s)>d/2$ and admits a meromorphic extension to the whole complex plane. It may have a pole at $s=0$. Let $f(s)$ be a meromorphic function on $\C$. For $s_0\in\C$ let \[ f(s)=\sum_{k\ge k_0}a_k(s-s_0)^k \] be the Laurent expansion of $f$ at $s_0$. Put $\operatorname{FP}_{s=s_0}:=a_0$. Now we define the analytic torsion $T_X(\tau)\in\R^+$ by \begin{equation}\label{analtor3} \log T_X(\tau)=\frac{1}{2}\sum_{p=0}^d (-1)^p p \left(\operatorname{FP}_{s=0}\frac{\zeta_p(s;\tau)}{s}\right). \end{equation} Put \begin{equation}\label{altheat} K(t,\tau):=\sum_{p=1}^d (-1)^p p \operatorname{Tr}_{\reg}\left(e^{-t\Delta_p(\tau)}\right). \end{equation} Then $K(t,\tau)=O(e^{-ct})$ as $t\to\infty$ and the Mellin transform \[ \int_0^\infty K(t,\tau)t^{s-1} dt \] converges absolutely and uniformly on compact subsets of $\Re(s)>d/2$ and admits a meromorphic extension to $\C$. Moreover, by \eqref{analtor3} we have \begin{equation}\label{analtor1} \log T_X(\tau)=\operatorname{FP}_{s=0}\left(\frac{1}{\Gamma(s)}\int_0^\infty K(t,\tau)t^{s-1} dt\right). \end{equation} Let \[ \phi_t^\tau:=\sum_{p=1}^d (-1)^p p \phi_t^{\tau,p} \quad\text{and}\quad k_t^\tau:=\sum_{p=1}^d (-1)^p p h_t^{\tau,p}. \] Then by \eqref{regtr-heat} we have \begin{equation}\label{alt-reg-tr} K(t,\tau)=J_{\spec}(\phi_t^\tau). \end{equation} For $\pi\in\Pi(G(\R))$ let $\Theta_\pi$ be the global character. Then we get \begin{equation}\label{alt-reg-tr1} J_{\spec,G}(\phi^\tau_t)=\sum_{\pi\in\Pi_{\di}(G(\A))} m(\pi)\dim\left(\cH_{\pi_f}^{K_f} \right)\Theta_{\pi_\infty}(k_t^\tau). \end{equation} For $n\in\N$, $n\ge 2$, let $\delta_n:=\operatorname{rank}_\C \operatorname{SL}(n)- \operatorname{rank}_\C\operatorname{SO}(n)$ be the fundamental rank of $\operatorname{SL}(n)$. \begin{lem} For $G=\operatorname{GL}(n)$ and $n\ge 5$ we have $J_{\spec,G}(\phi^\tau_t)=0$. \end{lem} \begin{proof} Let $Q$ be a standard cuspidal parabolic subgroup of $G(\R)$. Let $Q=M_QA_QN_Q$ be the Langlands decomposition of $Q$. Let $(\xi,W_\xi)$ be a discrete series representation of $M_Q$ and let $\nu\in\af_{Q,\C}^\ast$. Let $\pi_{\xi,\nu}$ be the induced representation. By \cite[Proposition 4.1]{MP2} we have $\Theta_{\xi,\nu}(k_t^\tau)=0$, if $\dim \af_Q\ge 2$. If $\delta_n\ge 2$, it follows that $\dim\af_Q\ge 2$ for every cuspidal parabolic subgroup $Q$ of $G(\R)$. Thus $\Theta_{\xi,\nu}(k_t^\tau)=0$ for all cuspidal parabolic subgroups $Q$ and pairs $(\xi,\nu)$ as above. Now observe that for $\operatorname{GL}(n)$ the $R$-group is trivial. Therefore, it follows from \cite[section 2.2]{De} that the Grothendieck group of all admissible representations of $G(\R)$ is generated by the induced representations $\pi_{\xi,\nu}$ as above. Hence $\Theta_\pi(k_t^\tau)=0$ for all $\pi\in\Pi(G(\R))$. If $n\ge 5$, then $\delta_n \ge 2$ and the lemma follows from \eqref{alt-reg-tr1}. \end{proof} \begin{remark} If $\Gamma$ is cocompact and $n\ge 5$, then it follows that $T_X(\tau)=1$. In the noncompact case this need not be true. In \cite{MP1} the case of finite volume hyperbolic manifolds has been studied. It has been shown that in even dimensions, the renormalized analytic torsion has a simple expression, but is not trivial. This includes the case of $\operatorname{SL}(2)$. \end{remark} \section{The case $G=\operatorname{GL}(3)$}\label{sect-gl3} \setcounter{equation}{0} If $G=\operatorname{GL}(3)$, the weight functions are explicitly given by \eqref{unipot3}- \eqref{unipot5}. Using the explicit form of the weight function, we can extract more precise information about the pole at $s=0$. To this end we need to show that the coefficients $c_{ij}(\nu)$ in \eqref{unipo-asympexp} with $j=d-k$ and $i=1,\dots,r_{d-k}$ vanish for the corresponding orbital integrals. In our case $d=5$. Now consider the first integral \eqref{unipot3}. Then $k=2$ and the weight function is $\log(y^2+z^2)$. Thus the highest power with which $\log t$ occurs in the asymptotic expansion of \eqref{unipot3} is 1. This means that $c_{13}(\nu)$ is the only coefficient that we need to consider. It is of the form \eqref{coeffi}. We are in the case $i=r_j$. Hence \[ c_{13}(\nu)=\int_{\R^2}e^{-a\|x\|^2}p(x)\;dx, \] where $p(x)$ is a homogeneous polynomial. Moreover, from its construction it follows that $p(x)$ is odd, i.e., $p(-x)=-p(x)$. Hence $c_{13}(\nu)=0$. Thus the asymptotic expansion of the first integral has the form \begin{equation}\label{unipot6} J_{M_1}(1,h_t^\nu)\sim t^{-3/2}\sum_{j=0}^\infty a_j(\nu) t^{j/2} +t^{-3/2}\sum_{k=0}^\infty b_k(\nu) t^{k/2}\log t, \end{equation} as $t\to +0$, and $b_3(\nu)=0$. Now consider the second integral \eqref{unipot4}. Then $k=3$. By \eqref{unipot4} we need only to consider $c_{12}(\nu)$, which we denote by $c_2(\nu)$. Let $p_1(x)$ and $p_2(x)$ be the polynomials occurring on the right hand side of \eqref{taylor6} and $a^\nu_j(g)$ the coefficients on the right hand side of \eqref{asympexp1}. If we collect all possible contributions, we get \begin{equation}\label{const1} \begin{split} c_{2}(\nu)=&a_0^\nu(1)\int_{\R^3}p_2(x) e^{-\|x\|^2}\;dx+\sum_{i=1}^3 \frac{\partial}{\partial x_i}a_0^\nu(n(x))\big|_{x=0} \int_{\R^3} x_ip_1(x)e^{-\|x\|^2}\;dx\\ &+ \sum_{i,j=1}^3\frac{\partial^2}{\partial x_i\partial x_j} a_0^\nu(n(x))\big|_{x=0}\int_{R^3} x_i x_j e^{-\|x\|^2} dx+ a_1^\nu(1)\int_{\R^3} e^{-\|x\|^2}\;dx. \end{split} \end{equation} By definition we have \[ p_1(x)=\sum_{|\alpha|=3} D^\alpha r^2(x)\big|_{x=0} x^\alpha. \] Now recall that for $g\in\operatorname{SL}(n,\R)$ the distance $r(g(x_0),x_0)$ is given as follows. Let $\lambda_1,\dots,\lambda_n$ be the eigenvalues of the positive definite matrix $g^\top\cdot g$. Then \[ r(g(x_0),x_0)^2=\sum_{i=1}^n (\log\lambda_i)^2. \] An explicit computation shows that \[ r^2(x_1,x_2,0)=2\log^2\left(1+\frac{x_1^2+x_2^2}{2}+\sqrt{x_1^2+x_2^2+ \frac{(x_1^2+x_2^2)^2}{4}}\right). \] Thus $r(x_1,x_2,0)$ is even in $x_1$ and $x_2$. The same holds for $r(x_1,0,x_3)$ and $r(0,x_2,x_3)$. This implies that for $\alpha\neq(1,1,1)$ we have $D^\alpha r^2(x)\big|_{x=0}=0$. Finally note that \[ \int_{\R^3} x_i x_1 x_2 x_3 e^{-\|x\|^2} dx=0,\quad\text{and}\quad \int_{\R^3} x_i x_j e^{-\|x\|^2}dx=0,\; i\neq j. \] Thus \eqref{const1} is reduced to \begin{equation}\label{const2} \begin{split} c_{2}(\nu)=&a_0^\nu(1)\int_{\R^3}p_2(x) e^{-\|x\|^2}\;dx + a_1^\nu(1)\int_{\R^3} e^{-\|x\|^2}\;dx\\ &+\sum_{i=1}^3\frac{\partial^2}{\partial x_i^2} a_0^\nu(n(x))\big|_{x=0}\int_{\R^3}x_i^2 e^{-\|x\|^2} dx. \end{split} \end{equation} Thus for the second integral we get an asymptotic expansion of the form \begin{equation}\label{unipot7} J_{M_1}(u(1,0,0), h_t^\nu)\sim t^{-1}\sum_{j=0}^\infty a_j(\nu) t^{j/2} + t^{-1}\sum_{k=0}^\infty c_k(\nu) t^{k/2}\log t \end{equation} with $c_2(\nu)$ given by \eqref{const2}. Finally consider the integral \eqref{unipot5}. Again $k=3$. By \eqref{unipot5} we only need to consider $c_{12}(\nu)$ and $c_{22}(\nu)$. By the same considerations as in the previous case, it follows that $c_{22}(\nu)=c_2(\nu)$. Furthermore, $c_{12}(\nu)$ has the same form as $c_2(\nu)$, except that the integrals contain in addition some factors $\log|x_i|$ for $i=1,2,3$. Thus we obtain \begin{equation}\label{unipot8} J_{M_0}(1,h_t^\nu)\sim t^{-1}\sum_{j=0}^\infty a_j(\nu) t^{j/2}+ t^{-1}\sum_{k=0}^\infty c_{1k}(\nu)t^{k/2}\log t +t^{-1}\sum_{l=0}^\infty c_{2l}(\nu)t^{l/2}(\log t)^2, \end{equation} with $c_{22}(\nu)=c_2(\nu)$, where $c_2(\nu)$ is given by \eqref{const2}, and $c_{12}(\nu)$ is given by a similar formula as described above. Now we specialize $\nu$ to $\nu_p(\tau)$, which is defined by \eqref{repr4}. \begin{lem}\label{lem-res1} Let $(\tau,V_\tau)$ be a finite dimensional representation of $G(\R)$. We have \[ \sum_{p=1}^5 (-1)^p p\cdot a_0^{\nu_p(\tau)}(1)=0, \quad \sum_{p=1}^5 (-1)^pp\cdot a_1^{\nu_p(\tau)}(1)=0. \] \end{lem} \begin{proof} By \eqref{leadcoeff} we have $a_0^{\nu_p(\tau)}=\dim(\Lambda^p\pg^\ast\otimes V_\tau)= \binom{5}{p}\cdot\dim V_\tau$. Now observe that $\sum_{p=1}^5 (-1)^p p\binom{5}{p} =0$. This proves the first statement. For the second statement we note that by \eqref{trace5} we have $a_1^{\nu_p(\tau)}(1)=\operatorname{tr}(\phi_1^{\nu_p(\tau)}(x_0,x_0))$, and by \eqref{asympexp}, $\phi_1^{\nu_p(\tau)}(x_0,x_0)$ is the second coefficient of the asymptotic expansion as $t\to +0$ of $\operatorname{tr} K^{\nu_p(\tau)}(t,x_0,x_0)$. Using the known structure of the coefficient, we get \begin{equation}\label{heat-coeff} a_1^{\nu_p(\tau)}(1)=-\frac{R\cdot\dim(\tau)}{6}\left\{\binom{5}{p} -6\binom{3}{p-1}\right\}, \end{equation} where $R$ is the scalar curvature (which is constant) and it is understood that $\binom{m}{p}=0$, if $p<0$ or $p>m$. For $\tau=1$, this follows from \cite[Theorem 4.1.7, (b)]{Gi}. It is easy to extend this to the twisted case. Using \eqref{heat-coeff}, the second statement follows. \end{proof} \begin{lem}\label{lem-res2} For every finite dimensional representation $(\tau,V_\tau)$ of $G(\R)$ we have \[ \frac{\partial^2}{\partial x_i^2}\bigg|_{x=0}\left(\sum_{p=1}^5 (-1)^p p\; a_0^{\nu_p(\tau)}(n(x))\right)=0. \] for $i=1,2,3$. \end{lem} \begin{proof} We consider the derivative with respect to $x_1$. Let \[ n_1(u)=\begin{pmatrix} 1&u&0\\0&1&0\\0&0&1\end{pmatrix}. \] Then \[ \frac{\partial^2}{\partial x_1^2}a_0^{\nu_p(\tau)}(n(x))\big|_{x=0}= \frac{\partial^2}{\partial u^2}a_0^{\nu_p(\tau)}(n_1(u))\big|_{u=0}. \] By \eqref{leadcoeff} we have \[ a_0^{\nu_p(\tau)}(n_1(u))=\operatorname{tr}(\nu_p(\tau)(k(u))\cdot j(x_0,n_1(u)x_0), \] where $k(u):=k(n_1(u))\in\operatorname{SO}(3)$ is determined by \eqref{cartan2}. Furthermore, by \eqref{repr4} we have \[ \operatorname{tr}(\nu_p(\tau)(k(u))= \operatorname{tr}(\Lambda^p\operatorname{Ad}_\pg^\ast(k(u))\cdot\operatorname{tr}(\tau(k(u)). \] Let \[ S:=\left\{A\in\operatorname{Mat}_3(\R)\colon A=A^t,\;\operatorname{tr}(A)=0\right\}, \] equipped with the inner product \[ \langle Y_1,Y_2\rangle=\operatorname{Tr}(Y_1Y_2),\quad Y_1,Y_2\in S. \] Then $\pg\cong S$ as inner product spaces. Moreover, the adjoint representation $\operatorname{Ad}_{\pg}$ of $\operatorname{SO}(3)$ on $S$ is given by \begin{equation}\label{adjoint} \operatorname{Ad}_{\pg}(k)Y=k\cdot Y\cdot k^\ast,\quad k\in\operatorname{SO}(3),\; Y\in S. \end{equation} With respect to this isomorphism, $k(u)$ is determined as follows. Let $A(u):=n_1(u)n_1(u)^\ast$. Then $A(u)=A(u)^t$ and $A(u)>0$. Let $S(u)=A(u)^{-1/2}$. Then $k(u)=S(u)\cdot n_1(u)$. Note that $k(u)$ is a block diagonal matrix of the form \[ \begin{pmatrix} r(\theta)&0\\0&1\end{pmatrix}, \] where $r(\theta)\in\operatorname{SO}(2)$ is the rotation by the angle $\theta$. Let \[ Y_1=\begin{pmatrix}-1/2&0&0\\0&-1/2&0\\0&0&1\end{pmatrix}. \] Then with respect to \eqref{adjoint} we have $\operatorname{Ad}_{\pg}(k(u))(Y_1)=Y_1$. Let $S_1:=\R Y_1$ and $S_0=S_1^\perp$. Then the decomposition $S=S_0\oplus S_1$ is invariant under $\operatorname{Ad}_{\pg}(k(u))$ and $\operatorname{Ad}_{\pg}(k(u))|_{S_1}=\operatorname{Id}$. Let $T(u):=\operatorname{Ad}_{\pg}(k(u))|_{S_0}$. Then we have \begin{equation}\label{det} \sum_{p=1}^5 (-1)^p p \;\operatorname{tr}(\Lambda^p\operatorname{Ad}_{\pg}^\ast(k(u)))=\sum_{p=0}^5 (-1)^p \operatorname{tr}(\Lambda^p T(u))=\det(\operatorname{Id}-T(u)). \end{equation} For $\lambda\in\C$ let \begin{equation}\label{det1} f(\lambda,u):=\det(\lambda\operatorname{Id}-T(u)),\quad \lambda\in\C,\;u\in\R. \end{equation} Recall that $T(u)$ is unitary. So every eigenvalue $\mu$ of $T(u)$ satisfies $|\mu|=1$. Assume that $|\lambda|\neq 1$. Then $f(\lambda,u)\neq 0$ for all $u\in\R$ and \begin{equation}\label{logder1} \frac{\partial}{\partial u}\log f(\lambda,u) =-\operatorname{tr}(T^\prime(u)(\lambda\operatorname{Id}-T(u))^{-1}), \end{equation} where $T^\prime(u)=\frac{d}{du}T(u)$. Note that $f(\lambda,0)=\det(\lambda\operatorname{Id}-T(0))=(\lambda-1)^4$. Thus \[ \frac{\partial}{\partial u}f(\lambda,u)\big|_{u=0} =-(\lambda-1)^3\operatorname{tr}(T^\prime(0)). \] Since $T(u)$ is orthogonal, it follows that $\operatorname{tr}(T^\prime(0))=0$, and therefore \begin{equation}\label{firstder} \frac{\partial}{\partial u}f(\lambda,u)\big|_{u=0}=0. \end{equation} Using \eqref{logder1}, we get \begin{equation}\label{deriv} \begin{split} \frac{\partial^2}{\partial u^2}f(\lambda,u)&=-\frac{\partial}{\partial u} f(\lambda,u)\cdot\operatorname{tr}(T^\prime(u)(\lambda\operatorname{Id}-T(u))^{-1})\\ &-f(\lambda,u)\operatorname{tr}(T^{\prime\prime}(u)(\lambda\operatorname{Id}-T(u))^{-1})\\ &-f(\lambda,u)\operatorname{tr}(T^\prime(u)(\lambda\operatorname{Id}-T(u))^{-1}T^\prime(u) (\lambda\operatorname{Id}-T(u))^{-1}). \end{split} \end{equation} Using \eqref{firstder}, we obtain \[ \frac{\partial^2}{\partial u^2}f(\lambda,u)\big|_{u=0}= -(\lambda-1)^3\operatorname{tr}(T^{\prime\prime}(0))-(\lambda-1)^2\operatorname{tr}(T^\prime(0)^2). \] Since $f(\lambda,u)$ is a polynomial in $\lambda$, it follows that this equality holds for all $\lambda\in\C$. In particular, we get \[ \frac{\partial^2}{\partial u^2}f(1,u)\big|_{u=0}=0. \] Combined with \eqref{det} and the definition of $f(\lambda,u)$, the statement follows for $i=1$. The proof of the other cases is similar. \end{proof} Using \eqref{unipot7}, \eqref{unipot8} and Lemmas \ref{lem-res1} and \ref{lem-res2}, it follows that \[ \sum_{p=1}^5 (-1)^p p\;\zeta_p(s;\tau) \] is holomorphic at $s=0$. Thus in this case we can define $\log T_{X(K_f)}(\tau)$ by \[ \log T_{X(K_f)}(\tau)=\frac{1}{2}\frac{d}{ds}\left(\sum_{p=1}^5 (-1)^p p\; \zeta_p(s;\tau)\right)\bigg|_{s=0}. \] \section{Example: Classes of finite order for $\operatorname{GL}(2)$ and $\operatorname{GL}(3)$}\label{sec:examples} \label{sec-finite-ord} \setcounter{equation}{0} In order to remove the assumption that $\Gamma\subseteq \Gamma(N)$ for some $N\ge 3$, we need to understand distributions $J_{\of}$ appearing in the coarse geometric expansion of the trace formula for which the equivalence classes $\of$ which are not necessarily unipotent. Let $K_f$ be an arbitrary subgroup of $G(\widehat{\Z})$ of finite index and let $f=f_{\infty}\cdot 1_{K_f}\in C_c^{\infty}(G(\A)^1)$ with $f_{\infty}\in C_c^{\infty} (G(\R)^1)$ and $1_{K_f}\in C_c^{\infty}(G(\A_f))$ the characteristic function of $K_f$. In this situation, more than just the unipotent orbit may contribute non-trivially to the coarse geometric expansion. The equivalence classes $\of\in\mathcal{O}$ are in bijection with semisimple orbits in $G(\Q)$. Hence there is a canonical bijection between $\mathcal{O}$ and monic polynomial of degree $n$ with rational coefficients and non-vanishing constant term if $G=\operatorname{GL}(n)$ by sending the semisimple conjugacy class to its characteristic polynomial. We may therefore speak of the characteristic polynomial and the eigenvalues of a class $\of$. The following lemma is explained in the proof of \cite[Lemma 5.1]{LM}. \begin{lem} We can choose a $K_{\infty}$-bi-invariant neighborhood $\omega\subseteq G(\R)^1$ of $K_{\infty}$ such that if $\of\in\mathcal{O}$ is such that there exists $f_{\infty}\in C_c^{\infty}(G(\R)^1)$ supported in $\omega$ with $J_\of(f_{\infty}\cdot 1_{K_f})\neq0$, then the eigenvalues of $\of$ are all roots of unity (over some algebraic closure of $\Q$). \end{lem} Let $\mathcal{O}_1$ denote the set of all $\of\in\mathcal{O}$ whose eigenvalues (in some algebraic closure of $\Q$) are all roots of unity. Note that this set is finite. By the preceding lemma we can choose a bi-$K_{\infty}$-invariant $f_{\infty}\in C_c^{\infty}(G(\R)^1)$ with $f(1)=1$ and \[ J_{\geo}(f_{\infty}\cdot 1_{K_f})=\sum_{\of\in \mathcal{O}_1} J_{\of}(f_{\infty}\cdot 1_{K_f}). \] Let $\of\in\mathcal{O}_1$, and let $\sigma\in G(\Q)\cap \of$ be a semisimple representative for $\of$. Then $\sigma$ is in $G(\R)$ conjugate to some element $\sigma_\infty$ in $\operatorname{O}(n)$. For each $\of$ and $f\in C_c^{\infty}(G(\A)^1)$ we have the fine expansion \[ J_{\of}(f)=\sum_{(M,\gamma)} a^M(\gamma, S) J_M(\gamma, f), \] where $S$ is a sufficiently large finite set of places of $\Q$ with $\infty\in S$, $a^M(\gamma, S)$ are certain global coefficients as defined in \cite{Ar7}, $(M, \gamma)$ runs over all pairs of Levi subgroups $M$ containing $M_0$ and $\gamma$ over representatives of the $M(\Q)$-conjugacy classes in $M(\Q)\cap \of$, and $J_M(\gamma, f)$ are $S$-adic weighted orbital integrals. Since the (finite) set $\mathcal{O}_1$ and the set $S$ are fixed in our setting, the value of the coefficients $a^M(\gamma, S)$ is not relevant for us. \subsection{Orbits of finite order for $\operatorname{GL}(2)$} If $G=\operatorname{GL}(2)$, then each $\of\in\mathcal{O}_1$ is represented by one of the following semisimple elements: \[ \sigma_0^\pm=\pm \begin{pmatrix}1&0\\0&1\end{pmatrix}, \quad \sigma_1= \begin{pmatrix}1&0\\0&-1\end{pmatrix}, \quad \sigma_i^\pm= \pm\begin{pmatrix}\cos\theta_i&\sin\theta_i\\-\sin\theta_i&\cos\theta_i\end{pmatrix},\quad i=2,3, \] with $\theta_2=\pi/2$ and $\theta_3=\pi/3$. We accordingly write $\of_i$ or $\of_i^\pm$ for the associated equivalence classes. Note that $\sigma_1, \sigma_2^\pm, \sigma_3^\pm$ are all regular semisimple so that the associated equivalence class is in fact equal to the conjugacy class of the respective element. (In fact, $\of_2^+=\of_2^-$, but we keep the superscript to make notation more uniform.) Moreover, since we assume that our test function $f$ is $K_\infty$-invariant, $J_{\operatorname{unip}}(f)=J_{\of_0^-}(f)$ for $\of_0^-$ the class attached to $\sigma_0^-$. Hence we only need to consider the regular elements. The element $\sigma_1$ is the only of the remaining elements which is split over $\R$. Since it is regular, the distribution $\of_1$ is of a simple form, namely, \[ J_{\of_1}(f)=\int_{U_0(\A)} f(u^{-1} \sigma_1 u) v_T(u)\, du \] for every bi-$K_{\infty}$-invariant $f\in C_c^{\infty}(G(\A)^1)$. It follows that if $f= f_{\infty}\cdot 1_{K_f}$, then \[ J_{\of_1}(f)=a_1\int_{\R} f_{\infty}(u(x)) \log(1+x^2) \, dx + a_2 \int_{\R}f_{\infty} (u(x))\, dx \] where $u(x)=\left(\begin{smallmatrix}1&x\\0&1\end{smallmatrix}\right)$, and $a_1, a_2\in\R$ are suitable constants depending only on $K_f$. Using Taylor expansion of $\log(1+x^2)$ around $x=0$, we get, as in \S~\ref{sec-asymp}, that for every $N>0$, \[ J_{\of_1}(\tilde\phi_t^\nu) = t^{-(d-1)/2}\sum_{k=0}^N c_k t^{k/2} + O_N(t^{(N-d+1)/2}) \] for suitable coefficients $c_k$. The remaining classes are regular elliptic and non-split over $\R$. In particular, for each $i\in\{2,3\}$ we have \[ J_{\of_i^\pm}(f)= a_i\int_{G_{\sigma_i}(\R)\backslash G(\R)} f_{\infty}(g^{-1}\sigma_i^\pm g)\, dg \] for a suitable constant $a_i\in\R$ again depending only on $K_f$ and $\of_i^\pm$. We have for $i=2,3$ that $G_{\sigma_i^\pm}(\R)=Z(\R) K_{\infty}$ so that using $KAK$ decomposition we get \[ J_{\of_i^\pm}(f)= a_i\int_{0}^{\infty} f_{\infty}(a^{-1}\sigma_i^\pm a) \sinh(2X)\, dX \] where $a=e^X$. We can write \[ \pm\begin{pmatrix} \cos\theta_i & e^{-2X}\sin\theta_i\\-e^{-2X}\sin\theta_i & \cos\theta_i\end{pmatrix} = a^{-1}\sigma_i^\pm a = k_1 \begin{pmatrix} e^Y&0\\0&e^{-Y}\end{pmatrix} k_2 \] with suitable $k_1, k_2\in K_{\infty}$ and $Y\ge0$. Hence \[ \sinh(Y)= \alpha_i\sinh(2X) \] with $\alpha_i=\sqrt{2}\sin\theta_i$. Hence \[ Y^2= 4\alpha_i^2 X^2+ O_{\of_i^\pm}(X^4) \] around $0$. Since $r(a^{-1}\sigma_i^\pm a)= \|(Y, -Y)\|$, we therefore get \[ r^2(a^{-1}\sigma_i^\pm a) = 8\alpha_i^2 X^2 + O_{\of_i^\pm}(X^4). \] Using the Taylor expansion of $\sinh(2X)$, we get, as in \S~\ref{sec-asymp}, that for any $N$, \[ J_{\of_i^\pm}(\tilde\phi_t^\nu) = t^{-(d-2)/2}\sum_{k=0}^N c_k t^{k/2}+ O_{N,\of_i^\pm}(t^{(N-d+2)/2}) \] as $t\rightarrow0^+$ for suitable coefficients $c_k$ depending on $\of_i^\pm$. \subsection{Orbits of finite order for $\operatorname{GL}(3)$} For $G=\operatorname{GL}(3)$ the real weighted orbital integrals associated to $G(\R)$-conjugacy classes of elements in the classes in $\mathcal{O}_1$ can have a more complicated form. Each $\of\in\mathcal{O}_1$ has a semisimple element which in $G(\R)$ is conjugate to one of the following matrices: \[ \sigma_0^\pm=\pm\begin{pmatrix}1&0&0\\0&1&0\\0&0&1\end{pmatrix}, \sigma_1^\pm=\pm\begin{pmatrix}1&0&0\\0&1&0\\0&0&-1\end{pmatrix}, \sigma_i^{\pm,\pm}=\pm \begin{pmatrix}\cos\theta_i&\sin\theta_i&0\\-\sin\theta_i&\cos\theta_i&0\\0&0&\pm1\end{pmatrix},\quad i=2,3, \] with $\theta_2$ and $\theta_3$ as for $\operatorname{GL}(2)$. Again, we write $\of_i^\pm$, $i=0,1$, and $J_{\of_i^{\pm,\pm}}$, $i=2,3$, for the associated equivalence classes. We already understand the distributions $J_{\of_0^\pm}$. If $i=2,3$, the only semi-standard Levi subgroups of $G(\R)$ containing $\sigma_i^{\pm,\pm}$ are $M(\R):=\operatorname{GL}_2(\R)\times\operatorname{GL}_1(\R)$ (diagonally embedded in $G(\R)$) and $G(\R)$ itself. Let $P(\R)$ be the standard parabolic subgroup of $G(\R)$ with Levi component $M(\R)$. Moreover, $\of_i^{\pm,\pm}$ in fact equals the conjugacy class of $\sigma_i^{\pm,\pm}$. Hence we need to understand the real weighted orbital integrals $J_M(\sigma_i^{\pm,\pm},\tilde\phi_{t,\infty}^\nu)$ and $J_G(\sigma_i^{\pm,\pm},\tilde\phi_{t,\infty}^\nu)$. The latter integral equals \begin{align*} J_G(\sigma_i^{\pm,\pm},\tilde\phi_{t,\infty}^\nu) & =\int_{M(\R)_{\sigma_i^{\pm,\pm}}\backslash M(\R)} \int_{U(\R)} \tilde\phi_{t,\infty}^\nu(u^{-1}m^{-1}\sigma_i^{\pm,\pm} mu)\, du\, dm \\ & = \int_{M(\R)_{\sigma_i^{\pm,\pm}}\backslash M(\R)} \int_{U(\R)} \tilde\phi_{t,\infty}^\nu(m^{-1}\sigma_i^{\pm,\pm} mu)\, du\, dm \\ \end{align*} where we used the $O(n)$-conjugation invariance of $\tilde\phi_{t,\infty}^\nu$ and that the centralizer of $\sigma_i^{\pm,\pm}$ in $G(\R)$ and $M(\R)$ coincide. Hence for $t\rightarrow0$ we get an asymptotic expansion \[ J_G(\sigma_i^{\pm,\pm},\tilde\phi_{t,\infty}^\nu) =t^{-(d-4)/2}\sum_{k=0}^N C_k t^{k/2}+ O_N(t^{(N-d+4)/2}) \] for any $N>0$ where $C_k$ are certain coefficients depending on $\of_i^{\pm,\pm}$. The other weighted orbital integral is of the form \[ J_M(\sigma_i^{\pm,\pm},\tilde\phi_{t,\infty}^\nu) =\int_{M(\R)_{\sigma_i^{\pm,\pm}}\backslash M(\R)} \int_{U(\R)} \tilde\phi_{t,\infty}^\nu(u^{-1}m^{-1}\sigma_i^{\pm,\pm} mu) v_M(u)\, du\, dm \] where the weight function is given by \[ v_M(u)=\log(1+x^2+y^2) \] for $u=\begin{pmatrix}1&0&x\\0&1&y\\0&0&1\end{pmatrix}\in U(\R)$. A change of variables therefore gives \[ J_M(\sigma_i^{\pm,\pm},\tilde\phi_{t,\infty}^\nu) =\int_{M(\R)_{\sigma_i^{\pm,\pm}}\backslash M(\R)} \int_{U(\R)} \tilde\phi_{t,\infty}^\nu(m^{-1}\sigma_i^{\pm,\pm} mu) \log\left(1+\|w\|^2\right)\, du\, dm \] where \[ w=\left({\rm id} - m^{-1}\sigma_i^{\pm,\pm} m\right)^{-1}\begin{pmatrix}x\\ y\end{pmatrix}\in\R^2. \] Using the series expansion of $\log$ around $1$, this integral also has an asymptotic expansion in $t$ as $t\rightarrow 0$. Altogether, we get \[ J_{\of_i^{\pm,\pm}}(\tilde\phi_t^\nu) = t^{-(d-4)/2}\sum_{k=0}^N B_k t^{k/2}+ O_N(t^{(N-d+4)/2}) \] for suitable constants $B_k$ and any $N>0$. The remaining two classes $\of_1^\pm$ contain more elements than just the conjugates of $\sigma_1^\pm$. Hence we have more orbital integrals to consider. Let $u_1=\left(\begin{smallmatrix}1&1&0\\0&1&0\\0&0&1\end{smallmatrix}\right)$. Then we need to consider the weighted orbital integrals $J_{M_0}(\sigma_1^\pm, f_\infty)$, $J_L(\sigma_1^\pm, f_\infty)$, and $J_L(\sigma_1^\pm u_1, f_\infty)$, $L=M, G$, for $f_\infty=\tilde\phi_{t,\infty}^\nu$. For the invariant integrals we get \[ J_G(\sigma_1^\pm, f_\infty) =\int_{U(\R)} f_\infty(u^{-1}\sigma_1^\pm u) \, du = c_1\int_{U(\R)} f_\infty(\sigma_1^\pm u)\, du, \] and similarly, after a change of variables, \[ J_G(\sigma_1^\pm u_1, f_\infty) =c_2\int_{U_0(\R)} f_\infty(\sigma_1^\pm u) \, du. \] Here $c_1, c_2>0$ are suitable constants. The weighted orbital integrals can also be written as integrals over $U_0(\R)$, but against a non-invariant measure. For $J_{M_0}(\sigma_1^\pm,f_{\infty})$ it involves a weight function of the form $\log(1+x^2+y^2)$ as above and a linear function in $\log|a|$ if we write $u=\left(\begin{smallmatrix}1&a&0\\0&1&0\\0&0&1\end{smallmatrix}\right)\left(\begin{smallmatrix}1&0&x\\0&1&y\\0&0&1\end{smallmatrix}\right)$. Similarly, $J_M(\sigma_1^\pm u_1,f_\infty)$ equals an integral over $U_0(\R)$ against $\log(1+x^2+y^2)$ times the invariant measure, and $J_M(\sigma_i^\pm, f_\infty)$ equals the integral over $U(\R)$ against $\log(1+x^2+y^2)$ times the invariant measure on $U(\R)$. Proceeding similarly as before, one can then show that \[ J_{\of_1^\pm}(\tilde\phi_t^\nu) = t^{-(d-3)/2}\sum_{k=0}^N C_k t^{k/2}+ t^{-(d-3)/2}\sum_{k=0}^NB_kt^{k/2} \log t + O_N(t^{(N-d+3)/2}) \] for suitable constants $C_k, B_k$ and any $N$.
cf3092317c38ec8c204c24754ac23a844d7f7c38
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Recently, the LHCb collaboration performed the first full amplitude analysis of the decays $B^+\to J/\psi \phi K^+$ with $J/\psi\to\mu^+\mu^-$, $\phi\to K^+K^-$ with a data sample of 3 fb$^{-1}$ of $pp$ collision data collected at $\sqrt{s}=7$ and $8$ TeV with the LHCb detector, confirmed the two old particles $X(4140)$ and $X(4274)$ in the $J/\psi \phi$ mass spectrum with statistical significance $8.4\sigma$ and $6.0\sigma$, respectively, determined the quantum numbers to be $J^{PC} =1^{++}$ with statistical significance $5.7\sigma$ and $5.8\sigma$, respectively \cite{LHCb-4500-1606}. Moreover, the LHCb collaboration observed two new particles $X(4500)$ and $X(4700)$ in the $J/\psi \phi$ mass spectrum with statistical significance $6.1\sigma$ and $5.6\sigma$, respectively, determined the quantum numbers to be $J^{PC} =0^{++}$ with statistical significance $4.0\sigma$ and $4.5\sigma$, respectively \cite{LHCb-4500-1606}. The measured masses and widths are \begin{flalign} & X(4140) : M = 4146.5 \pm 4.5 ^{+4.6}_{-2.8} \mbox{ MeV} \, , \, \Gamma = 83 \pm 21 ^{+21}_{-14} \mbox{ MeV} \, , \nonumber\\ & X(4274) : M = 4273.3 \pm 8.3 ^{+17.2}_{-3.6} \mbox{ MeV} \, , \, \Gamma = 56 \pm 11 ^{+8}_{-11} \mbox{ MeV} \, ,\nonumber \\ & X(4500) : M = 4506 \pm 11 ^{+12}_{-15} \mbox{ MeV} \, , \, \Gamma = 92 \pm 21 ^{+21}_{-20} \mbox{ MeV} \, , \nonumber\\ & X(4700) : M = 4704 \pm 10 ^{+14}_{-24} \mbox{ MeV} \, , \, \Gamma = 120 \pm 31 ^{+42}_{-33} \mbox{ MeV} \, . \end{flalign} There have been several possible assignments for the two new particles $X(4500)$ and $X(4700)$. In Ref.\cite{HXChen1606}, Chen et al study the newly observed $X(4500)$ and $X(4700)$ based on the diquark-antidiquark configuration within the framework of QCD sum rules, and interpret them as the D-wave $cs\bar{c}\bar{s}$ tetraquark states with $J^P =0^+$. In Ref.\cite{Liu1607}, Liu studies the possible rescattering effects contribute to the process $B^+\to J/\psi\phi K^+$, and observes that the $D_{s}^{*+}D_{s}^-$ rescattering via the open-charmed meson loops and the $\psi^\prime \phi$ rescattering via the $\psi^\prime K_1$ loops may simulate the structures of the $X(4140)$ and $X(4700)$, respectively, and it is hard to attribute the $X(4274)$ and $X(4500)$ to the P-wave threshold rescattering effects. In Ref.\cite{Maiani1607}, Maiani, Polosa and Riquer assign the $X(4500)$ and $X(4700)$ to be the 2S tetraquark states based on the constituent diquark model. Also in Ref.\cite{Zhu1607}, Zhu assigns the $X(4500)$ and $X(4700)$ to be the 2S tetraquark states based on the constituent diquark model. In Ref.\cite{Lebed-3915}, Lebed and Polosa propose that the $X(3915)$ is the ground state $cs\bar c \bar s$ state based on lacking of the observed $D\bar D$ and $D^*\bar{D}^*$ decays, and attribute the single known decay mode $J/\psi \omega$ to the $\omega-\phi$ mixing effect. The diquarks $\varepsilon^{ijk} q^{T}_j C\Gamma q^{\prime}_k$ in color antitriplet have five structures in Dirac spinor space, where $C\Gamma=C\gamma_5$, $C$, $C\gamma_\mu \gamma_5$, $C\gamma_\mu $ and $C\sigma_{\mu\nu}$ for the scalar, pseudoscalar, vector, axialvector and tensor diquarks, respectively. In Ref.\cite{Wang1606}, we study the masses and pole residues of the $X(3915)$, $X(4500)$ and $X(4700)$ in the scenario of tetraquark states with the QCD sum rules by calculating the contributions of the vacuum condensates up to dimension 10. The theoretical calculations support assigning the $X(3915)$ and $X(4500)$ to be the 1S and 2S $C\gamma_\mu\otimes\gamma^\mu C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states, respectively, and assigning the $X(4700)$ to be the 1S $C\gamma_\mu\gamma_5\otimes\gamma_5\gamma^\mu C$ type scalar $cs\bar{c}\bar{s}$ tetraquark state. In subsequent work, we take the $X(4140)$ as the diquark-antidiquark type $cs\bar{c}\bar{s}$ tetraquark state with $J^{PC}=1^{++}$, and study the mass and pole residue with the QCD sum rules in details by constructing two types interpolating currents. The numerical results $M_{X_{L,+}}=3.95\pm0.09\,\rm{GeV}$ and $M_{X_{H,+}}=5.00\pm0.10\,\rm{GeV}$ disfavor assigning the $X(4140)$ to be the $C\gamma_5\otimes \gamma_\mu C+C\gamma_\mu\otimes \gamma_5 C$ type or $C\otimes\gamma_5 \gamma_\mu C+C\gamma_\mu\gamma_5\otimes C$ type tetraquark state with $J^{PC}=1^{++}$ \cite{Wang1607}. The attractive interactions of one-gluon exchange favor formation of the diquarks in color antitriplet $\overline{3}_{ c}$, flavor antitriplet $\overline{3}_{ f}$ and spin singlet $1_s$ or flavor sextet $6_{ f}$ and spin triplet $3_s$ \cite{One-gluon}. The calculations based on the QCD sum rules also indicate that the favored configurations are the $C\gamma_5$ and $C\gamma_\mu$ diquark states \cite{WangDiquark,WangLDiquark}, and the heavy-light $C\gamma_5$ and $C\gamma_\mu$ diquark states have almost degenerate masses \cite{WangDiquark}. In Ref.\cite{Wang-Scalar}, we construct the $C\gamma_5\otimes \gamma_5C$, $C\gamma_\mu\otimes \gamma^\mu C$, $C\gamma_\mu\gamma_5\otimes \gamma_5\gamma^\mu C$ type interpolating currents to study the scalar tetraquark states with the QCD sum rules in a systematic way, and observe that the $C\gamma_5\otimes \gamma_5C$ type and $C\gamma_\mu\otimes \gamma^\mu C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states have almost degenerate masses, about $4.44\,\rm{GeV}$. The value $4.44\,\rm{GeV}$ is not robust as the masses are extracted from the QCD spectral densities at the energy scale $\mu=1\,\rm{GeV}$. In Refs.\cite{Wang-4660-2014,WangTetraquarkCTP,Wang-Huang-NPA-2014}, we explore the energy scale dependence of the masses $M_{X/Y/Z}$ of the hidden charm (bottom) tetraquark states in details for the first time, and suggest a formula, \begin{eqnarray} \mu&=&\sqrt{M^2_{X/Y/Z}-(2{\mathbb{M}}_Q)^2} \, , \end{eqnarray} with the effective heavy quark mass ${\mathbb{M}}_Q$ to determine the energy scales of the QCD spectral densities in the QCD sum rules, which works well. Now we take a short digression to discuss the energy scale dependence of the QCD sum rules for the hidden charm or hidden bottom tetraquark states. The correlation functions $\Pi(p)$ can be written as \begin{eqnarray} \Pi(p)&=&\int_{4m^2_Q(\mu)}^{s_0} ds \frac{\rho_{QCD}(s,\mu)}{s-p^2}+\int_{s_0}^\infty ds \frac{\rho_{QCD}(s,\mu)}{s-p^2} \, , \end{eqnarray} through dispersion relation at the QCD side, where the $s_0$ are continuum threshold parameters. The $\Pi(p)$ are energy scale independent, \begin{eqnarray} \frac{d}{d\mu}\Pi(p)&=&0\, , \end{eqnarray} which does not mean \begin{eqnarray} \frac{d}{d\mu}\int_{4m^2_Q(\mu)}^{s_0} ds \frac{\rho_{QCD}(s,\mu)}{s-p^2}\rightarrow 0 \, , \end{eqnarray} due to the following two reasons inherited from the QCD sum rules:\\ $\bullet$ Perturbative corrections are neglected, the higher dimensional vacuum condensates are factorized into lower dimensional ones therefore the energy scale dependence of the higher dimensional vacuum condensates is modified;\\ $\bullet$ Truncations $s_0$ set in, the correlation between the threshold $4m^2_Q(\mu)$ and continuum threshold $s_0$ is unknown. \\ In the QCD sum rules for the hidden charm or hidden bottom tetraquark states, the integrals \begin{eqnarray} \int_{4m^2_Q(\mu)}^{s_0} ds \,\rho_{QCD}(s,\mu)\,\exp\left(-\frac{s}{T^2} \right)\, , \end{eqnarray} are sensitive to the heavy quark masses $m_Q(\mu)$ or the energy scales $\mu$, where the $T^2$ denotes the Borel parameters. Variations of the heavy quark masses or the energy scales $\mu$ lead to changes of integral ranges $4m_Q^2(\mu)-s_0$ of the variable $ds$ besides the QCD spectral densities $\rho_{QCD}(s,\mu)$, therefore changes of the Borel windows and predicted masses and pole residues. We cannot obtain energy scale independent QCD sum rules, but we have an energy scale formula to determine the energy scales consistently. According to the formula, the energy scale $\mu=1\,\rm{GeV}$ is too low to result in robust predictions \cite{Wang-Scalar}. In Ref.\cite{Wang1606}, we use the energy scale formula in Eq.(2) to determine the energy scales of the QCD spectral densities in the QCD sum rules, and observe that the $X(3915)$ and $X(4500)$ can be assigned to be the 1S and 2S $C\gamma_\mu\otimes\gamma^\mu C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states, respectively, and the $X(4700)$ can be assigned to be the 1S $C\gamma_\mu\gamma_5\otimes\gamma_5\gamma^\mu C$ type scalar $cs\bar{c}\bar{s}$ tetraquark state. In Ref.\cite{Wang1606}, we obtain the values $M_{X(3915)}=3.91^{+0.21}_{-0.17}\,\rm{GeV}$, $M_{X(4500)}=4.50^{+0.08}_{-0.09}\,\rm{GeV}$ and $M_{X(4700)}=4.70^{+0.08}_{-0.09}\,\rm{GeV}$. The value $M_{X(3915)}=3.91^{+0.21}_{-0.17}\,\rm{GeV}$ is much smaller than the value $4.44\,\rm{GeV}$ obtained in Ref.\cite{Wang-Scalar}. If the $C\gamma_5\otimes \gamma_5C$ type and $C\gamma_\mu\otimes \gamma^\mu C$ type tetraquark states have degenerate masses, then the $X(3915)$ and $X(4500)$ can have another diquark-antidiquark structure, $C\gamma_5\otimes \gamma_5C$. In this article, we study the $C\gamma_5\otimes \gamma_5C$ type and $C\otimes C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states with the QCD sum rules by calculating the contributions of the vacuum condensates up to dimension 10 in a consistent way. In calculations, we use the energy scale formula to determine the optimal energy scales of the QCD spectral densities to extract to tetraquark masses to identify the $X(3915)$, $X(4500)$ and $X(4700)$. The article is arranged as follows: we derive the QCD sum rules for the masses and pole residues of the 1S $C\gamma_5\otimes \gamma_5C$ type and $C\otimes C$ type $cs\bar{c}\bar{s}$ tetraquark states in section 2; in section 3, we derive the QCD sum rules for the masses and pole residues of the 1S and 2S $C\gamma_5\otimes \gamma_5C$ type $cs\bar{c}\bar{s}$ tetraquark states; section 4 is reserved for our conclusion. \section{QCD sum rules for the scalar tetraquark states $X_L$ and $X_H$ } In the following, we write down the two-point correlation functions $\Pi_{L/H}(p)$ in the QCD sum rules, \begin{eqnarray} \Pi_{L/H}(p)&=&i\int d^4x e^{ip \cdot x} \langle0|T\left\{J_{L/H}(x)J_{L/H}^{\dagger}(0)\right\}|0\rangle \, , \end{eqnarray} where \begin{eqnarray} J_L(x)&=&\epsilon^{ijk}\epsilon^{imn}s^j(x)C\gamma_5 c^k(x) \bar{s}^m(x)\gamma_5 C \bar{c}^n(x) \, , \nonumber\\ J_H(x)&=&\epsilon^{ijk}\epsilon^{imn}s^j(x)C c^k(x) \bar{s}^m(x) C \bar{c}^n(x)\, , \end{eqnarray} where the $i$, $j$, $k$, $m$, $n$ are color indexes, the $C$ is the charge conjugation matrix. The scalar diquark states are more stable than the pseudoscalar diquark states, so we expect that the $C\gamma_5\otimes \gamma_5C$ type tetraquark states have much smaller masses than the corresponding $C\otimes C$ type tetraquark states, and add the subscripts $L$ and $H$ to denote the light and heavy tetraquark states, respectively. At the phenomenological side, we can insert a complete set of intermediate hadronic states with the same quantum numbers as the current operators $J_{L/H}(x)$ into the correlation functions $\Pi_{L/H}(p)$ to obtain the hadronic representation \cite{SVZ79,Reinders85}. After isolating the ground state contributions of the scalar $cs\bar{c}\bar{s}$ tetraquark states $X_{L/H}$, we get the results, \begin{eqnarray} \Pi_{L}(p)&=&\frac{\lambda_{ X_L}^2}{M_{X_L}^2-p^2} +\cdots \, \, , \nonumber\\ \Pi_{H}(p)&=&\frac{\lambda_{ X_H}^2}{M_{X_H}^2-p^2} +\cdots \, \, , \end{eqnarray} where the pole residues $\lambda_{X_{L/H}}$ are defined by $\langle 0|J_{L/H}(0)|X_{L/H}(p)\rangle = \lambda_{X_{L/H}}$. In the following, we briefly outline the operator product expansion for the correlation functions $\Pi_{L/H}(p)$ in perturbative QCD. We contract the $s$ and $c$ quark fields in the correlation functions $\Pi_{L/H}(p)$ with Wick theorem, and obtain the results: \begin{eqnarray} \Pi_{L}(p)&=&i\epsilon^{ijk}\epsilon^{imn}\epsilon^{i^{\prime}j^{\prime}k^{\prime}}\epsilon^{i^{\prime}m^{\prime}n^{\prime}}\int d^4x e^{ip \cdot x} \nonumber\\ &&{\rm Tr}\left[ \gamma_{5}C^{kk^{\prime}}(x)\gamma_{5} CS^{jj^{\prime}T}(x)C\right] {\rm Tr}\left[ \gamma_{5} C^{n^{\prime}n}(-x)\gamma_{5} C S^{m^{\prime}mT}(-x)C\right] \, , \nonumber\\ \Pi_{H}(p)&=&i\epsilon^{ijk}\epsilon^{imn}\epsilon^{i^{\prime}j^{\prime}k^{\prime}}\epsilon^{i^{\prime}m^{\prime}n^{\prime}}\int d^4x e^{ip \cdot x} \nonumber\\ &&{\rm Tr}\left[ C^{kk^{\prime}}(x) CS^{jj^{\prime}T}(x)C\right] {\rm Tr}\left[ C^{n^{\prime}n}(-x) C S^{m^{\prime}mT}(-x)C\right] \, , \end{eqnarray} where the $S_{ij}(x)$ and $C_{ij}(x)$ are the full $s$ and $c$ quark propagators, respectively \cite{Reinders85,Pascual-1984}, \begin{eqnarray} S^{ij}(x)&=& \frac{i\delta_{ij}\!\not\!{x}}{ 2\pi^2x^4} -\frac{\delta_{ij}m_s}{4\pi^2x^2}-\frac{\delta_{ij}\langle \bar{s}s\rangle}{12} +\frac{i\delta_{ij}\!\not\!{x}m_s \langle\bar{s}s\rangle}{48}-\frac{\delta_{ij}x^2\langle \bar{s}g_s\sigma Gs\rangle}{192}+\frac{i\delta_{ij}x^2\!\not\!{x} m_s\langle \bar{s}g_s\sigma Gs\rangle }{1152}\nonumber\\ && -\frac{ig_s G^{a}_{\alpha\beta}t^a_{ij}(\!\not\!{x} \sigma^{\alpha\beta}+\sigma^{\alpha\beta} \!\not\!{x})}{32\pi^2x^2} -\frac{i\delta_{ij}x^2\!\not\!{x}g_s^2\langle \bar{s} s\rangle^2}{7776} -\frac{\delta_{ij}x^4\langle \bar{s}s \rangle\langle g_s^2 GG\rangle}{27648}-\frac{1}{8}\langle\bar{s}_j\sigma^{\mu\nu}s_i \rangle \sigma_{\mu\nu} \nonumber\\ && -\frac{1}{4}\langle\bar{s}_j\gamma^{\mu}s_i\rangle \gamma_{\mu }+\cdots \, , \end{eqnarray} \begin{eqnarray} C_{ij}(x)&=&\frac{i}{(2\pi)^4}\int d^4k e^{-ik \cdot x} \left\{ \frac{\delta_{ij}}{\!\not\!{k}-m_c} -\frac{g_sG^n_{\alpha\beta}t^n_{ij}}{4}\frac{\sigma^{\alpha\beta}(\!\not\!{k}+m_c)+(\!\not\!{k}+m_c) \sigma^{\alpha\beta}}{(k^2-m_c^2)^2}\right.\nonumber\\ &&\left. +\frac{g_s D_\alpha G^n_{\beta\lambda}t^n_{ij}(f^{\lambda\beta\alpha}+f^{\lambda\alpha\beta}) }{3(k^2-m_c^2)^4} -\frac{g_s^2 (t^at^b)_{ij} G^a_{\alpha\beta}G^b_{\mu\nu}(f^{\alpha\beta\mu\nu}+f^{\alpha\mu\beta\nu}+f^{\alpha\mu\nu\beta}) }{4(k^2-m_c^2)^5}+\cdots\right\} \, , \nonumber \\ \end{eqnarray} \begin{eqnarray} f^{\lambda\alpha\beta}&=&(\!\not\!{k}+m_c)\gamma^\lambda(\!\not\!{k}+m_c)\gamma^\alpha(\!\not\!{k}+m_c)\gamma^\beta(\!\not\!{k}+m_c)\, ,\nonumber\\ f^{\alpha\beta\mu\nu}&=&(\!\not\!{k}+m_c)\gamma^\alpha(\!\not\!{k}+m_c)\gamma^\beta(\!\not\!{k}+m_c)\gamma^\mu(\!\not\!{k}+m_c)\gamma^\nu(\!\not\!{k}+m_c)\, , \end{eqnarray} and $t^n=\frac{\lambda^n}{2}$, the $\lambda^n$ is the Gell-Mann matrix, $D_\alpha=\partial_\alpha-ig_sG^n_\alpha t^n$ \cite{Reinders85}. Then we compute the integrals both in coordinate space and in momentum space, and obtain the correlation functions $\Pi_{L/H}(p)$, therefore the QCD spectral densities through dispersion relation. In this article, we take into account the vacuum condensates which are vacuum expectations of the operators of the orders $\mathcal{O}( \alpha_s^{k})$ with $k\leq 1$ consistently. For the technical details, one can consult Ref.\cite{WangHuangTao}. We neglect the radiative $\mathcal{O}(\alpha_s)$ corrections for the perturbative contributions, it is a challenging or formidable work to calculate the radiative $\mathcal{O}(\alpha_s)$ corrections in the QCD sum rules for hidden charm or hidden bottom tetraquark states, though the corrections may be large in the presence of two heavy quarks, just like in the QCD sum rules for the vector and axialvector $B_c$ mesons \cite{Wangzg-Bc}. Once the analytical QCD spectral densities are obtained, we take the quark-hadron duality below the continuum thresholds $s^0_{L/H}$ and perform Borel transform with respect to the variable $P^2=-p^2$ to obtain the QCD sum rules: \begin{eqnarray} \lambda^2_{X_L}\, \exp\left(-\frac{M^2_{X_L}}{T^2}\right)= \int_{4m_c^2}^{s_L^0} ds\, \rho_L(s) \, \exp\left(-\frac{s}{T^2}\right) \, , \\ \lambda^2_{X_H}\, \exp\left(-\frac{M^2_{X_H}}{T^2}\right)= \int_{4m_c^2}^{s_H^0} ds\, \rho_H(s) \, \exp\left(-\frac{s}{T^2}\right) \, , \end{eqnarray} where \begin{eqnarray} \rho_L(s)&=&\rho_{0}(s)+\rho_{3}(s) +\rho_{4}(s)+\rho_{5}(s)+\rho_{6}(s)+\rho_{7}(s) +\rho_{8}(s)+\rho_{10}(s)\, , \nonumber\\ \rho_H(s)&=&\rho_L(s)|_{m_c\to -m_c} \, , \end{eqnarray} \begin{eqnarray} \rho_{0}(s)&=&\frac{1}{512\pi^6}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, yz\, (1-y-z)^3\left(s-\overline{m}_c^2\right)^2\left(7s^2-6s\overline{m}_c^2+\overline{m}_c^4 \right) \nonumber\\ &&+\frac{m_sm_c}{256\pi^6}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, (y+z)\, (1-y-z)^2\left(s-\overline{m}_c^2\right)^2\left(5s-2\overline{m}_c^2 \right)\, , \end{eqnarray} \begin{eqnarray} \rho_{3}(s)&=&-\frac{m_c\langle \bar{s}s\rangle}{16\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, (y+z)(1-y-z)\left(s-\overline{m}_c^2\right)\left(2s-\overline{m}_c^2\right) \nonumber\\ &&+\frac{m_s\langle \bar{s}s\rangle}{16\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, yz\, (1-y-z)\left(10s^2-12s\overline{m}_c^2+3\overline{m}_c^4 \right)\nonumber\\ &&-\frac{m_sm_c^2\langle \bar{s}s\rangle}{8\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left( s - \overline{m}_c^2\right) \, , \end{eqnarray} \begin{eqnarray} \rho_{4}(s)&=&-\frac{m_c^2}{384\pi^4} \langle\frac{\alpha_s GG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left( \frac{z}{y^2}+\frac{y}{z^2}\right)(1-y-z)^3 \nonumber\\ &&\left\{ 2s-\overline{m}_c^2+\frac{s^2}{6}\delta\left(s-\overline{m}_c^2\right)\right\} \nonumber\\ &&+\frac{1}{512\pi^4} \langle\frac{\alpha_s GG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left( y+z\right)(1-y-z)^2 \left( 10s^2-12s\overline{m}_c^2+3\overline{m}_c^4\right) \nonumber\\ &&-\frac{m_sm_c^3}{384\pi^4} \langle\frac{\alpha_s GG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left( \frac{1}{y^3}+\frac{1}{z^3}\right)\left( y+z\right)(1-y-z)^2 \left\{1+\frac{s}{2}\delta\left(s-\overline{m}_c^2\right)\right\} \nonumber\\ &&+\frac{m_sm_c}{256\pi^4} \langle\frac{\alpha_s GG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left( \frac{z}{y^2}+\frac{y}{z^2}\right)(1-y-z)^2 \left(3s-2\overline{m}_c^2\right)\nonumber\\ &&+\frac{m_sm_c}{128\pi^4} \langle\frac{\alpha_s GG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \,(1-y-z) \left(3s-2\overline{m}_c^2\right)\, , \end{eqnarray} \begin{eqnarray} \rho_{5}(s)&=&\frac{m_c\langle \bar{s}g_s\sigma Gs\rangle}{64\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, (y+z) \left(3s-2\overline{m}_c^2 \right) \nonumber\\ &&-\frac{m_c\langle \bar{s}g_s\sigma Gs\rangle}{64\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, \left( \frac{y}{z}+\frac{z}{y}\right) (1-y-z) \left(3s-2\overline{m}_c^2 \right) \nonumber\\ &&-\frac{m_s\langle \bar{s}g_s\sigma Gs\rangle}{16\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \,yz \left\{2s-\overline{m}_c^2 +\frac{s^2}{6}\delta\left(s-\overline{m}_c^2 \right)\right\} \nonumber\\ &&+\frac{m_sm_c^2\langle \bar{s}g_s\sigma Gs\rangle}{32\pi^4}\int_{y_i}^{y_f}dy \nonumber\\ &&-\frac{m_sm_c^2\langle \bar{s}g_s\sigma Gs\rangle}{64\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, \left( \frac{1}{y}+\frac{1}{z}\right) \, , \end{eqnarray} \begin{eqnarray} \rho_{6}(s)&=&\frac{m_c^2\langle\bar{s}s\rangle^2}{12\pi^2}\int_{y_i}^{y_f}dy +\frac{g_s^2\langle\bar{s}s\rangle^2}{108\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz\, yz \left\{2s-\overline{m}_c^2 +\frac{s^2}{6}\delta\left(s-\overline{m}_c^2 \right)\right\}\nonumber\\ &&-\frac{g_s^2\langle\bar{s}s\rangle^2}{512\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, (1-y-z)\left\{ 2\left(\frac{z}{y}+\frac{y}{z} \right)\left(3s-2\overline{m}_c^2 \right)+\left(\frac{z}{y^2}+\frac{y}{z^2} \right)\right.\nonumber\\ &&\left.m_c^2\left[ 2+ s\,\delta\left(s-\overline{m}_c^2 \right)\right] \right\} \nonumber\\ &&-\frac{g_s^2\langle\bar{s}s\rangle^2}{3888\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \, (1-y-z)\left\{ 3\left(\frac{z}{y}+\frac{y}{z} \right)\left(3s-2\overline{m}_c^2 \right)+\left(\frac{z}{y^2}+\frac{y}{z^2} \right)\right. \nonumber\\ &&\left.m_c^2\left[ 2+s\,\delta\left(s-\overline{m}_c^2\right)\right]+(y+z)\left[12\left(2s-\overline{m}_c^2\right) +2s^2\delta\left(s-\overline{m}_c^2\right)\right] \right\}\nonumber\\ && -\frac{m_s m_c\langle\bar{s}s\rangle^2}{12\pi^2}\int_{y_i}^{y_f}dy \left\{1 +\frac{s}{2}\delta\left(s-\widetilde{m}_c^2 \right)\right\}\nonumber\\ && +\frac{m_s m_c g_s^2\langle\bar{s}s\rangle^2}{648\pi^4}\int_{y_i}^{y_f}dy \left\{1 +\frac{s}{2}\delta\left(s-\widetilde{m}_c^2 \right)\right\}\nonumber\\ && -\frac{m_s m_c g_s^2\langle\bar{s}s\rangle^2}{216\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left\{\frac{1}{y}+\frac{1}{z}+\left( \frac{1}{y^2}+\frac{1}{z^2}\right)\frac{m_c^2}{2}\delta\left(s-\overline{m}_c^2 \right)\right\}\nonumber\\ && -\frac{m_s m_c g_s^2\langle\bar{s}s\rangle^2}{648\pi^4}\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left( \frac{y}{z}+\frac{z}{y}\right) \left\{1+\frac{s}{2}\delta\left(s-\overline{m}_c^2 \right)\right\}\, , \end{eqnarray} \begin{eqnarray} \rho_7(s)&=&\frac{m_c^3\langle\bar{s}s\rangle}{288\pi^2 }\langle\frac{\alpha_sGG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left(\frac{1}{y^3}+\frac{1}{z^3} \right)(y+z)(1-y-z)\left(1+\frac{ s}{T^2}\right) \delta\left(s-\overline{m}_c^2\right)\nonumber\\ &&-\frac{m_c\langle\bar{s}s\rangle}{96\pi^2}\langle\frac{\alpha_sGG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left(\frac{y}{z^2}+\frac{z}{y^2}\right)(1-y-z) \left\{2+s\,\delta\left(s-\overline{m}_c^2\right) \right\}\nonumber\\ &&-\frac{m_c\langle\bar{s}s\rangle}{96\pi^2}\langle\frac{\alpha_sGG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz\left\{2+ s\, \delta\left(s-\overline{m}_c^2\right) \right\} \nonumber\\ &&-\frac{m_c\langle\bar{s}s\rangle}{576\pi^2}\langle\frac{\alpha_sGG}{\pi}\rangle\int_{y_i}^{y_f}dy \left\{2+ s \, \delta \left(s-\widetilde{m}_c^2\right) \right\} \nonumber\\ &&-\frac{m_s m_c^2\langle\bar{s}s\rangle}{144\pi^2 }\langle\frac{\alpha_sGG}{\pi}\rangle\int_{0}^{1}dy \int_{0}^{1-y}dz \left(\frac{z}{y^2}+\frac{y}{z^2} \right)(1-y-z)\left(1+\frac{ s}{T^2}+\frac{ 2s^2}{T^4}\right) \delta\left(s-\overline{m}_c^2\right)\nonumber\\ &&+\frac{m_s m_c^4\langle\bar{s}s\rangle}{144\pi^2T^2 }\langle\frac{\alpha_sGG}{\pi}\rangle\int_{0}^{1}dy \int_{0}^{1-y}dz \left(\frac{1}{y^3}+\frac{1}{z^3} \right) \delta\left(s-\overline{m}_c^2\right)\nonumber\\ &&-\frac{m_s m_c^2\langle\bar{s}s\rangle}{48\pi^2 }\langle\frac{\alpha_sGG}{\pi}\rangle\int_{0}^{1}dy \int_{0}^{1-y}dz \left(\frac{1}{y^2}+\frac{1}{z^2} \right) \delta\left(s-\overline{m}_c^2\right)\nonumber\\ &&+\frac{m_s \langle\bar{s}s\rangle}{64\pi^2 }\langle\frac{\alpha_sGG}{\pi}\rangle\int_{y_i}^{y_f}dy \int_{z_i}^{1-y}dz \left(y+z \right) \left\{ 1+\left( \frac{2s}{3}+\frac{s^2}{6T^2}\right)\delta\left(s-\overline{m}_c^2\right)\right\}\nonumber\\ &&-\frac{m_s m_c^2\langle\bar{s}s\rangle}{288\pi^2 }\langle\frac{\alpha_sGG}{\pi}\rangle\int_{0}^{1}dy \left( 1+\frac{s}{T^2}\right)\delta\left(s-\widetilde{m}_c^2\right) \, , \end{eqnarray} \begin{eqnarray} \rho_8(s)&=&-\frac{m_c^2\langle\bar{s}s\rangle\langle\bar{s}g_s\sigma Gs\rangle}{24\pi^2}\int_0^1 dy \left(1+\frac{s}{T^2} \right)\delta\left(s-\widetilde{m}_c^2\right)\nonumber \\ &&+\frac{ \langle\bar{s}s\rangle\langle\bar{s}g_s\sigma Gs\rangle}{48\pi^2}\int_{0}^{1} dy \,s\,\delta\left(s-\widetilde{m}_c^2\right) \nonumber\\ &&+\frac{ 5m_s m_c\langle\bar{s}s\rangle\langle\bar{s}g_s\sigma Gs\rangle}{144\pi^2}\int_{0}^{1} dy\left(1+\frac{s}{T^2}+\frac{s^2}{2T^4} \right)\delta\left(s-\widetilde{m}_c^2\right) \nonumber\\ &&-\frac{ m_s m_c\langle\bar{s}s\rangle\langle\bar{s}g_s\sigma Gs\rangle}{96\pi^2}\int_{0}^{1} dy \left( \frac{1-y}{y}+\frac{y}{1-y}\right)\left(1+\frac{s}{T^2} \right)\delta\left(s-\widetilde{m}_c^2\right)\, , \end{eqnarray} \begin{eqnarray} \rho_{10}(s)&=&\frac{m_c^2\langle\bar{s}g_s\sigma Gs\rangle^2}{192\pi^2T^6}\int_0^1 dy \, s^2 \, \delta \left( s-\widetilde{m}_c^2\right) \nonumber \\ &&-\frac{m_c^4\langle\bar{s}s\rangle^2}{216T^4}\langle\frac{\alpha_sGG}{\pi}\rangle\int_0^1 dy \left\{ \frac{1}{y^3}+\frac{1}{(1-y)^3}\right\} \delta\left( s-\widetilde{m}_c^2\right)\nonumber\\ &&+\frac{m_c^2\langle\bar{s}s\rangle^2}{72T^2}\langle\frac{\alpha_sGG}{\pi}\rangle\int_0^1 dy \left\{ \frac{1}{y^2}+\frac{1}{(1-y)^2}\right\} \delta\left( s-\widetilde{m}_c^2\right)\nonumber\\ &&-\frac{\langle\bar{s}g_s\sigma Gs\rangle^2}{192 \pi^2T^4} \int_0^1 dy \,s^2 \, \delta\left( s-\widetilde{m}_c^2\right)\nonumber\\ &&+\frac{ \langle\bar{s}g_s\sigma Gs\rangle^2}{128 \pi^2T^2} \int_0^1 dy \,s\, \delta\left( s-\widetilde{m}_c^2\right)\nonumber \\ &&+\frac{m_c^2\langle\bar{s} s\rangle^2}{216 T^6}\langle\frac{\alpha_sGG}{\pi}\rangle\int_0^1 dy \, s^2 \, \delta \left( s-\widetilde{m}_c^2\right) \nonumber\\ &&-\frac{m_sm_c \langle\bar{s}g_s\sigma Gs\rangle^2}{576\pi^2T^8}\int_0^1 dy \, s^3 \, \delta \left( s-\widetilde{m}_c^2\right) \nonumber \\ &&-\frac{m_s m_c^3\langle\bar{s}s\rangle^2}{432T^4}\langle\frac{\alpha_sGG}{\pi}\rangle\int_0^1 dy \left\{ \frac{1}{y^3}+\frac{1}{(1-y)^3}\right\}\left(1-\frac{s}{T^2} \right) \delta\left( s-\widetilde{m}_c^2\right)\nonumber\\ &&-\frac{m_s m_c\langle\bar{s}s\rangle^2}{144T^4}\langle\frac{\alpha_sGG}{\pi}\rangle\int_0^1 dy \left\{ \frac{1-y}{y^2}+\frac{y}{(1-y)^2}\right\}\,s\, \delta\left( s-\widetilde{m}_c^2\right)\nonumber\\ &&+\frac{m_sm_c \langle\bar{s}g_s\sigma Gs\rangle^2}{576\pi^2T^6}\int_0^1 dy \left( \frac{1-y}{y}+\frac{y}{1-y}\right)\, s^2 \, \delta \left( s-\widetilde{m}_c^2\right) \nonumber \\ &&-\frac{m_s m_c\langle\bar{s}s\rangle^2}{864T^8}\langle\frac{\alpha_sGG}{\pi}\rangle\int_0^1 dy \,s^3\, \delta\left( s-\widetilde{m}_c^2\right) \, , \end{eqnarray} the subscripts $0$, $3$, $4$, $5$, $6$, $7$, $8$, $10$ denote the dimensions of the vacuum condensates, $y_{f}=\frac{1+\sqrt{1-4m_c^2/s}}{2}$, $y_{i}=\frac{1-\sqrt{1-4m_c^2/s}}{2}$, $z_{i}=\frac{y m_c^2}{y s -m_c^2}$, $\overline{m}_c^2=\frac{(y+z)m_c^2}{yz}$, $ \widetilde{m}_c^2=\frac{m_c^2}{y(1-y)}$, $\int_{y_i}^{y_f}dy \to \int_{0}^{1}dy$, $\int_{z_i}^{1-y}dz \to \int_{0}^{1-y}dz$, when the $\delta$ functions $\delta\left(s-\overline{m}_c^2\right)$ and $\delta\left(s-\widetilde{m}_c^2\right)$ appear. We derive Eqs.(14-15) with respect to $\frac{1}{T^2}$, then eliminate the pole residues $\lambda_{X_{L/H}}$, and obtain the QCD sum rules for the masses $ M_{X_{L/H}}$ of the scalar tetraquark states, \begin{eqnarray} M^2_{X_{L/H}}&=& -\frac{\int_{4m_c^2}^{s_0} ds\,\frac{d}{d \left(1/T^2\right)}\,\rho_{L/H}(s)\,\exp\left(-\frac{s}{T^2}\right)}{\int_{4m_c^2}^{s_0} ds \, \rho_{L/H}(s)\, \exp\left(-\frac{s}{T^2}\right)}\, . \end{eqnarray} The vacuum condensates are taken to be the standard values $\langle\bar{q}q \rangle=-(0.24\pm 0.01\, \rm{GeV})^3$, $\langle\bar{s}s \rangle=(0.8\pm0.1)\langle\bar{q}q \rangle$, $\langle\bar{s}g_s\sigma G s \rangle=m_0^2\langle \bar{s}s \rangle$, $m_0^2=(0.8 \pm 0.1)\,\rm{GeV}^2$, $\langle \frac{\alpha_s GG}{\pi}\rangle=(0.33\,\rm{GeV})^4 $ at the energy scale $\mu=1\, \rm{GeV}$ \cite{SVZ79,Reinders85,Colangelo-00}. The quark condensates and mixed quark condensates evolve with the renormalization group equation, $\langle\bar{s}s \rangle(\mu)=\langle\bar{s}s \rangle(Q)\left[\frac{\alpha_{s}(Q)}{\alpha_{s}(\mu)}\right]^{\frac{4}{9}}$, and $\langle\bar{s}g_s \sigma Gs \rangle(\mu)=\langle\bar{s}g_s \sigma Gs \rangle(Q)\left[\frac{\alpha_{s}(Q)}{\alpha_{s}(\mu)}\right]^{\frac{2}{27}}$ \cite{Narison-book}. We take the $\overline{MS}$ masses $m_{c}(m_c)=(1.275\pm0.025)\,\rm{GeV}$ and $m_s(\mu=2\,\rm{GeV})=(0.095\pm0.005)\,\rm{GeV}$ from the Particle Data Group \cite{PDG}, and take into account the energy-scale dependence of the $\overline{MS}$ masses from the renormalization group equation, \begin{eqnarray} m_c(\mu)&=&m_c(m_c)\left[\frac{\alpha_{s}(\mu)}{\alpha_{s}(m_c)}\right]^{\frac{12}{25}} \, ,\nonumber\\ m_s(\mu)&=&m_s({\rm 2GeV} )\left[\frac{\alpha_{s}(\mu)}{\alpha_{s}({\rm 2GeV})}\right]^{\frac{4}{9}} \, ,\nonumber\\ \alpha_s(\mu)&=&\frac{1}{b_0t}\left[1-\frac{b_1}{b_0^2}\frac{\log t}{t} +\frac{b_1^2(\log^2{t}-\log{t}-1)+b_0b_2}{b_0^4t^2}\right]\, , \end{eqnarray} where $t=\log \frac{\mu^2}{\Lambda^2}$, $b_0=\frac{33-2n_f}{12\pi}$, $b_1=\frac{153-19n_f}{24\pi^2}$, $b_2=\frac{2857-\frac{5033}{9}n_f+\frac{325}{27}n_f^2}{128\pi^3}$, $\Lambda=213\,\rm{MeV}$, $296\,\rm{MeV}$ and $339\,\rm{MeV}$ for the flavors $n_f=5$, $4$ and $3$, respectively \cite{PDG}. In the four-quark system $q\bar{q}^{\prime}Q\bar{Q}$, the $Q$-quark serves as a static well potential and attracts the light quark $q$ to form a heavy diquark $\mathcal{D}$ in color antitriplet, while the $\bar{Q}$-quark serves as another static well potential and attracts the light antiquark $\bar{q}^\prime$ to form a heavy antidiquark $\mathcal{\bar{D}}$ in color triplet \cite{Wang-4660-2014,WangTetraquarkCTP,Wang-Huang-NPA-2014,WangHuangTao}. Then the $\mathcal{D}$ and $\mathcal{\bar{D}}$ attract each other to form a compact tetraquark state \cite{Wang-4660-2014,WangTetraquarkCTP,Wang-Huang-NPA-2014,WangHuangTao}, the two heavy quarks $Q$ and $\bar{Q}$ stabilize the tetraquark state \cite{Brodsky-2014}. The tetraquark states $\mathcal{D\bar{D}}$ are characterized by the effective heavy quark masses ${\mathbb{M}}_Q$ and the virtuality $V=\sqrt{M^2_{X/Y/Z}-(2{\mathbb{M}}_Q)^2}$. It is natural to take the energy scale $\mu=V$. We cannot obtain energy scale independent QCD sum rules, but we have an energy scale formula to determine the energy scales consistently. We fit the effective $Q$-quark mass ${\mathbb{M}}_{Q}$ to reproduce the experimental value $M_{Z_c(3900)/Z_b(10610)}$, the empirical effective $Q$-quark mass ${\mathbb{M}}_Q$ is universal in the QCD sum rules for the hidden-charm or hidden bottom tetraquark states and embodies the net effects of the complex dynamics. We take the empirical energy scale formula and reproduce the experimental values of the masses of the $X(3872)$, $Z_c(3900)$, $Z_c(4020)$, $Z_c(4025)$, $Z(4430)$, $Y(4660)$, $Z_b(10610)$ and $Z_b(10650)$ in the scenario of tetraquark states \cite{Wang-4660-2014,WangTetraquarkCTP,Wang-Huang-NPA-2014,WangHuangTao,WangEPJC1601,Wang4430}. In this article, we take the updated value ${\mathbb{M}}_c=1.82\,\rm{GeV}$ \cite{WangEPJC1601}. We search for the Borel parameters $T^2$ and continuum threshold parameters $s^0_{L/H}$ according to the four criteria: $\bf{1_\cdot}$ Pole dominance at the phenomenological side; $\bf{2_\cdot}$ Convergence of the operator product expansion; $\bf{3_\cdot}$ Appearance of the Borel platforms; $\bf{4_\cdot}$ Satisfying the energy scale formula. Now we take a short digression to discuss how to choose the Borel parameters. At the phenomenological side of the QCD sum rules, we prefer smaller Borel parameters so as to depress the contributions of the higher excited states and continuum states and determine the upper bound of the Borel parameters $T^2_{max}$. At the QCD side of the QCD sum rules, we prefer larger Borel parameters so as to warrant the convergence of the operator product expansion and determine the lower bound of the Borel parameters $T^2_{min}$. In the QCD sum rules for the tetraquark states, the operator product expansion converges slowly, the $T^2_{min}$ is postponed to large value, the Borel window $T^2_{max}-T^2_{min}$ is rather small. However, the small Borel window does exist, the Borel parameter is just a free parameter, the predicted masses and pole residues should be independent on this parameter, in other words, there appears Borel platform. The resulting Borel parameters $T^2$ and threshold parameters $s^0_{L/H}$ are \begin{eqnarray} X_L &:& T^2 = (2.7-3.1) \mbox{ GeV}^2 \, ,\, s^0_{L} = (4.4\pm0.1 \mbox{ GeV})^2 \, , \\ X_H &:& T^2 = (5.2-5.6) \mbox{ GeV}^2 \, ,\, s^0_{H} = (6.0\pm0.1 \mbox{ GeV})^2 \, . \end{eqnarray} The pole contributions are \begin{eqnarray} X_L &:& {\rm pole} = (39-62) \% \, \,\, {\rm at}\, \, \, \mu = 1.40 \mbox{ GeV} \, , \\ X_H &:& {\rm pole} = (43-58) \% \, \,\, {\rm at}\, \, \, \mu = 4.10 \mbox{ GeV} \, , \end{eqnarray} the pole dominance condition is satisfied. The contributions come from the vacuum condensates of dimension $i$ $D_{i}$ are \begin{eqnarray} X_L &:& D_{0}=(37-39)\%\, , \,\, D_{3}=(65-67)\%\, , \,\, D_{4}=1\%\, , \,\,D_{5}=-(14-17)\%\, , \,\,D_{6}=(9-11)\%\, ,\nonumber\\ &&D_{7}=2\%\, , \,\, D_{8}=-(1-2)\%\, , \,\, D_{10} \ll 1 \% \, \,\, {\rm at}\, \,\, \mu = 1.40 \mbox{ GeV} \, , \\ X_H &:& D_{0}=(165-174)\%\, , \,\, D_{3}=-(69-78)\%\, , \,\, D_{4}=1\%\, , \,\,D_{5}=1\%\, , \,\,D_{6}=(2-3)\%\, ,\nonumber\\ &&D_{7}=-1\%\, , \,\, D_{8}\ll 1\%\, , \,\, D_{10} \ll 1 \% \, \,\, {\rm at}\, \,\, \mu = 4.10 \mbox{ GeV} \, , \end{eqnarray} where $i=0,\,3,\,4,\,5,\,6,\,7,\,8,\,10$, the operator product expansion is well convergent. Now we take into account uncertainties of all the input parameters, and obtain the values of the masses and pole residues of the $X_L$ and $X_H$, \begin{eqnarray} M_{X_L}&=&3.89\pm 0.05\,\rm{GeV} \, , \,\,\, {\rm Experimental\,\, value} \,\,\,\,3918.4\pm 1.9\,\rm{ MeV} \, \cite{PDG}\, , \\ M_{X_H}&=&5.48\pm0.10\,\rm{GeV} \, , \,\,\, {\rm Experimental\,\, value} \,\,\,\,4704 \pm 10 ^{+14}_{-24} \mbox{ GeV}\, \cite{LHCb-4500-1606}\, , \\ \lambda_{X_L}&=&(2.21 \pm 0.22)\times 10^{-2}\,\rm{GeV}^5 \, , \nonumber\\ \lambda_{X_H}&=&(1.98\pm0.08)\times 10^{-1}\,\rm{GeV}^5 \, , \end{eqnarray} which are also shown in Figs.1-2. In Figs.1-2, we plot the masses and pole residues with variations of the Borel parameters at much larger intervals than the Borel windows shown in Eqs.(27-28). From Eqs.(33-34) and Figs.1-2, we can see that the energy scale formula is satisfied and there appear platforms in the Borel windows, the uncertainties originate from the Borel parameters in the Borel windows are very small, $\delta M_{X_L}/M_{X_L},\,\delta M_{X_H}/M_{X_H}\ll 1\%$, while the uncertainties originate from the Borel parameters outside of the Borel windows are rather large compared to the central values, where the criterion $\bf{1}$ or the criterion $\bf{2}$ is not satisfied. In the Borel windows, the four criteria are all satisfied, we expect to make reliable predictions. In calculations, we observe that the predicted masses $M_{X_L}$ and $M_{X_H}$ decrease monotonously with increase of the energy scales, the uncertainty $\delta \mu=\pm0.1\,\rm{GeV}$ can lead to uncertainties $\delta M_{X_L}=\pm20\,\rm{MeV}$ and $\delta M_{X_H}=\pm3\,\rm{MeV}$. The present prediction $M_{X_L}=3.89\pm 0.05\,\rm{GeV}$ is compatible with the experimental value $M_{X(3915)}=3918.4\pm 1.9\,\rm{ MeV}$ \cite{PDG}, so it is reasonable to assign the $X_L$ to be the $X(3915)$. The predicted mass $M_{X_H}=5.48\pm0.10\,\rm{GeV}$ lies above the upper bound of the experimental value $M_{X(4700)} = 4704 \pm 10 ^{+14}_{-24} \mbox{ GeV}$, it is impossible to assign the $X(4700)$ to be the $C-C$ type scalar tetraquark state. In Ref.\cite{Wang1606}, we obtain the values $M_{X(3915)}=3.91^{+0.21}_{-0.17}\,\rm{GeV}$, $M_{X(4500)}=4.50^{+0.08}_{-0.09}\,\rm{GeV}$, $M_{X(4700)}=4.70^{+0.08}_{-0.09}\,\rm{GeV}$, which are in excellent agreement with the experimental data, and support assigning the $X(3915)$ and $X(4500)$ to be the 1S and 2S $C\gamma_\mu\otimes\gamma^\mu C$ type tetraquark states, and support assigning the $X(4700)$ to be the 1S $C\gamma_\mu\gamma_5\otimes\gamma_5\gamma^\mu C$ type tetraquark state. For the hidden-charm mesons, the energy gaps between the ground states and the first radial excited states are $M_{\psi^\prime}-M_{J/\psi}=589\,\rm{MeV}$, $M_{\eta_c^\prime}-M_{\eta_c}=656\,\rm{MeV}$ \cite{PDG}, $M_{Z(4430)}-M_{Z_c(3900)}=576\,\rm{MeV}$ \cite{Wang4430}. In this article, we choose the continuum threshold parameters as $\sqrt{s^0_{L}}-M_{X_L} = 0.4\sim 0.6 \mbox{ GeV} $ and $\sqrt{s^0_{H}}-M_{X_H} = 0.4\sim 0.6 \mbox{ GeV} $, the contaminations of the radial excited states can be neglected. On the other hand, the currents $J_{L/H}(x)$ couple potentially to the scattering states $ J/\psi\omega$, $J/\psi \phi$, $D_s^{*\pm} D_s^{*\mp}$, $\cdots$, we can take into account the contributions of the intermediate meson-loops. All the renormalized self-energies contribute a finite imaginary part to modify the dispersion relation, we can take into account the finite width effects by the simple replacement of the hadronic spectral densities, \begin{eqnarray} \delta \left(s-M^2_{L/H} \right) &\to& \frac{1}{\pi}\frac{\sqrt{s}\,\Gamma_{L/H}(s)}{\left(s-M_{L/H}^2\right)^2+s\,\Gamma_{L/H}^2(s)}\, . \end{eqnarray} The widths $\Gamma_{X(3915)} = 20 \pm 5 \mbox{ MeV}$ \cite{PDG} and $\Gamma_{X(4700)} = 120 \pm 31 {}^{+42}_{-33} \mbox{ MeV}$ \cite{LHCb-4500-1606} are not broad, the effects of the finite widths can be absorbed safely into the pole residues \cite{Wang1606,Wang1607,WangIJTP}, the present predictions of the masses are reasonable. Now we can obtain the conclusion tentatively that the $X(3915)$ maybe have both $C\gamma_\mu\otimes\gamma^\mu C$ type and $C\gamma_5\otimes\gamma_5 C$ type tetraquark components. \begin{figure} \centering \includegraphics[totalheight=6cm,width=7cm]{mass-X-L.EPS} \includegraphics[totalheight=6cm,width=7cm]{mass-X-H.EPS} \caption{ The masses $M_{X_L}$ and $M_{X_H}$ with variations of the Borel parameters $T^2$, where the experimental value denotes the experimental value of the mass $M_{X(3915)}$. } \end{figure} \begin{figure} \centering \includegraphics[totalheight=6cm,width=7cm]{residue-X-L.EPS} \includegraphics[totalheight=6cm,width=7cm]{residue-X-H.EPS} \caption{ The pole residues $\lambda_{X_L}$ and $\lambda_{X_H}$ with variations of the Borel parameters $T^2$. } \end{figure} In the following, we perform Fierz re-arrangement to the currents $J_L$ and $J_H$ both in the color and Dirac-spinor spaces to obtain the results, \begin{eqnarray} J_L&=&\frac{1}{4}\left\{\,-\bar{c} c\,\bar{s} s+\bar{c}i\gamma_5 c\,\bar{s}i\gamma_5 s-\bar{c} \gamma^\mu c\,\bar{s}\gamma_\mu s-\bar{c} \gamma^\mu\gamma_5 c\,\bar{s}\gamma_\mu\gamma_5 s+\frac{1}{2}\bar{c}\sigma_{\mu\nu} c\,\bar{s}\sigma^{\mu\nu} s\right. \nonumber\\ &&\left.+\bar{c} s\,\bar{s} c-\bar{c}i\gamma_5 s\,\bar{s}i\gamma_5 c+\bar{c} \gamma^\mu s\,\bar{s}\gamma_\mu c+\bar{c} \gamma^\mu\gamma_5 s\,\bar{s}\gamma_\mu\gamma_5 c-\frac{1}{2}\bar{c}\sigma_{\mu\nu} s\,\bar{s}\sigma^{\mu\nu} c \,\right\} \, , \end{eqnarray} \begin{eqnarray} J_H&=&\frac{1}{4}\left\{\,\bar{c} i\gamma_5c\,\bar{s} i\gamma_5 s-\bar{c}c\,\bar{s} s+\bar{c} \gamma^\mu \gamma_5c\,\bar{s}\gamma_\mu \gamma_5s +\bar{c} \gamma^\mu c\,\bar{s} \gamma_\mu s+\frac{1}{2}\bar{c}\sigma_{\mu\nu} \gamma_5c\,\bar{s}\sigma^{\mu\nu}\gamma_5 s\right. \nonumber\\ &&\left.+\bar{c} s\,\bar{s} c-\bar{c}i\gamma_5 s\,\bar{s}i \gamma_5c-\bar{c} \gamma^\mu s\,\bar{s}\gamma_\mu c-\bar{c} \gamma^\mu\gamma_5 s\,\bar{s}\gamma_\mu \gamma_5c-\frac{1}{2}\bar{c}\sigma_{\mu\nu} s\,\bar{s}\sigma^{\mu\nu} c \,\right\} \, , \end{eqnarray} the components couple to the meson pairs, for example, $J/\psi \phi$, $D_s\bar{D}_s$, $D_s^*\bar{D}_s^*$. The two-body strong decays \begin{eqnarray} X_L &\to& J/\psi \phi \to J/\psi\omega \,\,\, (\phi-\omega\,\,{\rm mixing})\, , \nonumber\\ X_H&\to& J/\psi\phi \,\, , \, \, D_{s}\bar{D}_{s}\,\, , \, \, D^*_{s}\bar{D}^*_{s}\,\, , \, \, D_{s0}\bar{D}_{s0}\,\, , \, \, D_{s1}\bar{D}_{s1} \end{eqnarray} are Okubo-Zweig-Iizuka super-allowed. We can search for the $X_L$ and $X_H$ in those decays in the future. The diquark-antidiquark type tetraquark state can be taken as a special superposition of a series of meson-meson pairs, and embodies the net effects. The decays to its components, for example, $J/\psi \phi$, are Okubo-Zweig-Iizuka super-allowed, but the re-arrangements in the color-space are non-trivial. \section{QCD sum rules for the $X(3915)$ and $X(4500)$ as the $C\gamma_5 \otimes \gamma_5C$ type tetraquark states } Now we tentatively assign the $X(3915)$ and $X(4500)$ to be the 1S and 2S $C\gamma_5 \otimes \gamma_5C$ type tetraquark states, respectively, and study their masses and pole residues with the QCD sum rules. At the phenomenological side, we isolate the 1S and 2S scalar $cs\bar{c}\bar{s}$ tetraquark states in the correlation function $\Pi_{L}(p)$, and get the following result, \begin{eqnarray} \Pi_{L}(p)&=&\frac{\lambda_{ X_L{(\rm 1S)}}^2}{M_{X_L{(\rm 1S)}}^2-p^2} +\frac{\lambda_{ X_L{(\rm 2S)}}^2}{M_{X_L{(\rm 2S)}}^2-p^2} +\cdots \, \, , \end{eqnarray} where the pole residues $\lambda_{X_L{(\rm 1S,2S)}}$ are defined by $\langle 0|J_{L}(0)|X_L{(\rm 1S,2S)}(p)\rangle = \lambda_{X_L{(\rm 1S,2S)}}$. As the current $J_{L}(x)$ couples potentially to the scattering states $ J/\psi\omega$, $J/\psi \phi$, $D_s^{*\pm} D_s^{*\mp}$, $\cdots$, which contribute a finite imaginary part to modify the dispersion relation, we can take into account the finite width effects according to Eq.(36), and absorb the finite width effects into the pole residues \cite{Wang1606,Wang1607,WangIJTP}. The contributions of the scattering states $ J/\psi\omega$, $J/\psi \phi$, $D_s^{*\pm} D_s^{*\mp}$, $\cdots$ can be safely neglected if only the predicted masses $M_{X_L{(\rm 1S,2S)}}$ are concerned. We take into account the contribution of the $X(4500)$ and postpone the continuum threshold $s_L^0$ to the large value $s_{X(4500)}^0$ in the QCD sum rule in Eq.(14) to obtain the QCD sum rule, \begin{eqnarray} \lambda^2_{X_L{(\rm 1S)}}\, \exp\left(-\frac{M^2_{X_L{(\rm 1S)}}}{T^2}\right)+\lambda^2_{X_L{(\rm 2S)}}\, \exp\left(-\frac{M^2_{X_L{(\rm 2S)}}}{T^2}\right)= \int_{4m_c^2}^{s_{X(4500)}^0} ds\, \rho_L(s) \, \exp\left(-\frac{s}{T^2}\right) \, , \end{eqnarray} then we introduce the notations $\tau=\frac{1}{T^2}$, $D^n=\left( -\frac{d}{d\tau}\right)^n$, and use the subscripts $1$ and $2$ to denote the 1S state $X(3915)$ and the 2S state $X(4500)$ respectively for simplicity. The QCD sum rule can be rewritten as \begin{eqnarray} \lambda_1^2\exp\left(-\tau M_1^2 \right)+\lambda_2^2\exp\left(-\tau M_2^2 \right)&=&\Pi_{QCD}(\tau) \, , \end{eqnarray} here we add the subscript $QCD$ to denote the QCD side of the correlation function $\Pi_L(\tau)$. We derive the QCD sum rule in Eq.(42) with respect to $\tau$ to obtain \begin{eqnarray} \lambda_1^2M_1^2\exp\left(-\tau M_1^2 \right)+\lambda_2^2M_2^2\exp\left(-\tau M_2^2 \right)&=&D\Pi_{QCD}(\tau) \, . \end{eqnarray} Then we have two equations, and obtain the QCD sum rules, \begin{eqnarray} \lambda_i^2\exp\left(-\tau M_i^2 \right)&=&\frac{\left(D-M_j^2\right)\Pi_{QCD}(\tau)}{M_i^2-M_j^2} \, , \end{eqnarray} where $i \neq j$. Now we derive the QCD sum rules in Eq.(44) with respect to $\tau$ to obtain \begin{eqnarray} M_i^2&=&\frac{\left(D^2-M_j^2D\right)\Pi_{QCD}(\tau)}{\left(D-M_j^2\right)\Pi_{QCD}(\tau)} \, , \nonumber\\ M_i^4&=&\frac{\left(D^3-M_j^2D^2\right)\Pi_{QCD}(\tau)}{\left(D-M_j^2\right)\Pi_{QCD}(\tau)}\, . \end{eqnarray} The squared masses $M_i^2$ satisfy the following equation, \begin{eqnarray} M_i^4-b M_i^2+c&=&0\, , \end{eqnarray} where \begin{eqnarray} b&=&\frac{D^3\otimes D^0-D^2\otimes D}{D^2\otimes D^0-D\otimes D}\, , \nonumber\\ c&=&\frac{D^3\otimes D-D^2\otimes D^2}{D^2\otimes D^0-D\otimes D}\, , \nonumber\\ D^j \otimes D^k&=&D^j\Pi_{QCD}(\tau) \, D^k\Pi_{QCD}(\tau)\, , \end{eqnarray} $i=1,2$, $j,k=0,1,2,3$. We solve the equation in Eq.(46) and obtain the solutions \begin{eqnarray} M_1^2=\frac{b-\sqrt{b^2-4c} }{2} \, , \\ M_2^2=\frac{b+\sqrt{b^2-4c} }{2} \, . \end{eqnarray} In Ref.\cite{Baxi-G}, M. S. Maior de Sousa and R. Rodrigues da Silva study the masses and decay constants of the $\rho({\rm 1S,2S})$, $\psi({\rm 1S,2S})$, $\Upsilon({\rm 1S,2S})$ using Eqs.(48-49), and observe that the ground state masses are (much) smaller than the experimental values. In Ref.\cite{Wang4430}, we apply this approach to study the $Z_c(3900)$ and $Z(4430)$ as the 1S and 2S axialvector hidden-charm tetraquark states, respectively, and use the energy scale formula $\mu=\sqrt{M^2_{X/Y/Z}-(2{\mathbb{M}}_c)^2}$ to overcome the shortcoming \cite{Baxi-G}, and reproduce the experimental values of the masses $M_{Z_c(3900)}$ and $M_{Z(4430)}$. In Ref.\cite{Wang1606}, we use the same approach to study the $X(3915)$ and $X(4500)$ as the 1S and 2S $C\gamma_\mu\otimes\gamma^\mu C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states, respectively, and reproduce the experimental values of the masses $M_{X(3915)}$ and $M_{X(4500)}$. Now we tentatively assign the $X(3915)$ and $X(4500)$ to be the 1S and 2S $C\gamma_5 \otimes \gamma_5C$ type tetraquark states, respectively, and resort to the same approach study the masses and pole residues with the QCD sum rules. Again, we search for the Borel parameter $T^2$ and continuum threshold parameter $s^0_{X(4500)}$ according to the four criteria: $\bf{1_\cdot}$ Pole dominance at the phenomenological side; $\bf{2_\cdot}$ Convergence of the operator product expansion; $\bf{3_\cdot}$ Appearance of the Borel platforms; $\bf{4_\cdot}$ Satisfying the energy scale formula.\\ The resulting Borel parameter and threshold parameter are \begin{eqnarray} X(3890/4350) &:& T^2 = (1.8-2.2) \mbox{ GeV}^2 \, ,\, s^0_{X(4350)} = (4.9\pm0.1 \mbox{ GeV})^2 \, , \end{eqnarray} here we use the notations $3890$ and $4350$ in stead of $3915$ and $4500$ according to the masses extracted from the QCD sum rules. The pole contributions are \begin{eqnarray} X(3890)+X(4350) &:& {\rm pole} = (87-98) \% \, \,\, {\rm at}\, \,\, \mu = 1.25 \mbox{ GeV} \, , \\ X(3890)+X(4350) &:& {\rm pole} = (93-99) \% \, \,\, {\rm at}\, \,\, \mu = 2.40 \mbox{ GeV} \, , \end{eqnarray} the pole dominance condition is well satisfied. The contributions come from the vacuum condensates of dimension 10 $D_{10}$ are \begin{eqnarray} X(3890)+X(4350) &:& D_{10} = (1-3) \% \, \,\, {\rm at}\, \,\, \mu = 1.25 \mbox{ GeV} \, , \\ X(3890)+X(4350) &:& D_{10} = (0-1) \% \, \,\, {\rm at}\, \,\, \mu = 2.40 \mbox{ GeV} \, , \end{eqnarray} the operator product expansion is well convergent. Now we take into account uncertainties of all the input parameters, and obtain the masses and pole residues of the $X(3890)$ and $X(4350)$, \begin{eqnarray} M_{X(3890)}&=&3.85^{+0.18}_{-0.17}\,\rm{GeV} \, , \,\,\, {\rm Experimental\,\, value} \,\,\,\,3918.4\pm 1.9\,\rm{ MeV} \, \cite{PDG}\, , \nonumber\\ M_{X(4350)}&=&4.83 \,\rm{GeV} \, , \nonumber\\ \lambda_{X(3890)}&=&1.84^{+0.76}_{-0.54}\times 10^{-2}\,\rm{GeV}^5 \, , \nonumber\\ \lambda_{X(4350)}&=&5.82\times 10^{-2}\,\rm{GeV}^5 \, , \end{eqnarray} at the energy scale $\mu=1.25\,\rm{GeV}$, \begin{eqnarray} M_{X(3890)}&=&3.29\,\rm{GeV} \, , \nonumber\\ M_{X(4350)}&=&4.35^{+0.10}_{-0.11}\,\rm{GeV} \, , \,\,\, {\rm Experimental\,\, value} \,\,\,\,4506 \pm 11 ^{+12}_{-15}\, \rm {MeV}\, \cite{LHCb-4500-1606}\, , \nonumber\\ \lambda_{X(3890)}&=&1.16\times 10^{-2}\,\rm{GeV}^5 \, , \nonumber\\ \lambda_{X(4350)}&=&6.01^{+0.93}_{-0.80}\times 10^{-2}\,\rm{GeV}^5 \, , \end{eqnarray} at the energy scale $\mu=2.40\,\rm{GeV}$, where we have neglected the uncertainties out of control. \begin{figure} \centering \includegraphics[totalheight=9cm,width=12cm]{mass-3915-4500.EPS} \caption{ The masses $M_{X(3890)}$ and $M_{X(4350)}$ with variations of the Borel parameters $T^2$, where the experimental value denotes the experimental values of the masses $M_{X(3915)}$ and $M_{X(4500)}$. } \end{figure} \begin{figure} \centering \includegraphics[totalheight=9cm,width=12cm]{residue-3915-4500.EPS} \caption{ The residues $\lambda_{X(3890)}$ and $\lambda_{X(4350)}$ with variations of the Borel parameters $T^2$. } \end{figure} Then we take the central values of the masses and pole residues as the input parameters, and obtain the pole contributions of the $X(3890)$ and $X(4350)$ respectively, \begin{eqnarray} {\rm pole}_{X(3890)}&=&(72-90)\% \, , \nonumber \\ {\rm pole}_{X(4350)}&=&(8-15)\% \, , \end{eqnarray} at the energy scale $\mu=1.25\,\rm{GeV}$ and \begin{eqnarray} {\rm pole}_{X(3890)}&=&(55-76)\% \, , \nonumber \\ {\rm pole}_{X(4350)}&=&(23-38)\% \, , \end{eqnarray} at the energy scale $\mu=2.40\,\rm{GeV}$. The pole contribution of the $X(3890/4350)$ at $\mu=1.25/2.40\,\rm{GeV}$ is larger than that at $\mu=2.40/1.25\,\rm{GeV}$, we prefer to extract the mass and pole residue of the $X(3890/4350)$ from the QCD spectral density at $\mu=1.25/2.40\,\rm{GeV}$ and neglect the ones at $\mu=2.40/1.25\,\rm{GeV}$. The neglected values of the masses do not satisfy the energy scale formula in Eq.(2). In this article, we choose the values \begin{eqnarray} M_{X(3890)}&=&3.85^{+0.18}_{-0.17}\,\rm{GeV} \, , \,\,\, {\rm Experimental\,\, value} \,\,\,\,3918.4\pm 1.9\,\rm{ MeV} \, \cite{PDG}\, , \nonumber\\ M_{X(4350)}&=&4.35^{+0.10}_{-0.11}\,\rm{GeV} \, , \,\,\, {\rm Experimental\,\, value} \,\,\,\,4506 \pm 11 ^{+12}_{-15}\, \rm {MeV}\, \cite{LHCb-4500-1606}\, ,\\ \lambda_{X(3890)}&=&1.84^{+0.76}_{-0.54}\times 10^{-2}\,\rm{GeV}^5 \, , \nonumber\\ \lambda_{X(4350)}&=&6.01^{+0.93}_{-0.80}\times 10^{-2}\,\rm{GeV}^5 \, , \end{eqnarray} which are also shown in Figs.3-4. In Figs.3-4, we plot the masses and pole residues with variations of the Borel parameter $T^2$ at much larger interval than the Borel window shown in Eq.(50). There appear platforms not very flat in the Borel window, which result in uncertainties $\delta M_{X_L}/M_{X_L},\,\delta M_{X_H}/M_{X_H}=\pm 1\%$, while the uncertainties originate from the Borel parameter outside of the Borel window are rather large compared to the central values. The predicted masses $M_{X(3890)}=3.85^{+0.18}_{-0.17}\,\rm{GeV}$ and $M_{X(4350)}=4.35^{+0.10}_{-0.11}\,\rm{GeV}$ satisfy the energy scale formula. From Fig.3, we can see that the predicted mass $M_{X(3890)}=3.85^{+0.18}_{-0.17}\,\rm{GeV}$ is compatible with the experimental value $M_{X(3915)}=3918.4\pm 1.9\,\rm{ MeV}$ \cite{PDG}, so it is reasonable to assign the $X(3890)$ to be the $X(3915)$. The predicted mass $M_{X(4350)}=4.35^{+0.10}_{-0.11}\,\rm{GeV}$ lies below the lower bound of the experimental value $M_{X(4500)} = 4506 \pm 11 ^{+12}_{-15}\, \rm {MeV}$ \cite{LHCb-4500-1606}, it is not favored to assign the $X(4500)$ to be the 2S $C\gamma_5\otimes \gamma_5C$ type scalar tetraquark state. Now we can draw the conclusion tentatively that the $X(3915)$ maybe have both $C\gamma_\mu\otimes\gamma^\mu C$ type and $C\gamma_5\otimes\gamma_5 C$ type tetraquark components, while the $X(4500)$ can be assigned to the 2S $C\gamma_\mu\otimes\gamma^\mu C$ type tetraquark state \cite{Wang1606}. In the Borel or Laplace QCD sum rules, we introduce a new parameter $T^2$ therefore an exponential factor $\exp\left( -\frac{s}{T^2}\right)$ to suppress the experimentally unknown higher resonances. The Borel parameter $T^2$ has no physical significance other than being a mathematical artifact, leads to rather narrow stability window for the hidden charm or hidden bottom tetraquark states. On the other hand, the continuum threshold $s_0$, which has a clear physical interpretation, is often exponentially suppressed $\exp\left( -\frac{s_0}{T^2}\right)$, or at best reduced in importance \cite{Dominguez-IJMPA}. While in the finite energy QCD sum rules or Hilbert moment QCD sum rules, the threshold $s_0$ shows a power-like welcome feature, it is interesting to study the hidden charm or hidden bottom tetraquark states with the finite energy QCD sum rules or Hilbert moment QCD sum rules, this may be our next work. \section{Conclusion} In this article, we study the $C\gamma_5\otimes \gamma_5C$ type and $C\otimes C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states with the QCD sum rules by calculating the contributions of the vacuum condensates up to dimension 10 in a consistent way. In calculations, we use the energy scale formula $\mu=\sqrt{M^2_{X/Y/Z}-(2{\mathbb{M}}_c)^2}$ to determine the optimal energy scales of the QCD spectral densities. The ground state masses $M_{C\gamma_5\otimes \gamma_5C}=3.89\pm 0.05\,\rm{GeV}$ and $M_{C\otimes C}=5.48\pm0.10\,\rm{GeV}$ support assigning the $X(3915)$ to be the 1S $C\gamma_5\otimes \gamma_5C$ type tetraquark state with $J^{PC}=0^{++}$, but do not support assigning the $X(4700)$ to be the 1S $C\otimes C$ type $cs\bar{c}\bar{s}$ tetraquark state with $J^{PC}=0^{++}$. Then we tentatively assign the $X(3915)$ and $X(4500)$ to be the $\rm 1S$ and $\rm 2S$ $C\gamma_5\otimes \gamma_5C$ type scalar $cs\bar{c}\bar{s}$ tetraquark states respectively, and obtain the $\rm{1S}$ mass $M_{\rm 1S}=3.85^{+0.18}_{-0.17}\,\rm{GeV}$ and $\rm{2S}$ mass $M_{\rm 2S}=4.35^{+0.10}_{-0.11}\,\rm{GeV}$ from the QCD sum rules, which support assigning the $X(3915)$ to be the $\rm{1S}$ $C\gamma_5\otimes \gamma_5C$ type tetraquark state, but do not support assigning the $X(4500)$ to be the ${\rm 2S}$ $C\gamma_5\otimes \gamma_5C$ type tetraquark state. The $X(3915)$ maybe have both $C\gamma_\mu\otimes\gamma^\mu C$ type and $C\gamma_5\otimes\gamma_5 C$ type tetraquark components, while the $X(4500)$ can be assigned to the 2S $C\gamma_\mu\otimes\gamma^\mu C$ type tetraquark state. \section*{Acknowledgements} This work is supported by National Natural Science Foundation, Grant Numbers 11375063, and Natural Science Foundation of Hebei province, Grant Number A2014502017.
7a28d1db7346e9425d54cfbcafef1da2904ae4ae
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Let ${\mathcal{P}}$ be a set of $v \ge 1$ elements, and let ${\mathcal{B}}$ be a set of $k$-subsets of ${\mathcal{P}}$, where $k$ is a positive integer with $1 \leq k \leq v$. Let $t$ be a positive integer with $t \leq k$. The pair ${\mathbb{D}} = ({\mathcal{P}}, {\mathcal{B}})$ is called a $t$-$(v, k, \lambda)$ {\em design\index{design}}, or simply {\em $t$-design\index{$t$-design}}, if every $t$-subset of ${\mathcal{P}}$ is contained in exactly $\lambda$ elements of ${\mathcal{B}}$. The elements of ${\mathcal{P}}$ are called points, and those of ${\mathcal{B}}$ are referred to as blocks. We usually use $b$ to denote the number of blocks in ${\mathcal{B}}$. A $t$-design is called {\em simple\index{simple}} if ${\mathcal{B}}$ does not contain repeated blocks. In this paper, we consider only simple $t$-designs. A $t$-design is called {\em symmetric\index{symmetric design}} if $v = b$. It is clear that $t$-designs with $k = t$ or $k = v$ always exist. Such $t$-designs are {\em trivial}. In this paper, we consider only $t$-designs with $v > k > t$. A $t$-$(v,k,\lambda)$ design is referred to as a {\em Steiner system\index{Steiner system}} if $t \geq 2$ and $\lambda=1$, and is denoted by $S(t,k, v)$. A necessary condition for the existence of a $t$-$(v, k, \lambda)$ design is that \begin{eqnarray}\label{eqn-tdesignnecessty} \binom{k-i}{t-i} \mbox{ divides } \lambda \binom{v-i}{t-i} \end{eqnarray} for all integer $i$ with $0 \leq i \leq t$. There has been an interplay between codes and $t$-designs for decades. The incidence matrix of any $t$-design spans a linear code over any finite field ${\mathrm{GF}}(q)$. A lot of progress in this direction has been made and documented in the literature (see, for examples, \cite{AK92}, \cite{DingBook}, \cite{Tonchev,Tonchevhb}). On the other hand, both linear and nonlinear codes may hold $t$-designs. Some linear and nonlinear codes were employed to construct $2$-designs and $3$-designs \cite{AK92,Tonchev,Tonchevhb}. Binary and ternary Golay codes of certain parameters hold $4$-designs and $5$-designs \cite{AK92}. However, the largest $t$ for which an infinite family of $t$-designs is derived directly from codes is $t=3$. It looks that not much progress on the construction of $t$-designs from codes has been made so far, while many other constructions of $t$-designs are documented in the literature (\cite{BJL,CMhb,KLhb,RR10}). The main objective of this paper is to construct infinite families of $2$-designs and $3$-designs from linear codes. In addition, we determine the parameters of some known $t$-designs, and present many conjectured infinite families of $2$-designs that are based on projective ternary cyclic codes. \section{The classical construction of $t$-designs from codes and highly nonlinear functions} Let ${\mathcal{C}}$ be a $[v, \kappa, d]$ linear code over ${\mathrm{GF}}(q)$. Let $A_i:=A_i({\mathcal{C}})$, which denotes the number of codewords with Hamming weight $i$ in ${\mathcal{C}}$, where $0 \leq i \leq v$. The sequence $(A_0, A_1, \cdots, A_{v})$ is called the \textit{weight distribution} of ${\mathcal{C}}$, and $\sum_{i=0}^v A_iz^i$ is referred to as the \textit{weight enumerator} of ${\mathcal{C}}$. For each $k$ with $A_k \neq 0$, let ${\mathcal{B}}_k$ denote the set of supports of all codewords of Hamming weight $k$ in ${\mathcal{C}}$, where the coordinates of a codeword are indexed by $(0,1,2, \cdots, v-1)$. Let ${\mathcal{P}}=\{0, 1, 2, \cdots, v-1\}$. The pair $({\mathcal{P}}, {\mathcal{B}}_k)$ may be a $t$-$(v, k, \lambda)$ design for some positive integer $\lambda$. The following theorems, developed by Assumus and Mattson, show that the pair $({\mathcal{P}}, {\mathcal{B}}_k)$ defined by a linear code is a $t$-design under certain conditions. \begin{theorem}\label{thm-AM1}[Assmus-Mattson Theorem \cite{AM74}, \cite[p. 303]{HP03}] Let ${\mathcal{C}}$ be a binary $[v, \kappa, d]$ code. Suppose ${\mathcal{C}}^\perp$ has minimum weight $d^\perp$. Suppose that $A_i=A_i({\mathcal{C}})$ and $A_i^\perp=A_i({\mathcal{C}}^\perp)$, for $0 \leq i \leq v$, are the weight distributions of ${\mathcal{C}}$ and ${\mathcal{C}}^\perp$, respectively. Fix a positive integer $t$ with $t < d$, and let $s$ be the number of $i$ with $A_i^\perp \ne 0$ for $0 < i \leq v-t$. Suppose that $s \leq d -t$. Then \begin{itemize} \item the codewords of weight $i$ in ${\mathcal{C}}$ hold a $t$-design provided that $A_i \ne 0$ and $d \leq i \leq v$, and \item the codewords of weight $i$ in ${\mathcal{C}}^\perp$ hold a $t$-design provided that $A_i^\perp \ne 0$ and $d^\perp \leq i \leq v$. \end{itemize} \end{theorem} The Assmus-Mattson Theorem for nonbinary codes is given as follows [Assmus-Mattson Theorem \cite{AM74}, \cite[p. 303]{HP03}] \begin{theorem}\label{thm-AM2} Let ${\mathcal{C}}$ be a $[v, \kappa, d]$ code over ${\mathrm{GF}}(q)$. Suppose ${\mathcal{C}}^\perp$ has minimum weight $d^\perp$. Let $w$ be the largest integer with $w \leq v$ satisfying $$ w - \left\lfloor \frac{w+q-2}{q-1} \right\rfloor < d. $$ (So $w=v$ when $q=2$.) Define $w^\perp$ analogously using $d^\perp$. Suppose that $A_i=A_i({\mathcal{C}})$ and $A_i^\perp=A_i({\mathcal{C}}^\perp)$, for $0 \leq i \leq v$, are the weight distributions of ${\mathcal{C}}$ and ${\mathcal{C}}^\perp$, respectively. Fix a positive integer $t$ with $t < d$, and let $s$ be the number of $i$ with $A_i^\perp \ne 0$ for $0 < i \leq v-t$. Suppose that $s \leq d -t$. Then \begin{itemize} \item the codewords of weight $i$ in ${\mathcal{C}}$ hold a $t$-design provided that $A_i \ne 0$ and $d \leq i \leq w$, and \item the codewords of weight $i$ in ${\mathcal{C}}^\perp$ hold a $t$-design provided that $A_i^\perp \ne 0$ and $d^\perp \leq i \leq \min\{v-t, w^\perp\}$. \end{itemize} \end{theorem} The Assmus-Mattson Theorems documented above are very powerful tools in constructing $t$-designs from linear codes. We will employ them heavily in this paper. It should be noted that the conditions in Theorems \ref{thm-AM1} and \ref{thm-AM2} are sufficient, but not necessary for obtaining $t$-designs. To construct $t$-designs via Theorems \ref{thm-AM1} and \ref{thm-AM2}, we will need the following lemma in subsequent sections, which is a variant of the MacWilliam Identity \cite[p. 41]{vanLint}. \begin{theorem} \label{thm-MI} Let ${\mathcal{C}}$ be a $[v, \kappa, d]$ code over ${\mathrm{GF}}(q)$ with weight enumerator $A(z)=\sum_{i=0}^v A_iz^i$ and let $A^\perp(z)$ be the weight enumerator of ${\mathcal{C}}^\perp$. Then $$A^\perp(z)=q^{-\kappa}\Big(1+(q-1)z\Big)^vA\Big(\frac {1-z} {1+(q-1)z}\Big).$$ \end{theorem} A function $f$ from ${\mathrm{GF}}(q^m)$ to itself is called {\em planar} or \textit{perfect nonlinear} (PN) if \[\max_{0\ne a\in{\mathrm{GF}}(q^m)}\max_{b\in{\mathrm{GF}}(q^m)}|\{x\in{\mathrm{GF}}(q^m): f(x+a)-f(x)=b\}|=1,\] and {\em almost perfect nonlinear} (APN) if \[\max_{0\ne a\in{\mathrm{GF}}(q^m)}\max_{b\in{\mathrm{GF}}(q^m)}|\{x\in{\mathrm{GF}}(q^m): f(x+a)-f(x)=b\}|=2.\] Later in this paper, we will employ such functions in the constructions of linear codes and thus our constructions of $t$-designs. \section{Infinite families of $3$-designs from the binary RM codes}\label{sec-brmdesigns} It was known that Reed-Muller codes give families of $3$-$(2^m, k, \lambda)$ designs (\cite[Chapter 15]{MS77}, \cite{Tonchevhb}). However, the parameters of $k$ and $\lambda$ may not be specifically given in the literature. The purpose of this section is to determine the parameters of some $3$-designs derived from binary Reed-Muller codes. We use ${\mathrm{RM}}(r, m)$ to denote the binary Reed-Muller code of length $2^m$ and order $r$. Note that ${\mathrm{RM}}(m-r, m)^\perp={\mathrm{RM}}(r-1, m)$, where $2 \leq r < m$. The definition and information about binary Reed-Muller codes can be found in \cite[Section 4.5]{vanLint} and \cite[Chapters 13 and 14]{MS77}. \begin{lemma}\label{lem-RMwt1} The weight distribution of ${\mathrm{RM}}(m-2, m)$ (except $A_i=0$) is given by \begin{eqnarray*} A_{4k}= \frac{1}{2^{m+1}} \left[2\binom{2^m}{4k} + (2^{m+1}-2) \binom{2^{m-1}}{2k}\right] \end{eqnarray*} for $0 \leq k \leq 2^{m-2}$, and by \begin{eqnarray*} A_{4k+2}= \frac{1}{2^{m+1}} \left[2\binom{2^m}{4k+2} - (2^{m+1}-2) \binom{2^{m-1}}{2k+1}\right] \end{eqnarray*} for $0 \leq k \leq 2^{m-2}-1$. \end{lemma} \begin{proof} It is well known that the weight enumerator of ${\mathrm{RM}}(1, m)$ is $$ 1 + (2^{m+1}-2)z^{2^{m-1}} + z^{2^m}. $$ By Theorem \ref{thm-MI}, the weight enumerator of RM$(m-2, m)$, which is the dual of RM$(1, m)$, is given by \begin{eqnarray*} B(z) &=& \frac{1}{2^{m+1}} (1+z)^{2^m}\left[ 1 + (2^{m+1}-2) \left(\frac {1-z} {1+z}\right)^{2^{m-1}} + \left(\frac {1-z} {1+z}\right)^{2^m} \right] \\ &=& \frac{1}{2^{m+1}} \left[ (1+z)^{2^m} + (2^{m+1}-2) (1-z^2)^{2^{m-1}} + (1-z)^{2^m} \right] \\ &=& \frac{1}{2^{m+1}} \left[ 2 \sum_{i=0}^{2^{m-1}} \binom{2^m}{2i} z^{2i} +(2^{m+1}-2) \sum_{i=0}^{2^{m-1}} \binom{2^{m-1}}{i} (-1)^i z^{2i}\right] \\ &=&\frac{1}{2^{m+1}} \sum_{k=0}^{2^{m-2}} \left[ 2\binom{2^m}{4k} + (2^{m+1}-2) \binom{2^{m-1}}{2k} \right] z^{4k} + \\ & & \frac{1}{2^{m+1}}\sum_{k=0}^{2^{m-2}-1} \left[ 2\binom{2^m}{4k+2} - (2^{m+1}-2) \binom{2^{m-1}}{2k+1}\right] z^{4k+2}. \end{eqnarray*} The desired conclusion then follows. \end{proof} The following theorem gives parameters of all the $3$-designs in both ${\mathrm{RM}}(m-2, m)$ and ${\mathrm{RM}}(1, m)$. \begin{theorem}\label{thm-brmdesign1} Let $m \geq 3$. Then ${\mathrm{RM}}(m-2, m)$ has dimension $2^m -m-1$ and minimum distance $4$. For even positive integer $\kappa$ with $4 \leq \kappa \leq 2^m-4$, the supports of the codewords with weight $\kappa$ in ${\mathrm{RM}}(m-2, m)$ hold a $3$-$(2^m, \kappa, \lambda)$ design, where \begin{eqnarray*} \lambda=\left\{ \begin{array}{ll} \frac{\frac{1}{2^{m+1}}\binom{\kappa}{3}\left(2\binom{2^m}{4k} + (2^{m+1}-2) \binom{2^{m-1}}{2k} \right)}{\binom{2^m}{3}} & \mbox{ if } \kappa = 4k, \\ \frac{\frac{1}{2^{m+1}}\binom{\kappa}{3}\left(2\binom{2^m}{4k+2} - (2^{m+1}-2) \binom{2^{m-1}}{2k+1} \right)}{\binom{2^m}{3}} & \mbox{ if } \kappa = 4k+2. \end{array} \right. \end{eqnarray*} The supports of all codewords of weight $2^{m-1}$ in ${\mathrm{RM}}(1, m)$ hold a $3$-$(2^m, 2^{m-1}, 2^{m-2}-1)$ design. \end{theorem} \begin{proof} Note that the weight distribution of ${\mathrm{RM}}(1,m)$ is given by $$ A_0=1, \ A_{2^m}=1, \ A_{2^{m-1}}=2^{m+1}-2, \ \mbox{ and } A_i =0 \mbox{ for all other $i$.} $$ It is known that the minimum distance $d$ of ${\mathrm{RM}}(m-2, m)$ is equal to $4$. Put $t=3$. The number of $i$ with $A_i^\perp \neq 0$ and $1 \leq i \leq 2^m -3$ is $s=1$. Hence, $s=d-t$. Notice that two binary vectors have the same support if and only if they are equal. The desired conclusions then follow from Theorem \ref{thm-AM1} and Lemma \ref{lem-RMwt1}. \end{proof} As a corollary of Theorem \ref{thm-brmdesign1}, we have the following \cite[p. 63]{MS77}, which is well known. \begin{corollary} The minimum weight codewords in ${\mathrm{RM}}(m-2, m)$ form a $3$-$(2^m, 4, 1)$ design, i.e., a Steiner system. \end{corollary} The following theorem is also well known, and tells us that Reed-Muller codes give much more $3$-designs \cite{Tonchevhb}. \begin{theorem}\label{thm-brmdesign2} Let $m \geq 4$ and $2 \leq r < m$. Then ${\mathrm{RM}}(m-r, m)$ has dimension $2^m -\sum_{i=0}^{r-1} \binom{m}{i}$ and minimum distance $2^r$. For every nonzero weight $\kappa$ in ${\mathrm{RM}}(m-r, m)$, the codewords of weight $\kappa$ in RM$(m-r, m)$ hold a $3$-$(2^m, \kappa, \lambda)$ design. \end{theorem} \begin{proof} Since $2 \leq r < m$, by Theorem 24 in \cite[p. 400]{MS77}, the automorphism group of ${\mathrm{RM}}(m-r, m)$ is triply transitive. The desired conclusion then follows from Theorem 8.4.7 in \cite[p. 308]{HP03}. \end{proof} Determining the weight distribution of ${\mathrm{RM}}(m-r, m)$ may be hard for $3 \leq r \leq m-3$ in general. Therefore, it may be difficult to find out the parameters $(\kappa, \lambda)$ of all the $3$-designs. The following problem is open in general. \begin{open} Determine the weight distribution of ${\mathrm{RM}}(m-r, m)$ for $3 \leq r \leq m-3$. \end{open} Some progress on the open problem above was made by Kasami and Tokura \cite{KT} and Kasami, Tokura and Azumi \cite{KTA}. Detailed information on this problem can be found in \cite[Chapter 15]{MS77}. \section{Designs from cyclic Hamming codes}\label{sec-hmdesigns} Let $\alpha$ be a generator of ${\mathrm{GF}}(q^m)^*$. Set $\beta=\alpha^{q-1}$. Let $g(x)$ be the minimal polynomial of $\beta$ over ${\mathrm{GF}}(q)$. Let ${\mathcal{C}}_{(q,m)}$ denote the cyclic code of length $v=(q^m-1)/(q-1)$ over ${\mathrm{GF}}(q)$ with generator polynomial $g(x)$. Then ${\mathcal{C}}_{(q,m)}$ has parameters $[(q^m-1)/(q-1), (q^m-1)/(q-1)-m, d]$, where $d \in \{2,3\}$. When $\gcd(q-1, m)=1$, ${\mathcal{C}}_{(q,m)}$ has minimum weight $3$ and is equivalent to the Hamming code. \begin{lemma}\label{lem-HCwt} The weight distribution of ${\mathcal{C}}_{(q,m)}$ is given by \begin{eqnarray*} A_{k}= \frac{1}{q^m} \sum_{\substack {0 \le i \le (q^{m-1}-1)/(q-1) \\0 \le j \le q^{m-1} \\ i+j=k}}\left[\binom{\frac{q^{m-1}-1}{q-1}}{i} \binom{q^{m-1}}{j}\Big((q-1)^k+(-1)^j(q-1)^i(q^m-1)\Big)\right] \end{eqnarray*} for $0 \leq k \leq (q^m-1)/(q-1)$. \end{lemma} \begin{proof} ${\mathcal{C}}_{(q,m)}^\perp$ is the simplex code, as $\gcd(q-1, (q^{m}-1)/(q-1))=1$. Its weight enumerator is $$ 1+(q^m-1)z^{q^{m-1}}. $$ By Theorem \ref{thm-MI}, the weight enumerator of ${\mathcal{C}}_{(q,m)}$ is given by \begin{eqnarray*} A(z) &=& \frac{1}{q^m} (1+(q-1)z)^{v}\left[ 1 + (q^{m}-1) \left(\frac {1-z} {1+(q-1)z}\right)^{q^{m-1}} \right] \\ &=& \frac{1}{q^m} \left[ (1+(q-1)z)^{v} + (q^m-1) (1-z)^{q^{m-1}}(1+(q-1)z)^{\frac {q^{m-1}-1}{q-1}} \right] \\ &=& \frac{1}{q^m} (1+(q-1)z)^{\frac {q^{m-1}-1}{q-1}} \left[ (1+(q-1)z)^{q^{m-1}} + (q^m-1) (1-z)^{q^{m-1}} \right]. \end{eqnarray*} The desired conclusion then follows. \end{proof} A code of minimum distance $d=2e+1$ is \textit{perfect}, if the spheres of radius $e$ around the codewords cover the whole space. The following theorem introduces a relation between perfect codes and $t$-designs and is due to Assmus and Mattson \cite{AM74}. \begin{theorem}\label{thm-perfectcodedesign} A linear $q$-ary code of length $v$ and minimum distance $d=2e+1$ is perfect if and only if the supports of the codewords of minimum weight form a simple $(e+1)$-$(v, 2e+1, (q-1)^e)$ design. In particular, the minimum weight codewords in a linear or nonlinear perfect code, which contains the zero vector, form a Steiner system $S(e+1, 2e+1, v)$. \end{theorem} It is known that the Hamming code over ${\mathrm{GF}}(q)$ is perfect, and the codewords of weight $3$ hold a $2$-design by Theorem \ref{thm-perfectcodedesign}. The $2$-designs documented in the following theorem may be viewed as an extension of this result. \begin{theorem}\label{thm-HMdesign171} Let $m \geq 3$ and $q = 2$ or $m \geq 2$ and $q >2$, and let $\gcd(q-1, m)=1$. Let ${\mathcal{P}}=\{0,1,2, \cdots, (q^m-q)/(q-1)\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords of Hamming weight $k$ with $A_k \neq 0$ in ${\mathcal{C}}_{(q,m)}$, where $3 \leq k \leq w$ and $w$ is the largest such that $w-\lfloor (w+q-2)/(q-1) \rfloor < 3$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$((q^m-1)/(q-1), k, \lambda)$ design. In particular, the supports of codewords of weight $3$ in ${\mathcal{C}}_{(q,m)}$ form a $2$-$((q^m-1)/(q-1), 3, q-1)$ design. The supports of all codewords of weight $q^{m-1}$ in ${\mathcal{C}}_{(q,m)}^\perp$ form a $2$-$((q^m-1)/(q-1), q^{m-1}, \lambda)$ design, where $$ \lambda=(q-1)q^{m-2}. $$ \end{theorem} \begin{proof} ${\mathcal{C}}_{(q,m)}^\perp$ is the simplex code, as $\gcd(q-1, (q^{m}-1)/(q-1))=1$. Its weight enumerator is $$ 1+(q^m-1)z^{q^{m-1}}. $$ A proof of this weight enumerator is straightforward and can be found in \cite[Theorem 15]{DY13}. Recall now Theorem \ref{thm-AM2} and the definition of $w$ for ${\mathcal{C}}_{(q,m)}$ and $w^\perp$ for ${\mathcal{C}}_{(q,m)}^\perp$. Since ${\mathcal{C}}_{(q,m)}$ has minimum weight $3$. Given that the weight enumerator of ${\mathcal{C}}_{(q,m)}^\perp$ is $1+(q^m-1)z^{q^{m-1}},$ we deduce that $w^\perp=q^{m-1}$. Put $t=2$. It then follows that $s=1=d-t$. The desired conclusion on the $2$-design property then follows from Theorem \ref{thm-AM2} and Lemma \ref{lem-HCwt}. We now prove that the supports of codewords of weight $3$ in ${\mathcal{C}}_{(q,m)}$ form a $2$-$((q^m-1)/(q-1), 3, q-1)$ design. We have already proved that these supports form a $2$-$((q^m-1)/(q-1), 3, \lambda)$ design. To determine the value $\lambda$ for this design, we need to compute the total number $b$ of blocks in this design. To this end, we first compute the total number of codewords of weight $3$ in ${\mathcal{C}}_{(q,m)}$. It follows from Lemma \ref{lem-HCwt} that $$ A_3=\frac{(q^m-1)(q^m-q)}{6}. $$ Since $3$ is the minimum nonzero weight in ${\mathcal{C}}_{(q,m)}$, it is easy to see that two codewords of weight $3$ in ${\mathcal{C}}_{(q,m)}$ have the same support if and only one is a scalar multiple of another. Thus, the total number $b$ of blocks is given by $$ b:=\frac{A_3}{q-1}=\frac{(q^m-1)(q^m-q)}{6(q-1)}. $$ It then follows that $$ \lambda=\frac{b\binom{3}{2}}{\binom{\frac{q^m-1}{q-1}}{2}}=q-1. $$ Let $\alpha$ be a generator of ${\mathrm{GF}}(q^m)^*$, and set $\beta=\alpha^{q-1}$. Then $\beta$ is a $v$-th primitive root of unity, where $v=(q^m-1)/(q-1)$. It is known that $$ {\mathcal{C}}_{(q,m)}^\perp=\{{\mathbf{c}}_u: u \in {\mathrm{GF}}(q^m)\}, $$ where ${\mathbf{c}}_u=(({\mathrm{Tr}}(u), {\mathrm{Tr}}(u\beta), \cdots, {\mathrm{Tr}}(u\beta^{v-1}))$ and ${\mathrm{Tr}}(x)$ is the trace function from ${\mathrm{GF}}(q^m)$ to ${\mathrm{GF}}(q)$. It is then easily seen that ${\mathbf{c}}_{u}$ and ${\mathbf{c}}_{v}$ have the same support if and only if $u=av$ for some $a \in {\mathrm{GF}}(q)^*$. We then deduce that the total number $b^\perp$ of blocks in the design is given by $$ b^\perp = \frac{q^m-1}{q-1}. $$ Consequently, $$ \lambda^\perp = \frac{\frac{q^m-1}{q-1} \binom{q^{m-1}}{2}}{\binom{\frac{q^{m-1}-1}{q-1}}{2}} =(q-1)q^{m-2}. $$ Thus, the supports of all codewords of weight $q^{m-1}$ in ${\mathcal{C}}_{(q,m)}^\perp$ form a $2$-design with parameters $$ \left((q^m-1)/(q-1), \ q^{m-1}, \ (q-1)q^{m-2} \right). $$ \end{proof} Theorem \ref{thm-HMdesign171} tells us that for some $k \geq 3$ with $A_k \neq 0$, the supports of the codewords with weight $k$ in ${\mathcal{C}}_{(q,m)}$ form $2$-$((q^m-1)/(q-1), k, \lambda)$ design. However, it looks complicated to determine the parameter $\lambda$ corresponding to this $k \geq 4$. We draw the reader's attention to the following open problem. \begin{open} Let $q \geq 3$ and $m \geq 2$. For $k \geq 4$ with $A_k \neq 0$, determine the value $\lambda$ in the $2$-$((q^m-1)/(q-1), k, \lambda)$ design, formed by the supports of the codewords with weight $k$ in ${\mathcal{C}}_{(q,m)}$. \end{open} Notice that two binary codewords have the same support if and only if they are equal. When $q=2$, Theorem \ref{thm-HMdesign171} becomes the following. \begin{corollary}\label{cor-HMdesign171} Let $m \geq 3$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords with Hamming weight $k$ in ${\mathcal{C}}_{(2,m)}$, where $3 \leq k \leq 2^m-3$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, k, \lambda)$ design, where $$ \lambda=\frac{(k-1)kA_k}{(2^m-1)(2^m-2)} $$ and $A_k$ is given in Lemma \ref{lem-HCwt}. The supports of all codewords of weight $2^{m-1}$ in ${\mathcal{C}}_{(2,m)}^\perp$ form a $2$-$(2^m-1, 2^{m-1}, 2^{m-2})$ design. \end{corollary} Corollary \ref{cor-HMdesign171} says that each binary Hamming code ${\mathcal{C}}_{(2,m)}$ and its dual code give a total number $2^m-4$ of $2$-designs with varying block sizes. The following are examples of the $2$-designs held in the binary Hamming code. \begin{example} Let $m \geq 4$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords with Hamming weight $3$ in ${\mathcal{C}}_{(2,m)}$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, \, 3, \, 1)$ design. \end{example} \begin{proof} By Lemma \ref{lem-HCwt}, we have $$ A_3=\frac{(2^{m-1}-1)(2^m-1)}{3}. $$ The desired value for $\lambda$ then follows from Corollary \ref{cor-HMdesign171}. \end{proof} \begin{example}\label{exam-hmdesign4} Let $m \geq 4$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords with Hamming weight $4$ in ${\mathcal{C}}_{(2,m)}$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, \, 4, \, 2^{m-1}-2)$ design. \end{example} \begin{proof} By Lemma \ref{lem-HCwt}, we have $$ A_4=\frac{(2^{m-1}-1)(2^{m-1}-2)(2^m-1)}{6}. $$ The desired value for $\lambda$ then follows from Corollary \ref{cor-HMdesign171}. \end{proof} \begin{example}\label{exam-hmdesign5} Let $m \geq 4$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords with Hamming weight $5$ in ${\mathcal{C}}_{(2,m)}$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, \, 5, \, \lambda)$ design, where $$ \lambda=\frac{2(2^{m-1}-2)(2^{m-1}-4)}{3} $$ \end{example} \begin{proof} By Lemma \ref{lem-HCwt}, we have $$ A_5=\frac{(2^{m-1}-1)(2^{m-1}-2)(2^{m-1}-4)(2^m-1)}{15}. $$ The desired value for $\lambda$ then follows from Corollary \ref{cor-HMdesign171}. \end{proof} \begin{example}\label{exam-hmdesign6} Let $m \geq 4$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords with Hamming weight $6$ in ${\mathcal{C}}_{(2,m)}$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, \, 6, \, \lambda)$ design, where $$ \lambda=\frac{(2^{m-1}-2)(2^{m-1}-3)(2^{m-1}-4)}{3} $$ \end{example} \begin{proof} By Lemma \ref{lem-HCwt}, we have $$ A_6=\frac{(2^{m-1}-1)(2^{m-1}-2)(2^{m-1}-3)(2^{m-1}-4)(2^m-1)}{45}. $$ The desired value for $\lambda$ then follows from Corollary \ref{cor-HMdesign171}. \end{proof} \begin{example}\label{exam-hmdesign7} Let $m \geq 4$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords with Hamming weight $7$ in ${\mathcal{C}}_{(2,m)}$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, \, 7, \, \lambda)$ design, where $$ \lambda=\frac{(2^{m-1}-2)(2^{m-1}-3)(4 \times 2^{2(m-1)}-30 \times 2^{m-1} +71)}{30}. $$ \end{example} \begin{proof} By Lemma \ref{lem-HCwt}, we have $$ A_7=\frac{(2^{m-1}-1)(2^{m-1}-2)(2^{m-1}-3)(2^m-1)(4 \times 2^{2(m-1)}-30 \times 2^{m-1} +71)}{630}. $$ The desired value for $\lambda$ then follows from Corollary \ref{cor-HMdesign171}. \end{proof} \section{Designs from a class of binary codes with two zeros and their duals}\label{sec-newdesigns} In this section, we construct many infinite families of $2$-designs and $3$-designs with several classes of binary cyclic codes whose duals have two zeros. These binary codes are defined by almost perfect nonlinear (APN) functions over ${\mathrm{GF}}(2^m)$. \begin{table}[ht] \caption{Weight distribution for odd $m$.}\label{tab-CG1} \centering \begin{tabular}{ll} \hline Weight $w$ & No. of codewords $A_w$ \\ \hline $0$ & $1$ \\ $2^{m-1}-2^{(m-1)/2}$ & $(2^m-1)(2^{(m-1)/2}+1)2^{(m-3)/2}$ \\ $2^{m-1}$ & $(2^m-1)(2^{m-1}+1)$ \\ $2^{m-1}+2^{(m-1)/2}$ & $(2^m-1)(2^{(m-1)/2}-1)2^{(m-3)/2}$ \\ \hline \end{tabular} \end{table} \begin{lemma}\label{lem-TZwt} Let $m \geq 5$ be odd. Let ${\mathcal{C}}_m$ be a binary linear code of length $2^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG1}. Then the weight distribution of ${\mathcal{C}}_m$ is given by \begin{eqnarray*} 2^{2m}A_k&=& \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m-1)/2} \\ 0\le j \le 2^{m-1}+2^{(m-1)/2}-1 \\i+j=k}}(-1)^ia\binom{2^{m-1}-2^{(m-1)/2}} {i} \binom{2^{m-1}+2^{(m-1)/2}-1}{j}\\ & & + \binom {2^m-1}{k}+\sum_{\substack{0 \le i \le 2^{m-1} \\ 0\le j \le 2^{m-1}-1 \\i+j=k}}(-1)^ib\binom{2^{m-1}} {i}\binom{2^{m-1}-1} {j} \\ & & + \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m-1)/2} \\ 0\le j \le 2^{m-1}-2^{(m-1)/2}-1 \\i+j=k}}(-1)^ic\binom{2^{m-1}+2^{(m-1)/2}}{i}\binom{2^{m-1}-2^{(m-1)/2}-1}{j} \end{eqnarray*} for $0 \le k \le 2^m-1$, where \begin{eqnarray*} a &=& (2^m-1)(2^{(m-1)/2}+1)2^{(m-3)/2}, \\ b &=& (2^m-1)(2^{m-1}+1), \\ c &=& (2^m-1)(2^{(m-1)/2}-1)2^{(m-3)/2}. \end{eqnarray*} In addition, ${\mathcal{C}}_m$ has parameters $[2^m-1, 2^m-1-2m, 5]$. \end{lemma} \begin{proof} By assumption, the weight enumerator of ${\mathcal{C}}_m^\perp$ is given by $$ A^\perp(z)=1+az^{2^{m-1}-2^{(m-1)/2}}+bz^{2^{m-1}}+cz^{2^{m-1}+2^{(m-1)/2}}. $$ It then follows from Theorem \ref{thm-MI} that the weight enumerator of ${\mathcal{C}}_m$ is given by \begin{eqnarray*} A(z) &=& \frac{1}{2^{2m}} (1+z)^{2^m-1}\left[ 1 + a\left(\frac {1-z} {1+z}\right)^{2^{m-1}-2^{(m-1)/2}} \right] + \\ & & \frac{1}{2^{2m}} (1+z)^{2^m-1}\left[ b\left(\frac {1-z} {1+z}\right)^{2^{m-1}}+c\left(\frac {1-z} {1+z}\right)^{2^{m-1}+2^{(m-1)/2}} \right] \\ &=& \frac{1}{2^{2m}} \Bigg[ (1+z)^{2^m-1} + a(1-z)^{2^{m-1}-2^{(m-1)/2}}(1+z)^{2^{m-1}+2^{(m-1)/2}-1} \\ & & + b(1-z)^{2^{m-1}}(1+z)^{2^{m-1}-1} + c(1-z)^{2^{m-1}+2^{(m-1)/2}}(1+z)^{2^{m-1}-2^{(m-1)/2}-1} \Bigg]. \end{eqnarray*} Obviously, we have \begin{eqnarray*} (1+z)^{2^m-1} &=& \sum_{k=0}^{2^m-1} \binom{2^m-1}{k}z^k. \end{eqnarray*} It is easily seen that \begin{eqnarray*} \lefteqn{(1-z)^{2^{m-1}-2^{(m-1)/2}}(1+z)^{2^{m-1}+2^{(m-1)/2}-1} } \\ &=& \sum_{k=0}^{2^m-1} \left[ \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m-1)/2} \\ 0\le j \le 2^{m-1}+2^{(m-1)/2}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m-1)/2}} {i} \binom{2^{m-1}+2^{(m-1)/2}-1}{j} \right] z^k \end{eqnarray*} and \begin{eqnarray*} \lefteqn{(1-z)^{2^{m-1}+2^{(m-1)/2}}(1+z)^{2^{m-1}-2^{(m-1)/2}-1} } \\ &=& \sum_{k=0}^{2^m-1} \left[ \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m-1)/2} \\ 0\le j \le 2^{m-1}-2^{(m-1)/2}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m-1)/2}}{i}\binom{2^{m-1}-2^{(m-1)/2}-1}{j} \right] z^k. \end{eqnarray*} Similarly, we have \begin{eqnarray*} (1-z)^{2^{m-1}}(1+z)^{2^{m-1}-1}= \sum_{k=0}^{2^m-1} \left[ \sum_{\substack{0 \le i \le 2^{m-1} \\ 0\le j \le 2^{m-1}-1 \\i+j=k}}(-1)^i \binom{2^{m-1}} {i}\binom{2^{m-1}-1} {j} \right] z^k. \end{eqnarray*} Combining these formulas above yields the weight distribution formula for $A_k$. The weight distribution in Table \ref{tab-CG1} tells us that the dimension of ${\mathcal{C}}_m^\perp$ is $2m$. Therefore, the dimension of ${\mathcal{C}}_m$ is equal to $2^m-1-2m$. Finally, we prove that the minimum distance $d$ of ${\mathcal{C}}_m$ equals $5$. After tedious computations with the formula of $A_k$ given in Lemma \ref{lem-TZwt}, one can verify that $A_1=A_2=A_3=A_4=0$ and \begin{eqnarray}\label{eqn-minimumwt5} A_5=\frac{4\times 2^{3m-5} - 22 \times 2^{2m-4} + 26 \times 2^{m-3} -2}{15}. \end{eqnarray} When $m \geq 5$, we have $$ 4\times 2^{3m-5} = 4 \times 2^{m-1} 2^{2m-4} \geq 64 \times 2^{2m-4} > 22 \times 2^{2m-4} $$ and $$ 26 \times 2^{m-3} -2 >0. $$ Consequently, $A_5>0$ for all odd $m$. This proves that $d=5$. \end{proof} \begin{theorem}\label{thm-newdesigns1} Let $m \geq 5$ be odd. Let ${\mathcal{C}}_m$ be a binary linear code of length $2^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG1}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords of ${\mathcal{C}}_m$ with weight $k$, where $A_k \neq 0$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(2^m-1, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{k(k-1)A_k}{(2^m-1)(2^m-2)}, \end{eqnarray*} where $A_k$ is given in Lemma \ref{lem-TZwt}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-2\}$, and let ${\mathcal{B}}^\perp$ be the set of the supports of the codewords of ${\mathcal{C}}_m^\perp$ with weight $k$ and $A_k^\perp \neq 0$. Then $({\mathcal{P}}, {\mathcal{B}}^\perp)$ is a $2$-$(2^m-1, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{k(k-1)A_k^\perp}{(2^m-1)(2^m-2)}, \end{eqnarray*} where $A_k^\perp$ is given in Lemma \ref{lem-TZwt}. \end{theorem} \begin{proof} The weight distribution of ${\mathcal{C}}_m$ is given in Lemma \ref{lem-TZwt} and that of ${\mathcal{C}}_m^\perp$ is given in Table \ref{tab-CG1}. By Lemma \ref{lem-TZwt}, the minimum distance $d$ of ${\mathcal{C}}_m$ is equal to $5$. Put $t=2$. The number of $i$ with $A_i^\perp \neq 0$ and $1 \leq i \leq 2^m-1 -t$ is $s=3$. Hence, $s=d-t$. The desired conclusions then follow from Theorem \ref{thm-AM1} and the fact that two binary vectors have the same support if and only if they are equal. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then ${\mathcal{C}}_m^\perp$ gives three $2$-designs with the following parameters: \begin{itemize} \item $(v,\, k, \, \lambda)=\left(2^m-1,\ 2^{m-1}-2^{(m-1)/2}, \ 2^{m-3} (2^{m-1} - 2^{(m-1)/2} -1) \right).$ \item $(v, \, k, \, \lambda)=\left(2^m-1, \ 2^{m-1}+2^{(m-1)/2}, \ 2^{m-3} (2^{m-1} + 2^{(m-1)/2} -1) \right).$ \item $(v, \, k, \, \lambda)=\left(2^m-1, \ 2^{m-1}, \ (2^m-1)(2^{m-1}+1) \right).$ \end{itemize} \end{example} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $5$ in ${\mathcal{C}}_m$ give a $2$-$(2^m-1,\, 5,\, (2^{m-1}-4)/3)$ design. \end{example} \begin{proof} By Lemma \ref{lem-TZwt}, $$ A_5 = \frac{(2^{m-1}-1) (2^{m-1}-4) (2^m-1)}{30} $$ The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns1}. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $6$ in ${\mathcal{C}}_m$ give a $2$-$(2^m-1,\, 6,\, \lambda)$ design, where $$ \lambda= \frac{(2^{m-2}-2)(2^{m-1}-3)}{3}. $$ \end{example} \begin{proof} By Lemma \ref{lem-TZwt}, $$ A_6 = \frac{(2^{m-1}-1) (2^{m-1}-4) (2^{m-1}-3) (2^m-1)}{90} $$ The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns1}. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $7$ in ${\mathcal{C}}_m$ give a $2$-$(2^m-1,\, 7,\, \lambda)$ design, where $$ \lambda= \frac{2 \times 2^{3(m-1)} - 25 \times 2^{2(m-1)} + 123 \times 2^{m-1} - 190}{30}. $$ \end{example} \begin{proof} By Lemma \ref{lem-TZwt}, $$ A_7 = \frac{(2^{m-1}-1) (2^m-1) (2 \times 2^{3(m-1)} - 25 \times 2^{2(m-1)} + 123 \times 2^{m-1} - 190)}{630}. $$ The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns1}. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $8$ in ${\mathcal{C}}_m$ give a $2$-$(2^m-1,\, 8,\, \lambda)$ design, where $$ \lambda= \frac{ (2^{m-2}-2) (2 \times 2^{3(m-1)} - 25 \times 2^{2(m-1)} + 123 \times 2^{m-1} - 190)}{45}. $$ \end{example} \begin{proof} By Lemma \ref{lem-TZwt}, $$ A_8 = \frac{(2^{m-1}-1) (2^{m-1}-4) (2^m-1) (2 \times 2^{3(m-1)} - 25 \times 2^{2(m-1)} + 123 \times 2^{m-1} - 190)}{8 \times 315}. $$ The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns1}. \end{proof} \begin{lemma}\label{lem-TZEwt} Let $m \geq 5$ be odd. Let ${\mathcal{C}}_m$ be a linear code of length $2^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG1}. Denote by $\overline{{\mathcal{C}}}_m$ the extended code of ${\mathcal{C}}_m$ and let $\overline{{\mathcal{C}}}_m^\perp$ denote the dual of $\overline{{\mathcal{C}}}_m$. Then the weight distribution of $\overline{{\mathcal{C}}}_m$ is given by \begin{eqnarray*} 2^{2m+1}\overline{A}_k&=& (1+(-1)^k) \binom{2^m}{k} + \frac{1+(-1)^k}{2} (-1)^{\lfloor k/2 \rfloor} \binom{2^{m-1}}{\lfloor k/2 \rfloor} v + \\ & & u \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m-1)/2} \\ 0\le j \le 2^{m-1}+2^{(m-1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m-1)/2}} {i} \binom{2^{m-1}+2^{(m-1)/2}}{j} + \\ & & u \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m-1)/2} \\ 0\le j \le 2^{m-1}-2^{(m-1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m-1)/2}}{i}\binom{2^{m-1}-2^{(m-1)/2}}{j} \end{eqnarray*} for $0 \le k \le 2^m$, where $$ u=2^{2m-1}-2^{m-1} \mbox{ and } v = 2^{2m}+2^m-2. $$ In addition, $\overline{{\mathcal{C}}}_m$ has parameters $[2^m, 2^m-1-2m, 6]$. The code $\overline{{\mathcal{C}}}_m^\perp$ has weight enumerator \begin{eqnarray}\label{eqn-wtenumerator} \overline{A}^\perp(z) = 1+uz^{2^{m-1}-2^{(m-1)/2}}+vz^{2^{m-1}}+uz^{2^{m-1}+2^{(m-1)/2}}+z^{2^m}, \end{eqnarray} and parameters $[2^m, \ 2m+1, \ 2^{m-1}-2^{(m-1)/2}]$. \end{lemma} \begin{proof} It was proved in Lemma \ref{lem-TZwt} that ${\mathcal{C}}_m$ has parameters $[2^m-1, 2^m-1-2m, 5]$. By definition, the extended code $\overline{{\mathcal{C}}}_m$ has parameters $[2^m, 2^m-1-2m, 6]$. By Table \ref{tab-CG1}, all weights of ${\mathcal{C}}_m^\perp$ are even. Note that ${\mathcal{C}}_m^\perp$ has length $2^m-1$ and dimension $2m$, while $\overline{{\mathcal{C}}}_m^\perp$ has length $2^m$ and dimension $2m+1$. By definition, $\overline{{\mathcal{C}}}_m$ has only even weights. Therefore, the all-one vector must be a codeword in $\overline{{\mathcal{C}}}_m^\perp$. It can be shown that the weights in $\overline{{\mathcal{C}}}_m^\perp$ are the following: $$ 0, \ w_1,\ w_2, \ w_3,\ 2^m-w_1,\ 2^m-w_2, \ 2^m-w_3, \ 2^m, $$ where $w_1, w_2$ and $w_3$ are the three nonzero weights in ${\mathcal{C}}_m^\perp$. Consequently, $\overline{{\mathcal{C}}}_m^\perp$ has the following four weights $$ 2^{m-1}-2^{(m-1)/2}, \ 2^{m-1}, \ 2^{m-1}+2^{(m-1)/2}, \ 2^m. $$ Recall that $\overline{{\mathcal{C}}}_m$ has minimum distance $6$. Employing the first few Pless Moments, one can prove that the weight enumerator of $\overline{{\mathcal{C}}}_m^\perp$ is the one given in (\ref{eqn-wtenumerator}). By Theorem \ref{thm-MI}, the weight enumerator of $\overline{{\mathcal{C}}}_m$ is given by \begin{eqnarray}\label{eqn-j18-1} 2^{2m+1}\overline{A}(z) &=& (1+z)^{2^m}\left[ 1 + u\left(\frac {1-z} {1+z}\right)^{2^{m-1}-2^{(m-1)/2}}+ v\left(\frac {1-z} {1+z}\right)^{2^{m-1}} \right] + \nonumber \\ && (1+z)^{2^m}\left[ u\left(\frac {1-z} {1+z}\right)^{2^{m-1}+2^{(m-1)/2}} + \left(\frac{1-z}{1+z}\right)^{2^m} \right] \nonumber \\ &=& (1+z)^{2^m} + (1-z)^{2^m} + v (1-z^2)^{2^{m-1}} + \nonumber \\ & & u(1-z)^{2^{m-1}-2^{(m-1)/2}}(1+z)^{2^{m-1}+2^{(m-1)/2}} + \nonumber \\ & & u(1-z)^{2^{m-1}+2^{(m-1)/2}}(1+z)^{2^{m-1}-2^{(m-1)/2}} . \end{eqnarray} We now treat the terms in (\ref{eqn-j18-1}) one by one. We first have \begin{eqnarray}\label{eqn-j18-2} (1+z)^{2^m} + (1-z)^{2^m} = \sum_{k=0}^{2^m} \left(1+(-1)^k \right) \binom{2^m}{k}. \end{eqnarray} One can easily see that \begin{eqnarray}\label{eqn-j18-3} (1-z^2)^{2^{m-1}} = \sum_{i=0}^{2^{m-1}} (-1)^i \binom{2^{m-1}}{i} z^{2i} = \sum_{k=0}^{2^{m}} \frac{1+(-1)^k}{2} (-1)^{\lfloor k/2 \rfloor} \binom{2^{m-1}}{\lfloor k/2 \rfloor} z^{k}. \end{eqnarray} Notice that \begin{eqnarray*} (1-z)^{2^{m-1}-2^{(m-1)/2}}=\sum_{i=0}^{2^{m-1}-2^{(m-1)/2}} \binom{2^{m-1}-2^{(m-1)/2}}{i} (-1)^i z^i \end{eqnarray*} and \begin{eqnarray*} (1+z)^{2^{m-1}+2^{(m-1)/2}}=\sum_{i=0}^{2^{m-1}+2^{(m-1)/2}} \binom{2^{m-1}+2^{(m-1)/2}}{i} z^i \end{eqnarray*} We have then \begin{eqnarray}\label{eqn-j18-4} \lefteqn{(1-z)^{2^{m-1}-2^{(m-1)/2}} (1+z)^{2^{m-1}+2^{(m-1)/2}} } \nonumber \\ & & = \sum_{k=0}^{2^m} \left[ \sum_{\substack{0 \le i \le 2^{m-1}-2^{(m-1)/2} \\ 0\le j \le 2^{m-1}+2^{(m-1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}-2^{(m-1)/2}} {i} \binom{2^{m-1}+2^{(m-1)/2}}{j} \right] z^k. \end{eqnarray} Similarly, we have \begin{eqnarray}\label{eqn-j18-5} \lefteqn{(1-z)^{2^{m-1}+2^{(m-1)/2}} (1+z)^{2^{m-1}-2^{(m-1)/2}} } \nonumber \\ & & = \sum_{k=0}^{2^m} \left[ \sum_{\substack{0 \le i \le 2^{m-1}+2^{(m-1)/2} \\ 0\le j \le 2^{m-1}-2^{(m-1)/2} \\i+j=k}}(-1)^i \binom{2^{m-1}+2^{(m-1)/2}} {i} \binom{2^{m-1}-2^{(m-1)/2}}{j} \right] z^k. \end{eqnarray} Plugging (\ref{eqn-j18-2}), (\ref{eqn-j18-3}), (\ref{eqn-j18-4}), and (\ref{eqn-j18-5}) into (\ref{eqn-j18-1}) proves the desired conclusion. \end{proof} \begin{theorem}\label{thm-newdesigns2} Let $m \geq 5$ be odd. Let ${\mathcal{C}}_m$ be a linear code of length $2^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG1}. Denote by $\overline{{\mathcal{C}}}_m$ the extended code of ${\mathcal{C}}_m$ and let $\overline{{\mathcal{C}}}_m^\perp$ denote the dual of $\overline{{\mathcal{C}}}_m$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-1\}$, and let $\overline{{\mathcal{B}}}$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}}_m$ with weight $k$, where $\overline{A}_k \neq 0$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}})$ is a $3$-$(2^m, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{\overline{A}_k\binom{k}{3}}{\binom{2^m}{3}}, \end{eqnarray*} where $\overline{A}_k$ is given in Lemma \ref{lem-TZEwt}. Let ${\mathcal{P}}=\{0,1,2, \cdots, 2^m-1\}$, and let $\overline{{\mathcal{B}}}^\perp$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}}_m^\perp$ with weight $k$ and $\overline{A}_k^\perp \neq 0$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}}^\perp)$ is a $3$-$(2^m, k, \lambda)$ design, where \begin{eqnarray*} \lambda=\frac{\overline{A}_k^\perp\binom{k}{3}}{\binom{2^m}{3}}, \end{eqnarray*} where $\overline{A}_k^\perp$ is given in Lemma \ref{lem-TZEwt}. \end{theorem} \begin{proof} The weight distributions of $\overline{{\mathcal{C}}}_m$ and $\overline{{\mathcal{C}}}_m^\perp$ are described in Lemma \ref{lem-TZEwt}. Notice that the minimum distance $d$ of $\overline{{\mathcal{C}}}_m$ is equal to $6$. Put $t=3$. The number of $i$ with $\overline{A}_i^\perp \neq 0$ and $1 \leq i \leq 2^m -t$ is $s=3$. Hence, $s=d-t$. The desired conclusions then follow from Theorem \ref{thm-AM1} and the fact that two binary vectors have the same support if and only if they are identical. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then $\overline{{\mathcal{C}}}_m^\perp$ gives three $3$-designs with the following parameters: \begin{itemize} \item $(v,\, k, \, \lambda)=\left(2^m,\ 2^{m-1}-2^{(m-1)/2}, \ (2^{m-3}-2^{(m-3)/2}) (2^{m-1}-2^{(m-1)/2}-1) \right).$ \item $(v, \, k, \, \lambda)=\left(2^m, \ 2^{m-1}+2^{(m-1)/2}, \ (2^{m-3}+2^{(m-3)/2}) (2^{m-1}-2^{(m-1)/2}-1) \right).$ \item $(v, \, k, \, \lambda)=\left(2^m, \ 2^{m-1}, \ (2^{m-1}+1)(2^{m-2}-1) \right).$ \end{itemize} \end{example} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $6$ in $\overline{{\mathcal{C}}}_m$ give a $3$-$(2^m,\, 6,\, \lambda)$ design, where $$ \lambda= \frac{2^{m-1}-4}{3}. $$ \end{example} \begin{proof} By Lemma \ref{lem-TZEwt}, $$ \overline{A}_6 = \frac{2^{m-1} (2^{m-1}-1) (2^{m-1}-4) (2^m-1)}{90} $$ The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns2}. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $8$ in $\overline{{\mathcal{C}}}_m$ give a $3$-$(2^m,\, 8,\, \lambda)$ design, where $$ \lambda= \frac{2 \times 2^{3(m-1)} - 25 \times 2^{2(m-1)} + 123 \times 2^{m-1} - 190}{30}. $$ \end{example} \begin{proof} By Lemma \ref{lem-TZEwt}, $$ \overline{A}_8 = \frac{2^{m-1}(2^{m-1}-1) (2^m-1) (2 \times 2^{3(m-1)} - 25 \times 2^{2(m-1)} + 123 \times 2^{m-1} - 190)}{8 \times 315}. $$ The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns2}. \end{proof} \begin{example} Let $m \geq 5$ be odd. Then the supports of all codewords of weight $10$ in $\overline{{\mathcal{C}}}_m$ give a $3$-$(2^m,\, 10,\, \lambda)$ design, where $$ \lambda= \frac{ (2^{m-1}-4) (2 \times 2^{4(m-1)} - 34 \times 2^{3(m-1)} + 235 \times 2^{2(m-1)} - 931 \times 2^{m-1} + 1358)}{315}. $$ \end{example} \begin{proof} By Lemma \ref{lem-TZEwt}, $$ \overline{A}_{10} = \frac{2^{h} (2^{h}-1) (2^{h}-4) (2^{h+1}-1) (2\times 2^{4h} - 34 \times 2^{3h} + 235 \times 2^{2h} - 931 \times 2^{h} + 1358)}{4\times 14175}, $$ where $h=m-1$. The desired value for $\lambda$ then follows from Theorem \ref{thm-newdesigns2}. \end{proof} To demonstrate the existence of the $2$-designs and $3$-designs presented in Theorems \ref{thm-newdesigns1} and \ref{thm-newdesigns2}, respectively, we describe a list of binary codes that have the weight distribution of Table \ref{tab-CG1} below. Let $\alpha$ be a generator of ${\mathrm{GF}}(2^m)^*$. Let $g_s(x)=\mathbb{M}_1(x)\mathbb{M}_s(x)$, where $\mathbb{M}_i(x)$ is the minimal polynomial of $\alpha^i$ over ${\mathrm{GF}}(2)$. Let ${\mathcal{C}}_m$ denote the cyclic code of length $v=2^m-1$ over ${\mathrm{GF}}(2)$ with generator polynomial $g_s(x)$. It is known that ${\mathcal{C}}_m^\perp$ has dimension $2m$ and the weight distribution of Table \ref{tab-CG1} when $m$ is odd and $s$ takes on the following values \cite{DLLZ}: \begin{enumerate} \item $s=2^h+1$, where $\gcd(h, m)=1$ and $h$ is a positive integer. \item $s=2^{2h}-2^h+1$, where $h$ is a positive integer. \item $s=2^{(m-1)/2}+3$. \item $s=2^{(m-1)/2}+2^{(m-1)/4}-1$, where $m \equiv 1 \pmod{4}$. \item $s=2^{(m-1)/2}+2^{(3m-1)/4}-1$, where $m \equiv 3 \pmod{4}$. \end{enumerate} In all these cases, ${\mathcal{C}}_m$ has parameters $[2^m-1, 2^m-1-2m, 5]$ and is optimal. It is also known that the binary narrow-sense primitive BCH code with designed distance $2^{m-1}-2^{(m-1)/2}$ has also the weight distribution of Table \ref{tab-CG1} \cite{DFZ}. These codes and their extended codes give $2$-designs and $3$-designs when they are plugged into Theorems \ref{thm-newdesigns1} and \ref{thm-newdesigns2}. It is known that ${\mathcal{C}}_m$ has parameters $[2^m-1, 2^m-1-2m, 5]$ if and only if $x^e$ is an APN monomial over ${\mathrm{GF}}(2^m)$. However, even if $x^e$ is APN, the dual code ${\mathcal{C}}_m^\perp$ may have many weights, and thus the code ${\mathcal{C}}_m$ and its dual ${\mathcal{C}}_m^\perp$ may not give $2$-designs. One of such examples is the inverse APN monomial. \section{Infinite families of $2$-designs from a type of ternary linear codes}\label{sec-june28} In this section, we will construct infinite families of $2$-designs with a type of primitive ternary cyclic codes. \begin{table}[ht] \caption{Weight distribution of some ternary linear codes}\label{tab-CG3} \centering \begin{tabular}{|l|l|} \hline Weight $w$ & No. of codewords $A_w$ \\ \hline $0$ & $1$ \\ $2\times 3^{m-1}-3^{(m-1)/2}$ & $(3^m-1)(3^{m-1}+3^{(m-1)/2})$ \\ $2\times 3^{m-1}$ & $(3^m-1)(3^{m-1}+1)$ \\ $2\times 3^{m-1}+3^{(m-1)/2}$ & $(3^m-1)(3^{m-1}-3^{(m-1)/2})$ \\ \hline \end{tabular} \end{table} \begin{table}[ht] \caption{Weight distribution of some ternary linear codes}\label{tab-CG328} \centering \begin{tabular}{|l|l|} \hline Weight $w$ & No. of codewords $A_w$ \\ \hline $0$ & $1$ \\ $2\times 3^{m-1}-3^{(m-1)/2}$ & $3^{2m}-3^m$ \\ $2\times 3^{m-1}$ & $(3^m+3)(3^m-1)$ \\ $2\times 3^{m-1}+3^{(m-1)/2}$ & $3^{2m}-3^m$ \\ \hline $3^m$ & $2$ \\ \hline \end{tabular} \end{table} \begin{lemma}\label{lem-TZEwt28} Let $m \geq 3$ be odd. Assume that ${\mathcal{C}}_m$ is a ternary linear code of length $3^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG3}. Denote by $\overline{{\mathcal{C}}}_m$ the extended code of ${\mathcal{C}}_m$ and let $\overline{{\mathcal{C}}}_m^\perp$ denote the dual of $\overline{{\mathcal{C}}}_m$. Then we have the following conclusions. \begin{enumerate} \item The code ${\mathcal{C}}_m$ has parameters $[3^m-1, \, 3^m-1-2m, \, 4]$. \item The code ${\mathcal{C}}_m^\perp$ has parameters $[3^m-1, \, 2m, \, 2\times 3^{m-1}-3^{(m-1)/2}]$. \item The code $\overline{{\mathcal{C}}}_m^\perp$ has parameters $[3^m, \, 2m+1, \, 2\times 3^{m-1}-3^{(m-1)/2}]$, and its weight distribution is given in Table \ref{tab-CG328}. \item The code $\overline{{\mathcal{C}}}_m$ has parameters $[3^m, 3^m-1-2m, 5]$, and its weight distribution is given by \begin{eqnarray*} 3^{2m+1}\overline{A}_k &=& (2^k+(-1)^k 2) \binom{3^m}{k} + \\ & & v \sum_{\substack{0 \le i \le 2 \times 3^{m-1} \\ 0\le j \le 3^{m-1} \\i+j=k}}(-1)^i \binom{2 \times 3^{m-1}} {i} 2^j \binom{3^{m-1}}{j} + \\ & & u \sum_{\substack{0 \le i \le 2\times 3^{m-1}-3^{\frac{m-1}{2}} \\ 0\le j \le 3^{m-1}+3^{\frac{m-1}{2}} \\i+j=k}}(-1)^i \binom{2 \times 3^{m-1}-3^{\frac{m-1}{2}}} {i} 2^j \binom{3^{m-1}+3^{\frac{m-1}{2}}}{j} + \\ & & u \sum_{\substack{0 \le i \le 2 \times 3^{m-1}+3^{\frac{m-1}{2} } \\ 0\le j \le 3^{m-1}-3^{\frac{m-1}{2}} \\i+j=k}}(-1)^i \binom{2 \times 3^{m-1}+3^{\frac{m-1}{2}}}{i}2^j \binom{3^{m-1}-3^{\frac{m-1}{2}}}{j} \end{eqnarray*} for $0 \le k \le 3^m$, where $$ u=3^{2m}-3^{m} \mbox{ and } v = (3^m+3)(3^m-1). $$ \end{enumerate} \end{lemma} \begin{proof} The proof is similar to that of Lemma \ref{lem-TZEwt} and is omitted here. \end{proof} \begin{theorem}\label{thm-newdesigns228} Let $m \geq 3$ be odd. Let ${\mathcal{C}}_m$ be a linear code of length $3^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG3}. Denote by $\overline{{\mathcal{C}}}_m$ the extended code of ${\mathcal{C}}_m$ and let $\overline{{\mathcal{C}}}_m^\perp$ denote the dual of $\overline{{\mathcal{C}}}_m$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 3^m-1\}$, and let $\overline{{\mathcal{B}}}$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}}_m$ with weight $k$, where $5 \leq k \leq 10$ and $\overline{A}_k \neq 0$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}})$ is a $2$-$(3^m,\, k,\, \lambda)$ design for some $\lambda$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 3^m-1\}$, and let $\overline{{\mathcal{B}}}^\perp$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}}_m^\perp$ with weight $k$ and $\overline{A}_k^\perp \neq 0$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}}^\perp)$ is a $2$-$(3^m, \, k, \, \lambda)$ design for some $\lambda$. \end{theorem} \begin{proof} The weight distributions of $\overline{{\mathcal{C}}}_m$ and $\overline{{\mathcal{C}}}_m^\perp$ are described in Lemma \ref{lem-TZEwt28}. Notice that the minimum distance $d$ of $\overline{{\mathcal{C}}}_m$ is equal to $5$. Put $t=2$. The number of $i$ with $\overline{A}_i^\perp \neq 0$ and $1 \leq i \leq 3^m -t$ is $s=3$. Hence, $s=d-t$. The desired conclusions then follow from Theorem \ref{thm-AM2}. \end{proof} \begin{corollary} Let $m \geq 3$ be odd. Let ${\mathcal{C}}_m$ be a ternary linear code of length $3^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG3}. Denote by $\overline{{\mathcal{C}}}_m$ the extended code of ${\mathcal{C}}_m$ and let $\overline{{\mathcal{C}}}_m^\perp$ denote the dual of $\overline{{\mathcal{C}}}_m$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 3^m-1\}$, and let $\overline{{\mathcal{B}}}^\perp$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}}_m^\perp$ with weight $2\times 3^{m-1}-3^{(m-1)/2}$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}}^\perp)$ is a $2$-$(3^m, \, 2\times 3^{m-1}-3^{(m-1)/2}, \, \lambda)$, where $$ \lambda=\frac{(2\times 3^{m-1}- 3^{(m-1)/2})(2\times 3^{m-1}- 3^{(m-1)/2} -1)}{2}. $$ \end{corollary} \begin{proof} It follows from Theorem \ref{thm-newdesigns228} that $({\mathcal{P}}, \overline{{\mathcal{B}}}^\perp)$ is a $2$-design. We now determine the value of $\lambda$. Note that $\overline{{\mathcal{C}}}_m^\perp$ has minimum weight $2\times 3^{m-1}-3^{(m-1)/2}$. Any two codewords of minimum weight $2\times 3^{m-1}-3^{(m-1)/2}$ have the same support if and only if one is a scalar multiple of the other. Consequently, $$ \left|\overline{{\mathcal{B}}}^\perp \right|=\frac{3^{2m}-3^m}{2}. $$ It then follows that $$ \lambda=\frac{3^{2m}-3^m}{2}\frac{\binom{2\times 3^{m-1}-3^{(m-1)/2}}{2}}{\binom{3^m}{2}} = \frac{(2\times 3^{m-1}- 3^{(m-1)/2})(2\times 3^{m-1}- 3^{(m-1)/2} -1)}{2}. $$ \end{proof} \begin{corollary} Let $m \geq 3$ be odd. Let ${\mathcal{C}}_m$ be a ternary linear code of length $3^m-1$ such that its dual code ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{tab-CG3}. Denote by $\overline{{\mathcal{C}}}_m$ the extended code of ${\mathcal{C}}_m$ and let $\overline{{\mathcal{C}}}_m^\perp$ denote the dual of $\overline{{\mathcal{C}}}_m$. Let ${\mathcal{P}}=\{0,1,2, \cdots, 3^m-1\}$, and let $\overline{{\mathcal{B}}}$ be the set of the supports of the codewords of $\overline{{\mathcal{C}}}_m$ with weight $5$. Then $({\mathcal{P}}, \overline{{\mathcal{B}}})$ is a $2$-$(3^m,\, 5,\, \lambda)$ design, where $$ \lambda=\frac{5(3^{m-1}-1)}{2}. $$ \end{corollary} \begin{proof} It follows from Theorem \ref{thm-newdesigns228} that $({\mathcal{P}}, \overline{{\mathcal{B}}})$ is a $2$-design. We now determine the value of $\lambda$. Using the weight distribution formula in Lemma \ref{lem-TZEwt28}, we obtain that $$ \overline{A}_5=\frac{3^{3m-1}-4 \times 3^{2m-1}+3^m}{4}. $$ Recall that $\overline{{\mathcal{C}}}_m$ has minimum weight $5$. Any two codewords of minimum weight $5$ have the same support if and only if one is a scalar multiple of the other. Consequently, $$ \left|\overline{{\mathcal{B}}}^\perp \right|=\frac{\overline{A}_5}{2}. $$ It then follows that $$ \lambda= \frac{\overline{A}_5}{2} \frac{\binom{5}{2}}{\binom{3^m}{2}}=\frac{5(3^{m-1}-1)}{2}. $$ \end{proof} Theorem \ref{thm-newdesigns228} gives more $2$-designs. However, determining the corresponding value $\lambda$ may be hard, as the number of blocks in the design may be difficult to derive from $\overline{A}_k$ or $\overline{A}_k^\perp$. \begin{open} Determine the value of $\lambda$ of the $2$-$(3^m,\, k,\, \lambda)$ design for $6 \leq k \leq 10$, which are described in Theorem \ref{thm-newdesigns228}. \end{open} \begin{open} Determine the values of $\lambda$ of the $2$-$(3^m,\, 3^{m-1},\, \lambda)$ design and the $2$-$(3^m,\, 2\times 3^{m-1}-3^{(m-1)/2},\, \lambda)$ design, which are described in Theorem \ref{thm-newdesigns228}. \end{open} To demonstrate the existence of the $2$-designs presented in Theorem \ref{thm-newdesigns228}, we present a list of ternary cyclic codes that have the weight distribution of Table \ref{tab-CG3} below. Put $n=3^m-1$. Let $\alpha$ be a generator of ${\mathrm{GF}}(3^m)^*$. Let $g_s(x)=\mathbb{M}_{n-1}(x)\mathbb{M}_{n-s}(x)$, where $\mathbb{M}_i(x)$ is the minimal polynomial of $\alpha^i$ over ${\mathrm{GF}}(3)$. Let ${\mathcal{C}}_m$ denote the cyclic code of length $n=3^m-1$ over ${\mathrm{GF}}(3)$ with generator polynomial $g_s(x)$. It is known that ${\mathcal{C}}_m^\perp$ has dimension $2m$ and the weight distribution of Table \ref{tab-CG3} when $m$ is odd and $s$ takes on the following values \cite{CDY,YCD}: \begin{enumerate} \item $s=3^h+1$, $h \geq 0$ is an integer. \item $s=(3^h + 1)/2$, where $h$ is a positive integer and $\gcd(m, h)=1$. \end{enumerate} In these two cases, $x^s$ is a planar function on ${\mathrm{GF}}(3^m)$. Hence, these ternary codes are extremal in the sense that they are defined by planar functions whose differentiality is extremal. More classes of ternary codes such that their duals have the weight distribution of Table \ref{tab-CG3} are documented in \cite{DLLZ}. They give also $2$-designs via Theorem \ref{thm-newdesigns228}. There are also ternary cyclic codes with three weights but different weight distributions in \cite{DLLZ}. They may also hold $2$-designs. \section{Conjectured infinite families of $2$-designs from projective cyclic codes}\label{sec-conjectureddesigns} Throughout this section, let $m \geq 3$ be an odd integer, and let $v=(3^m-1)/2$. The objective of this section is to present a number of conjectured infinite families of $2$-designs derived from linear projective ternary cyclic codes. \begin{table} \begin{center} \caption{The weight distribution for odd $m \ge3$}\label{Tab-GG2} \begin{tabular}{|c|c|} \hline Weight & Frequency \\ \hline $0$ & $1$ \\ \hline $3^{m-1}-3^{(m-1)/2}$ & $\frac{(3^{m-1}+3^{(m-1)/2})(3^m-1)}{2}$ \\ \hline $3^{m-1}$ & $(3^m-3^{m-1}+1)(3^m-1)$ \\ \hline $3^{m-1}+3^{(m-1)/2}$ & $\frac{(3^{m-1}-3^{(m-1)/2})(3^m-1)}{2}$ \\ \hline \end{tabular} \label{table4} \end{center} \end{table} \begin{lemma}\label{lem-june21} Let ${\mathcal{C}}_m$ be a linear code of length $v$ over ${\mathrm{GF}}(3)$ such that its dual ${\mathcal{C}}_m^\perp$ has the weight distribution in Table \ref{Tab-GG2}. Then the weight distribution of ${\mathcal{C}}_m$ is given by \begin{eqnarray*} 3^{2m}A_k &=& \sum_{\substack{0 \le i \le 3^{m-1}-3^{(m-1)/2} \\ 0\le j \le \frac {3^{m-1}+2\cdot 3^{(m-1)/2}-1} {2} \\i+j=k}}(-1)^i2^ja\binom{3^{m-1}-3^{(m-1)/2}} {i} \binom{\frac {3^{m-1}+2 \cdot 3^{(m-1)/2}-1}{2}}{j}\\ & & + \binom {\frac {3^m-1} {2}}{k}2^k+ \sum_{\substack{0 \le i \le 3^{m-1} \\ 0\le j \le \frac {3^{m-1}-1} {2} \\i+j=k}}(-1)^i2^jb\binom{3^{m-1}} {i}\binom{\frac {3^{m-1}-1} {2}} {j} \\ & & + \sum_{\substack{0 \le i \le 3^{m-1}+3^{(m-1)/2} \\ 0\le j \le \frac {3^{m-1}-2 \cdot 3^{(m-1)/2}-1}{2} \\i+j=k}}(-1)^i2^jc\binom{3^{m-1}+3^{(m-1)/2}}{i}\binom{\frac {3^{m-1}-2 \cdot 3^{(m-1)/2}-1}{2}}{j} \end{eqnarray*} for $0 \le k \le \frac {3^m-1} {2}$, where \begin{eqnarray*} a &=& \frac{(3^{m-1}+3^{(m-1)/2})(3^m-1)}{2}, \\ b &=& (3^m-3^{m-1}+1)(3^m-1), \\ c &=& \frac{(3^{m-1}-3^{(m-1)/2})(3^m-1)}{2}. \end{eqnarray*} In addition, ${\mathcal{C}}_m$ has parameters $[(3^m-1)/2, (3^m-1)/2-2m, 4]$. \end{lemma} \begin{proof} Note that the weight enumerator of ${\mathcal{C}}_m^\perp$ is $$1+az^{3^{m-1}-3^{(m-1)/2}}+bz^{3^{m-1}}+cz^{3^{m-1}+3^{(m-1)/2}}.$$ The proof of this theorem is similar to that of Lemma \ref{lem-TZwt} and is omitted. \end{proof} Below we present two examples of ternary linear codes ${\mathcal{C}}_m$ such that their duals ${\mathcal{C}}_m^\perp$ have the weight distribution of Table \ref{Tab-GG2}. \begin{example}\label{exam-j271} Let $m \geq 3$ be odd. Let $\alpha$ be a generator of ${\mathrm{GF}}(3^m)^*$. Put $\beta=\alpha^2$. Let $\mathbb{M}_i(x)$ denote the minimal polynomial of $\beta^i$ over ${\mathrm{GF}}(3)$. Define $$ \delta=3^{m-1}-1-\frac{3^{(m+1)/2}-1}{2} $$ and $$ h(x)=(x-1){\mathrm{lcm}}(\mathbb{M}_1(x), \, \mathbb{M}_2(x), \, \cdots, \, \mathbb{M}_{\delta-1}(x)), $$ where ${\mathrm{lcm}}$ denotes the least common multiple of the polynomials. Let ${\mathcal{C}}_m$ denote the cyclic code of length $v=(3^m-1)/2$ over ${\mathrm{GF}}(3)$ with generator polynomial $g(x):=(x^v-1)/h(x)$. Then ${\mathcal{C}}_m$ has parameters $[(3^m-1)/2, (3^m-1)-2m, 4]$ and ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{Tab-GG2}. \end{example} \begin{proof} A proof of the desired conclusions was given in \cite{LDXG}. \end{proof} \begin{example}\label{exam-j272} Let $m \geq 3$ be odd. Let $\alpha$ be a generator of ${\mathrm{GF}}(3^m)^*$. Let $\beta=\alpha^2$. Let $g(x)=\mathbb{M}_{n-1}(x)\mathbb{M}_{n-2}(x)$, where $\mathbb{M}_i(x)$ is the minimal polynomial of $\beta^i$ over ${\mathrm{GF}}(3)$. Let ${\mathcal{C}}_m$ denote the cyclic code of length $v=(3^m-1)/2$ over ${\mathrm{GF}}(3)$ with generator polynomial $g(x)$. Then ${\mathcal{C}}_m$ has parameters $[(3^m-1)/2, (3^m-1)-2m, 4]$ and ${\mathcal{C}}_m^\perp$ has the weight distribution of Table \ref{Tab-GG2}. \end{example} \begin{proof} The desired conclusions can be proved similarly as Theorem 19 in \cite{LDXG}. \end{proof} \begin{conj}\label{conj-j261} Let ${\mathcal{P}}=\{0,1,2, \cdots, v-1\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords of ${\mathcal{C}}_m$ with Hamming weight $k$, where $A_k \neq 0$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(v, k, \lambda)$ design for all odd $m \ge 3$. \end{conj} \begin{conj}\label{conj-j262} Let ${\mathcal{P}}=\{0,1,2, \cdots, v-1\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords of ${\mathcal{C}}_m$ with Hamming weight $4$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a Steiner system $S(2, 4, (3^m-1)/2)$ for all odd $m \ge 3$. \end{conj} There are a survey on Steiner systems $S(2,4,v)$ \cite{RR10} and a book chapter on Steiner systems \cite{CMhb}. It is known that a Steiner system $S(2,4,v)$ exists if and only if $v \equiv 1 \mbox{ or } 4 \pmod{12}$ \cite{Hanani}. If Conjecture \ref{conj-j261} is true, so is Conjecture \ref{conj-j262}. In this case, a coding theory construction of a Steiner system $S(2, 4, (3^m-1)/2)$ for all odd $m \ge 3$ is obtained. \begin{conj}\label{conj-j263} Let ${\mathcal{P}}=\{0,1,2, \cdots, v-1\}$, and let ${\mathcal{B}}$ be the set of the supports of the codewords of ${\mathcal{C}}_m^\perp$ with Hamming weight $k$, where $A_k^\perp \neq 0$. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(v, k, \lambda)$ design for all odd $m \ge 3$. \end{conj} Even if some or all of the three conjectures are not true for ternary codes with the weight distribution of Table \ref{Tab-GG2}, these conjectures might still be valid for the two classes of ternary cyclic codes descried in Examples \ref{exam-j271} and \ref{exam-j272}. Note that Theorem \ref{thm-AM2} does not apply to the three conjectures above. We need to develop different methods for settling these conjectures. \section{Summary and concluding remarks} In the last section of this paper, we mention some applications of $t$-designs and summarize the main contributions of this paper. \subsection{Some applications of $2$-designs} Let ${\mathcal{P}}$ be an Abelian group of order $v$ under a binary operation denoted by $+$. Let ${\mathcal{B}}=\{B_1, B_2, \cdots, B_b\}$, where all $B_i$ are $k$-subsets of ${\mathcal{P}}$ and $k$ is a positive integer. We define $\Delta(B_i)$ to be the multiset $\{ x-y: x \in B_i,\ y \in B_i\}$. If every nonzero element of ${\mathcal{P}}$ appears exactly $\delta$ times in the multiset $\bigcup_{i=1}^b \Delta(B_i)$, we call ${\mathcal{B}}$ a $(v, k, \delta)$ difference family in $({\mathcal{P}}, +)$. The following theorems are straightforward and should be well known. \begin{theorem}\label{thm-june261} Let ${\mathcal{P}}$ be an Abelian group of order $v$ under a binary operation denoted by $+$. Let ${\mathcal{B}}=\{B_1, B_2, \cdots, B_b\}$, where all $B_i$ are $k$-subsets of ${\mathcal{P}}$ and $k$ is a positive integer. Then $({\mathcal{P}}, {\mathcal{B}})$ is a $2$-$(v, k, \lambda)$ design if and only if ${\mathcal{B}}$ is a $(v, k, \lambda v)$ difference family in $({\mathcal{P}}, +)$. \end{theorem} \begin{theorem}\label{thm-june262} Let $({\mathcal{P}}, {\mathcal{B}})$ be a $t$-$(v, k, \lambda)$ design, where ${\mathcal{P}}$ is an Abelian group. If $t \geq 2$, then ${\mathcal{B}}$ is a $(v, k, \delta)$ difference family in ${\mathcal{P}}$, where $$ \delta=\frac{v \lambda \binom{v-2}{t-2}}{\binom{k-2}{t-2}}. $$ \end{theorem} Difference families have applications in the design and analysis of optical orthogonal codes, frequency hopping sequences, and other engineering areas. By Theorems \ref{thm-june261} and \ref{thm-june262}, $t$-designs with $t \geq 2$ have also applications in these areas. In addition, $2$-designs give naturally linear codes \cite{AK92,DingBook}. These show the importance of $2$-designs in applications. \subsection{Summary} It is well known that binary Reed-Muller codes hold $3$-designs. Hence, the only contribution of Section \ref{sec-brmdesigns} is the determination of the specific parameters of the $3$-designs held in ${\mathrm{RM}}(m-2, m)$ and its dual code, which are documented in Theorem \ref{thm-brmdesign1}. It has also been known for a long time that the codewords of weight $3$ in the Hamming code hold a $2$-$((q^m-1)/(q-1), 3, q-1)$ design. The contribution of Section \ref{sec-hmdesigns} is Theorem \ref{thm-HMdesign171}, which may be viewed as an extension of the known $2$-$((q^m-1)/(q-1), 3, q-1)$ design held in the Hamming code, and also the parameters of the infinite families of $2$-designs derived from the binary Hamming codes, which are documented in Examples \ref{exam-hmdesign4}, \ref{exam-hmdesign5}, \ref{exam-hmdesign6}, and \ref{exam-hmdesign7}. A major contribution of this paper is presented in Section \ref{sec-newdesigns}, where Theorems \ref{thm-newdesigns1} and \ref{thm-newdesigns2} document many infinite families of $2$-design and $3$-designs. The parameters of these $2$-designs and $3$-designs are given specifically. These designs are derived from binary cyclic codes that are defined by special almost perfect nonlinear functions. Another major contribution of this paper is documented in Section \ref{sec-june28}, where Theorem \ref{thm-newdesigns228} and its two corollaries describe several infinite families of $2$-designs. These $2$-designs are related to planar functions. It is noticed that the total number of $3$-designs presented in this paper (see Theorems \ref{thm-brmdesign1} and \ref{thm-newdesigns2}) are exponential. All of them are derived from linear codes. After comparing the list of infinite families of $3$-designs in \cite{KLhb} with the $3$-designs presented in this paper, one may conclude that many, if not most, of the known infinite families of $3$-designs are from coding theory. Section \ref{sec-conjectureddesigns} presents many conjectured infinite families of $2$-designs. The reader is cordially invited to attack these conjectures and solve other open problems presented in this paper.
a8163218a41acf7181d97282cc5bfd3b33b02f79
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Several natural graph classes are known to have polynomially many minimal separators, w.r.t. the number $n$ of vertices of the graph. It is the case for \emph{chordal} graphs, which have at most $n$ minimal separators~\cite{RTL76}, \emph{weakly chordal}, \emph{circular-arc} and \emph{circle} graphs, which have $\mathcal{O}(n^2)$ minimal separators~\cite{BoTo01,KKW98}. The property of having polynomially many minimal separators has been used in algorithms for decades, initially in an ad-hoc manner, i.e., algorithms were based on minimal separators but also other specific features of particular graph classes (see, e.g., ~\cite{BKK93,KKW98}). Later, it was observed that minimal separators are sufficient for solving problems like \textsc{Treewidth} or \textsc{Minimum fill-in}~\cite{BoTo01,BoTo02}. Both problems are related to \emph{minimal triangulations}. Given an arbitrary graph $G$, a minimal triangulation is a minimal chordal supergraph $H$ of $G$, on the same vertex set. Bouchitt\'e and Todinca~\cite{BoTo01} introduced the notion of \emph{potential maximal clique}, that is, a vertex set of $G$ inducing a maximal clique in some minimal triangulation $H$ of $G$. Their algorithm for treewidth is based on dynamic programming over minimal separators and potential maximal clique s. The same authors proved that the number of potential maximal clique s is polynomially bounded in the number of minimal separators~\cite{BoTo02}. Fomin and Villanger~\cite{FoVi10} found a more surprising application of minimal separators and potential maximal clique s, proving that they were sufficient for solving problems like \textsc{Maximum Independent Set}, \textsc{Maximum Induced Forest}, and more generally for finding a maximum induced subgraph $G[F]$ of treewidth at most $t$, where $t$ is a constant. More formally, let $\operatorname{poly}$ be some polynomial. We call $\mathcal{G}_{\poly}$ the family of graphs such that $G \in \mathcal{G}_{\operatorname{poly}}$ if and only if $G$ has at most $\operatorname{poly}(n)$ minimal separators. By~\cite{FoVi10}, the problem of finding a maximum induced subgraph of treewidth at most $t$ can be solved in polynomial time on $\mathcal{G}_{\poly}$. The exponent of the polynomial depends on $\operatorname{poly}$ and on $t$. In~\cite{FTV15}, Fomin \textit{et al.} further extend the technique to compute large induced subgraphs of bounded treewidth, and satisfying some CMSO property (expressible in counting monadic second-order logic). That allows to capture problems like \textsc{Longest induced path}. They also point out some limits of the approach. It is asked in~\cite{FTV15} whether the techniques can be extended for solving the \textsc{Connected Vertex Cover} problem, which is equivalent to finding a maximum independent set $F$ such that $G - F$ is connected. More generally, their algorithm computes an induced subgraph $G[F]$ of treewidth at most $t$ satisfying some CMSO property, but is not able to ensure any property relating the induced subgraph to the initial graph. Here we make some progress in this direction. First, we consider the problem \textsc{Distance-$d$ Independent Set} on $\mathcal{G}_{\poly}$, where the goal is to find a maximum independent set $F$ of the input graph $G$, such that the vertices of $F$ are at pairwise distance at least $d$ in $G$ (in the literature this problem is also known as {\sc $d$-Scattered-Set}). This is equivalent to finding a maximum independent set in graph $G^{d-1}$, the $(d-1)$-th power of $G$. Eto \textit{et al.}~\cite{EGM14} already studied the problem on chordal graphs, and proved that it is polynomial for every even $d$, and NP-hard for any odd $d\geq 3$ (it is even $W[1]$-hard when parameterized by the solution size). Their positive result is based on the observation that for any even $d$, if $G$ is chordal then so is $G^{d-1}$. Eto \textit{et al.}~\cite{EGM14} ask if \textsc{Distance-$d$ Independent Set} is polynomial on chordal bipartite graphs (which are \emph{not} chordal but weakly chordal, see Section~\ref{se:prelim}), a subclass of $\mathcal{G}_{\poly}$. We bring a positive answer to their question for even values $d$, by a result of combinatorial nature: for any graph $G$ and any odd $k$, the graph $G^k$ has no more minimal separators than $G$ (see Section~\ref{se:pow}). Consequently, \textsc{Distance-$d$ Independent Set} is polynomial on $\mathcal{G}_{\poly}$, for any even value $d$ and any polynomial $\operatorname{poly}$, and NP-hard for any odd $d \geq 3$ and any $\operatorname{poly}(n)$ asymptotically larger than $n$. Such a dichotomy between odd and even values also appears when computing large $d$-clubs, that are induced subgraphs of diameter at most $d$~\cite{GHKR14}, and for quite similar reasons. Second, we consider \textsc{Connected Vertex Cover}, \textsc{Connected Feedback Vertex Set} and more generally the problem of finding a maximum induced subgraph $G[F]$ of treewidth at most $t$, such that $G - F$ is connected. We show (Section~\ref{se:cvc}) that the problems are polynomially solvable for subclasses of $\mathcal{G}_{\poly}$, like chordal and circular-arc graphs. This does not settle the complexity of these problems on $\mathcal{G}_{\poly}$. As we shall discuss in Section~\ref{se:ind}, when restricted to bipartite graphs in $\mathcal{G}_{\poly}$, \textsc{Connected Vertex Cover} can be reduced from \textsc{Red-Blue Dominating Set} (see~\cite{DLS14}). It might be that this latter problem is NP-hard on bipartite graphs of $\mathcal{G}_{\poly}$; that was our hope, since the very related problem \textsc{Independent Dominating Set} is NP-hard on chordal bipartite graphs~\cite{DMK90}, and on circle graphs~\cite{BGMPST14}. This question is still open, however we will observe that the \textsc{Red-Blue Dominating Set} is polynomial on the two natural classes of bipartite graphs with polynomially many minimal separators: chordal bipartite and circle bipartite graphs. \section{Preliminaries}\label{se:prelim} Let $G = (V,E)$ be a graph. Let $dist_G(u,v)$ denote the distance between vertices $u$ and $v$ (the minimum number of edges of a $uv$-path). We denote by $N_G^k[v]$ the set of vertices at distance at most $k$ from $v$. Let also $N_G^k(v) = N_G^k[v] \setminus \{v\}$, and we call these sets the \emph{closed} and \emph{open neighborhoods at distance} $k$ of $v$, respectively. Similarly, for a set of vertices $U \subseteq V$, we call the sets $N_G^k(U) = \cup_{u\in U} N_G^k(u) \backslash~ U$ and $N_G^k[U] = \cup_{u\in U} N_G^k[u]$ the open and closed neighborhoods at distance $k$ of $U$, respectively. For $k=1$, we simply denote by $N_G(U)$, respectively $N_G[U]$, the open and closed neighborhoods of $U$; the subscript is omitted if clear from the context. A \emph{clique} (resp. \emph{independent set}) of $G$ is a set of pairwise adjacent (resp. non-adjacent) vertices. A distance-$d$ independent set is a set of vertices at pairwise distance at least $d$. Equivalently, it is an independent set of the $(d-1)$-th power $G^{d-1}$ of $G$. Graph $G^k = (V,E^k)$ is obtained from $G$ by adding an edge between every pair of vertices at distance at most~$k$. Given a vertex subset $C$ of $G$, we denote by $G[C]$ the subgraph induced by $C$. We say that $C$ is a connected component of $G$ if $G[C]$ is connected and $C$ is inclusion-maximal for this property. For $S \subseteq V$, we simply denote $G - S$ the graph $G[V \setminus S]$. We say that $S$ is a \emph{$a,b$-minimal separator} of $G$ if $a$ and $b$ are in distinct components $C$ and $D$ of $G-S$, and $N(C) = N(D) = S$. We also say that $S$ is a minimal separator if it is an $a,b$-minimal separator for some pair of vertices $a$ and $b$. \begin{proposition}[\cite{BBH00}]\label{prop:min_sep} Let $G = (V,E)$ be a graph, $C$ be a connected set of vertices, and let $D$ be a component of $G - N[C]$. Then $N(D)$ is an $a,b$-minimal separator of $G$, for any $a \in C$ and $b \in D$. \end{proposition} \subsection{Graph classes} A graph is \emph{chordal} if it has no induced cycle with more than three vertices. A graph $G$ is \emph{weakly chordal} if $G$ and its complement $\overline{G}$ have no induced cycle with more than four vertices. The classes of \emph{circle} and \emph{circular-arc graphs} are defined by their intersection model. A graph $G$ is a \emph{circle graph} (resp. a \emph{circular-arc graph}) if every vertex of the graph can be associated to a chord (resp. to an arc) of a circle such that two vertices are adjacent in $G$ if and only if the corresponding chords (resp. arcs) intersect. We may assume w.l.o.g. that, in the intersection model, no two chords (resp. no two arcs) share an endpoint. On the circle, we add a \emph{scanpoint} between each two consecutive endpoints of the set of chords (resp. arcs). A \emph{scanline} is a line segment between two scanpoints. Given an intersection model of a circle (resp. circular-arc) graph $G$, for any minimal separator $S$ of $G$ there is a scanline such that the vertices of $S$ correspond exactly to the chords (resp. arcs) intersecting the scanline, see, e.g.,~\cite{KKW98}. Chordal graphs have at most $n$ minimal separators~\cite{RTL76}; weakly chordal, circle and circular-arc graphs all have $\mathcal{O}(n^2)$ minimal separators~\cite{BoTo01,KKW98}. \begin{definition}\label{de:gpoly} Let $\operatorname{poly}$ be some polynomial. We call $\mathcal{G}_{\operatorname{poly}}$ the family of graphs such that $G \in \mathcal{G}_{\operatorname{poly}}$ if and only if $G$ has at most $\operatorname{poly}(n)$ minimal separators, where $n = |V(G)|$. \end{definition} \subsection{Dynamic programming over minimal triangulations}\label{ss:FoVi10} Let $G=(V,E)$ be an arbitrary graph. A chordal supergraph $H = (V, E')$ (i.e., with $E \subseteq E'$), is called a \emph{triangulation} of $G$. If, moreover, $E'$ is inclusion-minimal among all possible triangulations, we say that $H$ is a \emph{minimal triangulation} of $G$. The \emph{treewidth} of a chordal graph is its maximum clique size, minus one. Forests have treewidth $1$, and graphs with no edges have treewidth $0$. The treewidth $\tw(G)$ of an arbitrary graph $G$ is the minimum treewidth over all (minimal) triangulations $H$ of $G$. Cliques of minimal triangulations play a central role in treewidth. A \emph{potential maximal clique} of $G$ is a set of vertices that induces a maximal clique in some minimal triangulation $H$ of $G$. By~\cite{BoTo01}, if $\Omega$ is a potential maximal clique, then for every component $C_i$ of $G - \Omega$, its neighborhood $S_i$ is a minimal separator. Moreover, the sets $S_i$ are exactly the minimal separators of $G$ contained in $\Omega$. \begin{proposition}[\cite{BBC00,BoTo02}] For any polynomial $\operatorname{poly}$, there is a polynomial-time algorithm enumerating the minimal separators and the potential maximal clique s of graphs on $\mathcal{G}_{\poly}$. \end{proposition} Minimal separators and potential maximal clique s have been used for computing treewidth and other parameters related to minimal triangulations, on $\mathcal{G}_{\poly}$. Fomin and Villanger~\cite{FoVi10} extend the techniques to a family of problems: \begin{proposition}[\cite{FoVi10}]\label{pr:FoVi10} For any polynomial $\operatorname{poly}$ and any constant $t$, there is a polynomial algorithm computing a \textsc{Maximum Induced Subgraph of Treewidth at most $t$} on $\mathcal{G}_{\poly}$. \end{proposition} Clearly, \textsc{Maximum Independent Set} (which is equivalent to \textsc{Minimum Vertex Cover}) and \textsc{Maximum Induced Forest} (which is equivalent to \textsc{Minimum Feedback Vertex Set}) fit into this framework: they consist in finding maximum induced subgraphs $G[F]$ of treewidth at most 0, respectively at most 1. The first ingredient of~\cite{FoVi10} is the following observation. \begin{proposition}[\cite{FoVi10}]\label{pr:compat} Let $G = (V,E)$ be a graph, $F \subseteq V$, and let $H_F$ be a minimal triangulation of $G[F]$. There exists a minimal triangulation $H_G$ of $G$ such that $H_G[F] = H_F$. We say that $H_G$ \emph{respects} the minimal triangulation $H_F$ of $G[F]$. \end{proposition} Note that, for any clique $\Omega$ of $H_G$, we have that $F \cap \Omega$ induces a clique in $H_F$. In particular, if $\tw(G[F]) \leq t$ and the clique size of $H_F$ is at most $t+1$, then every maximal clique of $H_G$ intersects $F$ in at most $t+1$ vertices. The second ingredient is a dynamic programming scheme that we describe below. Let $S$ be a minimal separator of $G$, and $C$ be a component of $G - S$ such that $N(C) = S$. The pair $(S,C)$ is called a \emph{block}. Let $\Omega$ be a potential maximal clique\ such that $S \subset \Omega \subseteq S \cup C$. Then $(S,C,\Omega)$ is called a \emph{good triple}. In the sequel, $W$ denotes a set of at most $t+1$ vertices. \begin{definition}\label{de:partcomp} Let $(S,C)$ (resp. $(S,\Omega,C)$) be a block (resp. a good triple) and let $W \subseteq S$ (resp. $W \subseteq \Omega$) be a set of vertices of size at most $t+1$ . We say that a vertex set $F$ is a \emph{partial solution compatible with $(S,C,W)$} (resp. with $(S,C,\Omega,W)$) if: \begin{enumerate} \item $G[F]$ is of treewidth at most $t$, \item $F \subseteq S \cup C$, \item $W = F \cap S$ (resp. $W = F \cap \Omega$), \item there is a minimal triangulation $H$ of $G$ respecting some minimal triangulation of $G[F]$ of treewidth at most $t$, such that $S$ is a minimal separator (resp. $S$ is a minimal separator and $\Omega$ is a maximal clique) of $H$. \end{enumerate} \end{definition} Observe that the two variants of compatibility differ by parameter $\Omega$ and the last two conditions. We denote by $\alpha(S,C,W)$ (resp. $\beta(S,C,\Omega,W)$) the size of a largest partial solution compatible with $(S,C,W)$ (resp. $(S,C,\Omega,W)$). We now show how these quantities can be computed over all blocks and all good triples. The dynamic programming will proceed by increasing size over the blocks $(S,C)$, the size of the block being $|S \cup C|$. It is based on the following equations (see~\cite{FoVi10,FTV15} for details and proofs and Figure~\ref{fi:alphabeta} for an illustration). \paragraph{Base case.} It occurs for good triples $(S,C,\Omega)$ such that $\Omega = S \cup C$. In this case, for each subset $W$ of $\Omega$ of size at most $t+1$, \begin{equation}\label{eq:base} \beta(S,C,\Omega,W) = |W|. \end{equation} \begin{figure}[h] \begin{center} \includegraphics[scale=0.5]{AlphaFromBeta.pdf}\hspace{1cm} \includegraphics[scale=0.5]{BetaFromAlpha.pdf} \end{center} \vspace{-1cm} \caption{Computing $\alpha$ form $\beta$ (left), and $\beta$ from $\alpha$ (right).} \label{fi:alphabeta} \end{figure} \paragraph{Computing $\alpha$ from $\beta$.} The following equation allows to compute the $\alpha$ values from $\beta$ values: \begin{equation}\label{eq:afb} \alpha(S,C,W) = \max_{\Omega, W'} \beta(S,C,\Omega,W'), \end{equation} where the maximum is taken over all potential maximal clique s $\Omega$ such that $(S,C,\Omega)$ is a good triple, and all subsets $W'$ of $\Omega$, of size at most $t+1$, such that $W = W' \cap S$. \paragraph{Computing $\beta$ from $\alpha$.} Let $(S,C,\Omega)$ be a good triple, and fix an order $C_1, C_2,\dots,C_p$ on the connected components of $G[C \setminus \Omega]$. Let $S_i = N_G(C_i)$, for all $1 \leq i \leq p$. By~\cite{BoTo01}, $(S_i,C_i)$ are also blocks of $G$. A partial solution $F$ compatible with $(S,C,\Omega,W)$ is obtained as a union of partial solutions $F_i$ compatible with $(S_i,C_i,W \cap S_i)$, for each $1 \leq i \leq p$, and the set $W$. Denote by $\gamma_i(S,C,\Omega,W)$ the size of the largest partial solution $F$ compatible\footnote{To be precise, the $\gamma$ function is not required at this stage, if we only compute largest induced subgraphs of treewidth at most $t$. However it becomes necessary when we request the solution to satisfy additional properties, as it will happen in Section~\ref{se:cvc}.} with $(S,C,\Omega,W)$, contained in $\Omega \cup C_1 \cup \dots \cup C_i$ (hence $F$ is not allowed to intersect the components $C_{i+1}$ to $C_p$). We have the following equations. \begin{equation}\label{eq:g1} \gamma_1(S,C,\Omega,W) = \alpha(S,C,\Omega,W \cap S_1) + |W| - |W \cap S_1|. \end{equation} For all $i$, $2 \leq i \leq p$, \begin{equation}\label{eq:gi} \gamma_i(S,C,\Omega,W) = \gamma_{i-1}(S,C,\Omega,W) + \alpha(S,C,\Omega,W \cap S_i) - |W \cap S_i|. \end{equation} and finally \begin{equation}\label{eq:bfg} \beta(S,C,\Omega,W) = \gamma_p(S,C,\Omega,W). \end{equation} For convenience we also consider that $\emptyset$ is a minimal separator, and $(\emptyset,V)$ is a block. Then the size of the optimal global solution is simply $\alpha(\emptyset,V,\emptyset)$. The algorithm can be adapted to output an optimal solution, not only its size. \section{Powers of graphs with polynomially many minimal separators}\label{se:pow} Let us prove that for any odd $k$, graph $G^k$ has no more minimal separators than $G$. \begin{theorem}\label{th:main} Consider a graph $G$, an odd number $k=2l+1$ with $l\geq 0$, and a minimal separator $\overline{S}$ of $G^k$. Then there exists a minimal separator $S$ of $G$ such that $\overline{S} = N^{l}_G[S]$. \end{theorem} \begin{proof} The lemma is trivially true if $\overline{S} = \emptyset$. Let $a,b \in V$ such that $\overline{S} \neq \emptyset$ is an $a,b$-minimal separator in $G^k$, and call $C_a$, $C_b$ the components of $G^k - \overline{S}$ that contain $a$ and $b$, respectively. Let us call $D_a = N^{l}_G[C_a]$ and $D_b = N^{l}_G[C_b]$. \smallskip {\bf Claim 1: $dist_G(D_a, D_b) \geq 2$.} Suppose that $dist_G(D_a, D_b) < 2$, and pick $x\in D_a, y \in D_b$ with $dist_G(x,y)\leq 1$ (notice that possibly $x=y$). Let $x_a \in C_a$ and $x_b \in C_b$ be such that there exists an $x_a, x$-path and a $y, x_b$ -path in $G$, each one of length at most $l$, called $P_a$ and $P_b$, respectively. This implies that there must be a $x_a, x_b$-path of length at most $2l + 1 = k$ in $G$, which means that $\{x_a, x_b\} \in E(G^k)$, a contradiction with the fact that $\overline{S}$ separates $C_a$ from $C_b$ in $G^k$. \smallskip {\bf Claim 2}: $\tilde{S} = \overline{S} ~\backslash \left( D_a \cup D_b \right)$ separates $a$ and $b$ in $G$. Notice first that $N_G(D_a) \subseteq N_G^{l+1}(C_a) \subseteq \overline{S}$. Suppose that $\tilde{S}$ does not separate $a$ and $b$, and let $P$ be an $a,b$-path in $G$ that does not pass through $\tilde{S}$. Let $x_1, \dots, x_{s-2}$ the internal nodes of $P$, where $s = |P|$, and consider $i = \max\{j ~|~ x_j \in D_a\cap P\}$. Since $P \cap \tilde{S} = \emptyset$, necessarily $x_{i+1} \in D_b$, a contradiction with Claim 1. \smallskip {\bf Claim 3}: $D_a$ and $D_b$ are connected subsets of $G$. This is straightforward from the definition of the sets, $D_a = N^{l}_G[C_a]$ and $D_b = N^{l}_G[C_b]$, and the fact that $C_a$ and $C_b$ are connected in $G$. \medskip Let $\tilde{C}_b$ be the connected component of $G - N_G[D_a]$ that contains $b$, and denote $S = N_G(\tilde{C_b})$. Note that $S \subseteq \tilde{S} \subset \overline{S}$. By applying Proposition \ref{prop:min_sep}, we have that $S$ is a minimal $a,b$-separator in $G$. Call $\tilde{C}_a$ the component of $G - S$ that contains $a$. Since $S \subseteq \tilde{S}$, we have that $D_b \subseteq \tilde{C}_b$ and $D_a \subseteq \tilde{C}_a$. \smallskip {\bf Claim 4}: $N^l_G[S] = \overline{S}$. We first prove that $N^l_G[S] \subseteq \overline{S}$. By construction, $S \subseteq N_G(D_a)$. Consequently $S\subseteq N_G^{l+1}(C_a) \backslash N_G^{l}(C_a)$, therefore $N_G^l[S] \subseteq N_G^{2l+1}(C_a) = N_{G^k}(C_a) = \overline{S}$. Conversely, we must show that every vertex $x$ of $\overline{S}$ is in $N^l_G[S]$. By contradiction, let $x \in \overline{S} \setminus N^l_G[S]$. We distinguish two cases~: $x \in \tilde{C}_a$, and $x \in \overline{S} \setminus \tilde{C}_a$. In the first case, since $N_{G^k}(C_b) = \overline{S}$, there exists a path from some vertex $y\in C_b$ to $x$ of length at most $k$, in graph $G$. Let us call $P$ one of those $y,x$-paths. Observe that the first $l+1$ vertices of the path belong to $D_b \subseteq \tilde{C_b}$, and none of the last $l+1$ vertices of the path belongs to $S$ (otherwise $x \in N_G^l[S]$). Then $P$ is a path that connects $\tilde{C_b}$ with $\tilde{C_a}$ without passing through $S$, a contradiction with the fact that $S$ separates $a$ and $b$ in graph $G$. It remains to prove the last case, when $x\in \overline{S} \setminus \tilde{C}_a$. Since $N_{G^k}(C_a) = \overline{S}$, there exists a node $y \in C_a$ such that there is a $y,x$-path $P$ of length at most $k$ in $G$. Since the first $l+1$ vertices of the path belong to $D_a$, and the last $l+1$ vertices of the path do not belong to $S$, we deduce that $P$ is an $y,x$-path in $G$ that does not intersect $S$. The path can be extended (through $C_a$) into an $a,x$-path that does not intersect $S$, a contradiction with the fact that $x$ does not belong to $\tilde{C}_a$. This concludes the proof of our theorem. \qed \end{proof} Recall that \textsc{Distance-$d$ Independent set} on $G$ is equivalent to \textsc{Maximum Independent Set} on $G^{d-1}$. Since the latter problem is polynomial on $\mathcal{G}_{\poly}$ by Proposition~\ref{pr:FoVi10}, we deduce: \begin{theorem}\label{th:odddis} For any even value $d$, and any polynomial $\operatorname{poly}$, problem \textsc{Distance-$d$ Independent set} is polynomially solvable on $\mathcal{G}_{\poly}$. \end{theorem} We remind that for any odd value $d$, problem \textsc{Distance-$d$ Independent set} is NP-hard on chordal graphs~\cite{EGM14}, thus on $\mathcal{G}_{\poly}$ for any polynomial $\operatorname{poly}$ asymptotically larger than $n$. The construction of~\cite{EGM14} also shows that even powers of chordal graphs may contain exponentially many minimal separators. \section{On \textsc{Connected Vertex Cover} and \textsc{Connected Feedback Vertex Set}}\label{se:cvc} Let us consider the problem of finding a maximum induced subgraph $G[F]$ such that $\tw(G[F])\leq t$ and $G - F$ is connected. One can easily observe that, for $t=0$ (resp. $t=1$), this problem is equivalent to \textsc{Connected Vertex Cover} (resp. \textsc{Connected Feedback Vertex Set}), in the sense that if $F$ is an optimal solution for the former, than $V(G) - F$ is an optimal solution for the latter. Our goal is to enrich the dynamic programming scheme described in Subsection~\ref{ss:FoVi10} in order to ensure the connectivity of $G - F$. One should think of this dynamic programming scheme of Subsection~\ref{ss:FoVi10} as similar to dynamic programming algorithms for bounded treewidth. The difference is that the bags (here, the potential maximal clique s) are not small but polynomially many, and we parse simultaneously through a set of decompositions. Nevertheless, we can borrow several classical ideas from treewidth-based algorithms. In general, for checking some property for the solution $F$, we add a notion of \emph{characteristics} of partial solutions. Then, for a characteristic $c$, we update the Definition~\ref{de:partcomp} in order to define partial solutions compatible with $(S,C,W,c)$ (resp. $(S,C,\Omega,W,c)$), by requesting the partial solution to be compatible with characteristic $c$. Parameter $c$ will also appear in the updated version of Equations~\ref{eq:base} to~\ref{eq:bfg}. As usual in dynamic programming, the characteristics must satisfy several properties: (1) we must be able to compute the characteristic for the base case, (2) the characteristic of a partial solution $F$ obtained from gluing smaller partial solutions $F_i$ must only depend on the characteristics of $F_i$, and (3) the characteristic of a global solution should indicate whether it is acceptable or not. Moreover, for a polynomial algorithm, we need the set of possible characteristics to be polynomially bounded. For checking connectivity conditions on $G-F$, we define the characteristics of partial solutions in a natural way. Consider a block $(S,C)$ (resp. a good triple $(S,C,\Omega)$) and a subset $W$ of $S$ (resp. of $\Omega$). Let $F$ be a partial solution compatible with $(S,C,W)$ (resp. $(S,C,\Omega,W)$), see Definition~\ref{de:partcomp}. The \emph{characteristic} $c$ of $F$ for $(S,C,W)$ (resp. for $(S,C,\Omega,W)$) is defined as the partition induced on $S \setminus W$ (resp. on $\Omega \setminus W$) by the connected components of $G[S \cup C] - F$. More formally, let $D_1, \dots, D_q$ denote the connected components of $G[S \cup C] - F$, and let $P_j = D_j \cap S$ (resp. $P_j = D_j \cap \Omega$), for all $1 \leq j \leq q$. Then $c = \{P_1,\dots, P_q\}$. We decide that if $S \neq \emptyset$, partial solutions $F$ having some component $D_j$ that does not intersect $S$ (resp. $\Omega$) are immediately rejected; indeed, for any extension $F'$ of $F$, the graph $G - F'$ remains disconnected. Hence we may assume that all sets $P_j$ are non-empty. We say that a partial solution $F$ is \emph{compatible with $(S,C,W,c)$} (resp. \emph{with $(S,C,\Omega,W,c)$}) if it satisfies the conditions of Defintion~\ref{de:partcomp}, and $c$ is the characteristic of $F$ for $(S,C,W)$ (resp. for $(S,C,\Omega,W)$). We also define functions $\alpha(S,C,W,c)$, $\beta(S,C,\Omega,W,c)$ and $\gamma_i(S,C,\Omega,W,c)$ like in Subsection~\ref{ss:FoVi10}, as the maximum size of partial solutions $F$ compatible with the parameters. We can update Equations~\ref{eq:base} to~\ref{eq:bfg} as follows. \paragraph{Base case.} For the good triples $(S,C,\Omega)$ such that $(S,C)$ is inclusion-minimal (hence $\Omega = S \cup C$), \begin{equation}\label{eq:base2} \beta(S,C,\Omega,W,c) = |W|~\text{if $c$ corresponds to the connected components of $G[\Omega \setminus W]$.} \end{equation} Otherwise we set $\beta(S,C,\Omega,W,c) = -\infty$. \paragraph{Computing $\alpha$ from $\beta$.} \begin{equation}\label{eq:afb2} \alpha(S,C,W,c) = \max_{\Omega, W',c'} \beta(S,C,\Omega,W',c'), \end{equation} where the maximum is taken over all potential maximal clique s $\Omega$ such that $(S,C,\Omega)$ is a good triple, and all subsets $W'$ of $\Omega$, of size at most $t+1$, such that $W = W' \cap S$, and all characteristics $c'$ such that each part of $c$ corresponds to the intersection between $S$ and a part of $c'$. If $S \neq \emptyset$ we also request that each part of $c'$ intersects $S$. This condition allows to reject partial solutions $F$ for which a component of $G[S \cup C] - F$ is strictly contained in $C$. Indeed, such partial solutions cannot extend to global solutions $F'$ such that $G - F'$ is connected. For the particular case $S = \emptyset$ (hence $C=V$ and $W = \emptyset$) we only consider characteristics $c'$ with a single part. This ensures that the global solution $F$ satisfies that $G - F$ is connected. If there is no such triple $(\Omega, W',c')$, then we set $\alpha(S,C,W,c) = - \infty$ (we can assume that, when it has no parameters, function $\max$ returns $-\infty$). \paragraph{Computing $\beta$ from $\alpha$.} \begin{equation}\label{eq:g12} \gamma_1(S,C,\Omega,W,c) = \max_{c'} (\alpha(S_1,C_1,\Omega,W \cap S_1, c') + |W| - |W \cap S_1|), \end{equation} over all characteristics if $c'$ that \emph{map correctly} on $c$, in the following sense. Consider a characteristic $c'$ and let $G_{c'}[\Omega \setminus W]$ be the graph obtained from $G[\Omega \setminus W]$ by completing each part $D \in c'$ into a clique. We say that a characteristic $c'$ maps correctly on $c$ if $c$ is the partition of $\Omega \setminus W$ defined by the connected components of $G_{c'}[\Omega \setminus W]$. The notion of mapping transforms the characteristic of the partial solution $F_1$ w.r.t. $(S_1,C_1,W \cap S_1)$ into the characteristic of $F_1 \cup W$ w.r.t. the quadruple $(S,C,\Omega,W)$. For all $i$, $2 \leq i \leq p$, \begin{equation}\label{eq:gi2} \gamma_i(S,C,\Omega,W,c) = \max_{c_{i-1},c_i}(\gamma_{i-1}(S,C,\Omega,W,c_{i-1}) + \alpha(S,C,\Omega,W \cap S_i,c_i) - |W \cap S_i|), \end{equation} over all pairs of characteristics $c_{i-1},c_i$ that \emph{map correctly} on $c$. That is, $c$ must correspond to the connected components of $G_{c_{i-1},c_i}[\Omega \setminus W]$, obtained from $G[\Omega \setminus W]$ by completing each part of $c_{i-1}$ and each part $c_i$ of into a clique. Finally \begin{equation}\label{eq:bfg2} \beta(S,C,\Omega,W,c) = \gamma_p(S,C,\Omega,W,c). \end{equation} The optimal solution size is, as before, $\alpha(\emptyset,V,\emptyset,\{\emptyset\})$. In general, the number of characteristics may be exponential. Nevertheless, there are classes of graphs with the property that each minimal separator $S$ and each potential maximal clique\ $\Omega$ can be partitioned into at most a constant number of cliques. With this constraint, the number of characteristics is polynomial (even constant, for any given triple $(S,C,W)$ or quadruple $(S,C,\Omega,W)$). This is the case for chordal graphs, where each minimal separator and each potential maximal clique\ induces a clique in $G$. It is also the case for circular-arc graphs. Recall that each minimal separator corresponds to the set of arcs intersecting a pair of scanpoints~\cite{KKW98}. Moreover, by~\cite{KKW98,BoTo01}, each potential maximal clique\ corresponds to the set of arcs intersecting a triple of scanpoints. Since arcs intersecting a given scanpoint form a clique, we have that each minimal separator can be partitioned into two cliques, and each potential maximal clique\ can be partitioned into three cliques. We deduce: \begin{theorem}\label{th:cvc} On chordal and circular-arc graphs, problems \textsc{Connected Vertex Cover} and~\textsc{Connected Feedback Vertex Set} are solvable in polynomial time. More generally, one can compute in polynomial time a maximum vertex subset $F$ such that $G[F]$ is of treewidth at most $t$ and $G - F$ is connected. \end{theorem} Note that Escoffier \textit{et al.}~\cite{EGM10} already observed that \textsc{Connected Vertex Cover} is polynomial for chordal graphs. \section{\textsc{Independent Dominating Set} and variants}\label{se:ind} The \textsc{Independent Dominating Set} problem consists in finding a \emph{minimum} independent set $F$ of $G$ such that $F$ dominates $G$. Hence the solution $F$ induces a graph of treewidth $0$ and it is natural to ask if similar techniques work in this case. The fact that we have a minimization problem is not a difficulty: the general dynamic programming scheme applies in this case, and for any weighted problem with polynomially bounded weights, including negative ones~\cite{FoVi10,FTV15}. \textsc{Independent Dominating Set} is known to be NP-complete in chordal bipartite graphs~\cite{DMK90} and in circle graphs~\cite{BGMPST14}. Therefore, it is NP-hard on $\mathcal{G}_{\poly}$ for some polynomials $\operatorname{poly}$. But, again, we can use our scheme in the case of circular-arc graphs, for this problem or any problem of the type minimum dominating induced subgraph of treewidth at most a constant~$t$. Let $(S,C)$ be a block an let $F \subseteq S \cup C$ be a partial solution compatible with $(S,C,W)$ for some $W \subseteq S$ of size at most $t+1$ (in the sense of Definition~\ref{de:partcomp}). The natural way for defining the characteristic of $F$ is to specify which vertices of $S$ are dominated by $F$ and which are not (we already know that $F \cap S = W$). It is thus enough to memorize which vertices of $S$ are dominated by $F \cap C$. In circular-arc graphs, this information can be encoded using a polynomial number of characteristics. Indeed, a minimal separator $S$ corresponds to arcs intersecting a scanline, between two scanpoints $p_1$ and $p_2$ of some intersection model of $G$. Moreover (see~\cite{KKW98}), the vertices of component $C$ correspond to the arcs situated on one of the sides of the scanline. Let $s^1_1,s^1_2,\dots, s^1_{l_1}$ be the arcs of the model containing scanpoint $p_1$, ordered by increasing intersection with the side of $p_1p_2$ corresponding to $C$. Simply observe that if $F \cap C$ dominates vertex $s^1_i$, it also dominates all vertices $s^1_j$ with $j>i$. Therefore we only have to store the vertex $s^1_{min_1}$ dominated by $F \cap C$ which has a minimum intersection with the side of the scanline corresponding to component $C$, and proceed similary for the arcs of $S$ containing scanpoint $p_2$. These two vertices of $S$ will define the characteristic of $F$, and they suffice to identify all vertices of $S$ dominated by $F \cap C$. \begin{figure} \centering \begin{tikzpicture}[scale=1] \draw(-1,0) arc (0:360:2) ; \draw[ dotted] (-0.66,0) arc (0:10:2.33); \draw[-|] (-0.66,0) arc (0:-30:2.33); \draw[ dotted] (-0.33,0) arc (0:10:2.66); \draw[-|] (-0.33,0) arc (0:-45:2.66) ; \draw[dotted] (0,0) arc (0:10:3); \draw[-|] (0,0) arc (0:-60:3); \draw[-|] (-6,0) arc (180:230:3) ; \draw[dotted] (-6,0) arc (180:170:3) ; \draw[-|] (-5.66,0) arc (180:200:2.66) ; \draw[dotted] (-5.66,0) arc (180:170:2.66) ; \draw[densely dotted, thick, |-|] (-4.8,-1.5) arc (220:310:2.33); \node (p1) at (-0.75,-0.3){$\phantom{p_1}$}; \node (p2) at (-5.25,-0.3){$\phantom{p_2}$}; \draw[-,very thick] (p1)--(p2); \node (pp1) at (-1.2,-0.1){$p_1$}; \node (pp2) at (-4.75,-0.1){$p_2$}; \node (C) at (-3.3,-2.6){$F \cap C$}; \node (pp2) at (-1.1,-1.35){$s_1^1$}; \node (pp2) at (-1.2,-2.1){$s_2^1$}; \node (pp2) at (-1.7,-2.6){$s_3^1$}; \node (pp2) at (-5.4,-1.1){$s_1^2$}; \node (pp2) at (-4.75,-2.4){$s_2^2$}; \end{tikzpicture} \caption{Domination in circular-arc graphs. The characteristic of $F$ w.r.t. $(S,C)$ is $(s_3^1, s_2^2)$.}\label{fi:ca} \end{figure} These characteristics can be used to compute a minimum dominating induced subgraph of treewidth at most $t$, for circular-arc graphs, in polynomial time. We will not show, in details, how to do it, since the technique is quite classical. Problem \textsc{Independent Dominating Set} is already known to be polynomial for this class~\cite{Ch98,Va12}. The algorithm of Vatshelle~\cite{Va12} is more general, based on parameters called \emph{boolean-width} and \emph{MIM-width}, which are small ($\mathcal{O}(\log n)$ for the former, constant for the latter) on circular-arc graphs and also other graph classes. Another problem of similar flavor, combining domination and independence, is \textsc{Red-Blue Dominating Set}. In this problem we are given a bipartite graph $G=(R,B,E)$ with red and blue vertices, and an integer $k$, and the goal is to find a set of at most $k$ blue vertices dominating all the red ones. \textsc{Red-Blue Dominating Set} can be reduced to \textsc{Connected Vertex Cover} as follows~\cite{DLS14}. Let $G'$ be the graph obtained from $G=(R,B,E)$ by adding a new vertex $u$ adjacent to all vertices of $B$ and then, for each $v \in R \cup \{u\}$, a pendant vertex $v'$ adjacent only to $v$. Then $G$ has a red-blue dominating set of size at most $k$ if and only if $G'$ has a connected vertex cover of size at most $k+|B|+1$. Indeed any minimum connected vertex cover of $G'$ must contain $u$, $R$, and a subset of $B$ dominating $R$. It is not hard to prove that this reduction increases the number of minimal separators by at most $\mathcal{O}(n)$, see Appendix~\ref{app:red}. Therefore, if \textsc{Red-Blue Dominating Set} is NP-hard on (bipartite) $\mathcal{G}_{\poly}$ for some $poly$, so is \textsc{Connected Vertex Cover}. There are two natural, well-studied classes of bipartite graphs with polynomial number of minimal separators, and it turns out that \textsc{Red-Blue Dominating Set} is polynomial for both. One is the class of chordal bipartite graphs (which are actually defined as the bipartite, \emph{weakly} chordal graphs). For this class, \textsc{Red-Blue Dominating Set} is polynomial by~\cite{DMK90}. Reference~\cite{DMK90} considers the total domination problem for the class, but the approach is based on red-blue domination. The second natural class is the class of circle bipartite graphs, i.e., bipartite graphs that are also circle graphs. They have an elegant characterization established by de Fraysseix~\cite{Fr81}. Let $H = (V,E)$ be a planar multigraph, and partition its edge set into two parts $E_R$ and $E_B$ such that $T=(V,E_R)$ is a spanning tree of $H$. Let $B(H, E_R) = (E_R, E_B, E')$ be the bipartite graph defined as follows: $E_R$ is the set of red vertices, $E_B$ is the set of blue vertices, and $e_R \in E_R$ is adjacent to $e_B \in E_B$ if the unique cycle obtained from the spanning tree $T$ by adding $e_B$ contains the edge $e_R$. We say that $B(H, E_R)$ is a fundamental graph of $H$. By~\cite{Fr81}, a graph is circle bipartite if and only if it is the fundamental graph $B(H, E_R)$ of a planar multigraph $H$. Consider now the \textsc{Tree augmentation} problem that consists in finding, on input $G$ and a spanning tree $T$ of $G$, a minimum set of edges $D \subseteq E(G)-E(T)$ such that each edge in $E(T)$ is contained in at least one cycle of $G' = (V, E(T) \cup D)$. In \cite{Provan199987} is shown that \textsc{Tree augmentation} is polynomial when the input graph is planar. Is direct to see that a set $S \subseteq E_B$ is a solution of the \textsc{Tree augmentation} problem on input $H = (V, E_R \cup E_B)$ and $T = (V, E_R)$, if and only if $S$ is a solution of \textsc{Red-Blue Dominating Set} on input $B(H) = (E_R, E_B, E')$. This observation, together with \cite{Fr81} and \cite{Provan199987}, impliy that \textsc{Red-Blue Dominating Set} is polynomial in circle bipartite graphs. \section{Discussion} We showed how the dynamic programming scheme of~\cite{FoVi10,FTV15} can be extended for other optimization problems, on \emph{subclasses} of $\mathcal{G}_{\poly}$. Note that the algorithm of~\cite{FTV15} allows to find in polynomial time, on $\mathcal{G}_{\poly}$, a maximum (weight) subgraph $G[F]$ of treewidth at most $t$, satisfying some property expressible in CMSO. It also handles annotated versions, where the vertices/edges of $G[F]$ must be selected from a prescribed set. We have seen that \textsc{Distance-$d$ Independent Set} can be solved in polynomial time on $\mathcal{G}_{\poly}$ for any even $d$. This also holds for the more general problem of finding an induced subgraph $G[F]$ whose components are at pairwise distance at least $d$, and such that each component is isomorphic to a graph in a fixed family. E.g., each component could be an edge, to have a variant of \textsc{Maximum Induced Matching} where edges should be at pairwise distance at least $d$. For this we need to solve the corresponding problem on $G^{d-1}$, using only edges from $G$, as in~\cite{FTV15}. When seeking for maximum (resp. minimum) induced subgraphs $G[F]$ of treewidth at most $t$ such that $G - F$ is connected (resp. $F$ dominates $G$) on particular subclasses of $\mathcal{G}_{\poly}$, we can add any CMSO condition on $G[F]$. It is not unlikely that the techniques can be extended to other classes than circular-arc graphs (and chordal graphs, for connectivity constraints). We also believe that the interplay between graphs of bounded MIM-width~\cite{Va12} and $\mathcal{G}_{\poly}$ deserves to be studied. None of the classes contains the other, but several natural graph classes are in their intersection, and they are both somehow related to induced matchings. We leave as open problems the complexity of \textsc{Connected Vertex Cover} and \textsc{Connected Feedback Vertex set} in weakly chordal graphs, and on $\mathcal{G}_{\poly}$. We have examples showing that, even for weakly chordal graphs, the natural set of characteristics that we used in Section~\ref{se:cvc} is not polynomially bounded. \paragraph{Acknowledgements.} We thank Iyad Kanj for fruitful discussions on the subject. \bibliographystyle{plain}
203e9377263667eea402357d4f1fae62c27e78c9
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Dust grains are the raw material for planetesimals, which over time form rocky planet cores. The widely accepted core accretion scenario implies that coagulation of micron-sized dust grains into larger bodies is a first step for planet formation. Thus, it is important to investigate the grain size and its distribution in the birthplaces of planets: protoplanetary disks. If the emission is optically thin, the opacity index ($\beta$) can be used as an observational indication of grain size (possibly grain growth), where $\kappa_d$ $(\propto \lambda^{-\beta})$ is the dust opacity and $\lambda$ is the wavelength. A number of observations reveal that the value of $\beta$ is generally low for circumstellar disks around low and intermediate-mass young stellar objects (YSOs) \citep{beckwith91,dalessio01,andrews05,draine06,ricci10a,ricci10b,guilloteau11,menu14}. The exploration of radial and azimuthal dust distribution in various sizes of dust is essential for revealing where and how dust grain coagulation takes place. Since the $\beta$ offers suggestive evidence for the amount of dust grains whose size is approximately comparable to the wavelength of observation, the analysis of $\beta$ and its distribution help understand the initial condition for the subsequent formation of planetesimals. We have conducted $H$-band linear polarimetric observations that trace (sub)micron-sized dust at the disk scattering surface and 0.87 mm dust continuum observations that detect the thermal emission from approximately millimeter-sized grains around the disk mid-plane. LkH$\alpha$ 330 is a G3 star \citep{cohen79} of 2.5 M$_\sun$ \citep{salyk09} and 15 L$_\sun$ \citep{andrews11} located in the Perseus molecular clouds, which is about 250 pc away from the Sun \citep{enoch06}. The disk of LkH$\alpha$ 330 is one of the 4 transitional disks found in the Spitzer Space Telescope Cores to Disks (c2d) Legacy Program, where they took $5-35$ $\mu$m spectra with the Infrared Spectrograph (IRS) for 100 stellar systems \citep{brown07}. Later, \citet{brown08,brown09} constructed the first resolved submillimeter image of the disk with an angular resolution of 0\farcs3 using the Submillimeter Array (SMA) at 340 GHz. This observation directly confirmed the cavity in the dust disk. Then, a two-dimensional Monte Carlo radiative transfer code \citep[RADMC;][]{dullemond04} was used to model the disk. In this model, the disk is flared, density is a power law of radius, and the cavity is represented by the inner and outer gap radius and a density reduction factor. In addition to the best-fitting model of the spectral energy distribution (SED), they indicated that the radius of the cavity was 47 AU and that hot gas resided within 0.8 AU. \citet{andrews11} re-calibrated the data from \citet{brown08,brown09} and combined it with SMA compact (COM) data. They treated the surface density by the similarity solution with an exponential tail in their analysis and derived the cavity radius to be 68 AU. \citet{isella13} also studied the asymmetric component in the LkH$\alpha$ 330 disk with SMA archival data and CARMA data at $\lambda = 1.3$ mm. As results of the visibility fitting with parametric gaussian disk models, they found that the asymmetric component traces a narrow circular arc, which extends in the azimuthal direction by about 90$\arcdeg$ and accounts for about 1/3 of the total disk flux at $\lambda = 1.3$ mm. They discuss possible mechanisms of the arc formation including planets, Rossby waves instabilities (RWI), baroclinic instabilities, disk warping, and disk shadowing. Observations of molecular tracers might help to further confirm the origin of the cavity in the disk. In this paper, the details of the $H$-band linear polarimetric observations and 0.87 mm dust continuum observations of LkH$\alpha$ 330 are provided in section \ref{obs}, and the obtained images which show a spiral arm and asymmetric structures in the LkH$\alpha$ 330 disk are presented in section \ref{result}. In section \ref{discussion}, micron and millimeter-sized dust distributions and the origin of the spiral arm as well as the asymmetric structure are discussed. Finally, we examine the possibility of dust trapping by estimating the dust opacity index using our observations and the data from the literature. Section \ref{summary} summarizes our findings. \section{Observations and Data Reduction} \label{obs} \subsection{Subaru/HiCIAO $H$-band Observation} \label{subaru} $H$-band (1.6 $\mu$m) linear polarimetric observations of LkH$\alpha$ 330 were conducted on December 22nd 2011 (UT) as a part of the Strategic Explorations of Exoplanets and Disks with Subaru \citep[SEEDS;][]{tamura09} project. The High Contrast Instrument for the Subaru Next Generation Adaptive Optics \citep[HiCIAO;][]{tamura06} with a dual-beam polarimeter on the Subaru 8.2 m telescope was used. The polarimetric observation mode acquires the \textit{o}-ray and \textit{e}-ray simultaneously, and images a 10\arcsec $\times$ 20\arcsec field of view (FOV) with a pixel scale of 9.50 mas/pixel. The adaptive optics system \citep[AO188;][]{hayano08} functioned mostly stable during observation, providing stellar point-spread function (PSF) with a FWHM of 0\farcs15 in the \textit{H}-band. The adaptive optics corrections functioned well in the observation and achieve a saturated radius (inner working angle) of 0\farcs2, corresponding to $\sim$ 50 AU. As discussed in Section \ref{result}, the region less than 0\farcs2 from the central star in principle shows no saturation problems nor discernible misalignment of polarization vectors and thus could potentially be usable. However, we did not include the region within $r = 0\farcs2$, and to be conservative, polarization measurements at radii more distant than 0\farcs2 from the central star are regarded as more accurate and safe analysis. The data sets were taken with a combination of Angular Differential Imaging (ADI) and Polarimetric Differential Imaging (PDI) mode with a net field rotation of 12 degrees. A double Wollaston prism system was used for splitting the incident light, and as a result, four images with individual FOV of 5\arcsec $\times$ 5\arcsec are generated. The total integration time is 1600 s for the final PI image. All reduced data were corrected for field rotation and then integrated. Note that only data obtained by PDI mode were used and ADI analysis leaves for our future work. A standard reduction method \citep{hinkley09} was applied to the polarimetric data, including corrections for distortion and flat-field, using the Image Reduction and Analysis Facility (IRAF\footnote{IRAF is distributed by National Optical Astronomy Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.}) software as in previous SEEDS publications \citep[e.g.][]{hashimoto11}. Calibration for instrumental polarization of the Nasmyth HiCIAO instrument was made based on the Mueller-matrices technique, which quantizes polarization characteristics \citep{joos08}. Finally, a PI image was calculated from $\sqrt{Q^2+U^2}$, where $Q$ and $U$ are orthogonal linear polarizations. \subsection{Submillimeter Array Observation} \label{sma} LkH$\alpha$ 330 was observed at 343.796 GHz, corresponding to 0.87 mm in wavelength, by the SMA on December 18th 2014 (UT). The very extended configuration was selected, with eight 6 m diameter antennae whose baseline lengths ranged between 30 m and 590 m. The double sideband (DSB) receivers with a total bandwidth of 4 GHz was used to maximize the sensitivity to the continuum. The bright quasar 3c279 was used to calibrate the bandpass and Callisto was observed for determining an absolute flux scale, whose uncertainty is estimated to be approximately 10$\%$. The nearby quasar 3c84 was also observed to calibrate the atmospheric and instrumental phase before and after observing the target source. The system temperature was between 400 K and 600 K and the precipitable water vapor (PWV) was less than 2.5 mm during the observation. The integration time on source was 7.4 hours. All of the calibrations were done using the MIR software package\footnote{The package is open for public use and can be obtained from https://www.cfa.harvard.edu/~cqi/mircook.html.}. The deconvolution was done using the MIRIAD software and the natural weighting was applied when we CLEANed the image so that we can obtain higher sensitivity. Self-calibration was also introduced to the imaging by applying the continuum image itself as a model. \section{Results} \label{result} The polarized intensity image in $H$-band of the transitional disk around LkH$\alpha$ 330 is presented in Figure \ref{fig:fig1}a. The figure shows two bright components at approximately 70 AU in the east and west side of the disk, whose intensities are 6.1 and 3.6 mJy arcsec$^{-2}$, respectively. The rms is only 0.16 mJy arcsec$^{-2}$, producing S/N ratios of 38 and 22, respectively. The western bright component includes a spiral arm structure extending to the southeast direction, and the bright component in the eastern side also seems to include a spiral arm-like structure, although its significance is less clear than the other side. Figure \ref{fig:fig1}b displays the continuum image at $\lambda$ = 0.87 mm obtained from the SMA. The synthesized beam of $0\farcs31 \times 0\farcs27$ was obtained as shown in the lower left corner in the same Figure. The rms in this image is 1.67 mJy beam$^{-1}$. The emission extends 0\farcs6 in radius, corresponding to 150 AU, and a surface brightness depletion is seen around the central star. As in the $H$-band polarized image shown in Figure \ref{fig:fig1}a, asymmetric structure can be seen at the region of $r \la 1\arcsec$ from the central star. The continuum peak in the west side is 29.6 $\pm$ 1.6 mJy beam$^{-1}$, while the one in the east side is 24.8 $\pm$ 1.6 mJy beam$^{-1}$. The continuum image superimposed over the $H$-band polarized image in a larger view is provided in Figure \ref{fig:fig1}c. Both of the emissions are similarly distributed, and their peak intensities spatially match with each other. Since the Strehl ratio provided by the AO was modest (approximately 0.2), the seeing PSF was not perfectly corrected for, and thus a stellar halo partially remains in the image. The polarization by forward and backward scattering in the disk becomes smaller because the polarization direction significantly deviates from 90$\arcdeg$ of the scattering angle, making the halo appear more polarized along the minor axis of the disk. Figure \ref{fig:fig2} displays how the polarized halo affects the PI image. Figure \ref{fig:fig2}a shows the original PI image with no polarized halo correction. The polarization vectors align along the minor axis due to the polarized halo. However, the polarization vectors become circularly aligned when the remaining polarized halo is subtracted from the PI image as shown in Figure \ref{fig:fig2}b. The correction is made by measuring the Stokes parameters, I, Q, and U at the region between 0\farcs2375 and 2\farcs85 (corresponding to 25 to 300 pixels). The uniformly distributed polarized halo, which has a polarization degree of $0.51 \pm 0.01$ and a polarization angle of $172.7 \pm 0.5\degr$, is subtracted from the image in Figure \ref{fig:fig2}a. The correction method is described in detail by \citet{hashimoto12}. Note that the region beyond $r = 0\farcs2$ can be safely analyzed after the correction, since the PI of the disk is now distinguished from the polarized halo and no misalignment of polarization vectors can be seen beyond $r = 0\farcs2$. We further checked the effect of instrumental polarization by the radial polarization (${\rm U}_\phi$) image provided in the Appendix. The polarization component that is not originated from the source is approximately 10 times smaller than the component polarized by the disk source. As a result, the effect of the instrumental polarization component is negligible in our results and therefore not affecting the discussed structures significantly. See more details in the Appendix. \section{Discussion} \label{discussion} \subsection{Dust distribution and implications of a massive planet} \label{dist} Figure \ref{fig:fig3}a shows the radial surface brightness profiles along the directions of P. A. = 94$\arcdeg$ and 235$\arcdeg$ crossing the scattered emission peaks and P. A. = 258$\arcdeg$ crossing the continuum peak in the radial direction as denoted by red lines in Figure \ref{fig:fig1}c. Both of the scattering emission peaks are found at $r \sim 0\farcs25$ ($\sim$ 62.5 AU) from the central star indicated by dashed circle in Figure \ref{fig:fig1}c. Note that the polarized emission peaks at P. A. = 94$\arcdeg$ and 235$\arcdeg$ are almost located along the direction of semi-major axis (the position angle of LkH$\alpha$ 330 is 80$\arcdeg$ \citep{andrews11} measured from the north in counter-clockwise). While surface brightness asymmetries along the semi-minor axis can be strongly affected by anisotropic scattering, asymmetries along the semi-major axis are unaffected by the scattering efficiency and thus reflect intrinsic variations in the disk structure. Thus, the disk inclination of 35$\arcdeg$ \citep{isella13} does not affect the results very much. Figure \ref{fig:fig3}b represents the azimuthal surface brightness profile at $r = 0\farcs25$ and shows a non-axisymmetric torus in the disk. The surface brightnesses in the west and east sides are approximately 3 and 6 times higher than the base levels found in the north and south regions. As shown in Figure \ref{fig:fig1}c, the two continuum peaks at the east and west sides of the disk match with the peaks seen in the $H$-band PI image, indicating that small-sized dust (0.1 - 1 $\mu$m) has a similar distribution to the millimeter sized dust when phase function including polarization property is the same everywhere. The radial profiles of scattering and continuum emissions also support that they similarly distribute in the disk because their own peaks are seen at the same radial distance of 0$\farcs25$ as shown in Figure \ref{fig:fig3}a. It is important to know whether the disk is optically thick or not at $\lambda = 0.87$ mm continuum when they are compared with other emissions at different wavelength. It can be roughly checked by comparing the brightness temperature ($T_{\rm b}$) with the radiative equilibrium temperature. As a result, $T_{\rm b}$ at the continuum peak was 9.6 K, which is much lower than the radiative equilibrium temperature of 98 K at $r = 0\farcs25$, indicating that the continuum emission is optically thin. Our results suggest that dust grains between (sub)micron and millimeter in size accumulate at roughly the same radius, but not necessarily from an annular structure of equal thickness. Both the $H$-band PI image and the 0.87 mm continuum image shown in Figure \ref{fig:fig1}c indicate that there is a dust concentration in the disk. Recent theoretical studies have proposed several mechanisms to trap dust particles: snow lines, dead zones generated by magneto-rotational instability \citep[MRI;][]{balbus91,dzyurkevich10}, anticyclonic vortices that generate high pressure region triggered by RWI \citep{meheut12}, and pressure bumps generated by massive planets \citep{{paardekooper04},zhu12,pinilla12a,pinilla12b,pinilla15}. Among such possibilities, LkH$\alpha$ 330 implies the possible presence of a massive planet because at least spiral structure is associated in the disk. Spiral arms have been discovered in other transitional disks, such as SAO 206462, MWC 758, and HD 100453 and theoretical simulations reproduce the observed spiral features by embedded planet(s), indicating that spiral arms can be induced by planet-disk interaction \citep{muto12,garufi13,grady13,benisty15,dong15,dong16,wagner15}. A pair of arm-like features in the east and west sides of the LkH$\alpha$ 330 disk, extending from a continuum peak toward the southeast and northwest, may signify the presence of a massive planet probably located in the south east or north west direction as an extension of the spiral arms and which can generate a pressure bump that locally accumulates a large amount of dust. \citet{dong15} produced the $H$-band polarized intensity images by the newly developed Athena $+ +$ MHD code (J. M. Stone et al. 2016, in preparation) showed spiral structure with $m = 2$ mode (see Figure 2 in \citet{dong15}) by a single massive planet at outer disk region. They also predicts that the brightness of the observed arm is expected to be enhanced by 100 to 300 \% from the emission free background, which is consistent with our result of 380 \% enhancement. \citet{isella13} showed that the interaction between the disk and planet is a possible mechanism by utilizing hydrodynamic simulation code FARGO \citep{masset00} as a means to investigate the origin of the asymmetric disk structure. They explored disk structure during 100 to 2000 disk rotations with viscosity of 0.002 and 0.02 and the number of planets with different mass between 3 and 10 $M_J$, and could reproduce the observed azimuthal asymmetries at $\lambda$ = 1.3 mm emission. \citet{birnstiel13} also demonstrated the formation of azimuthal asymmetries in the surface density by disk-planet interactions during their orbital rotation. Therefore, large grains are probably trapped or piled up by a pressure bump generated by a massive planet that depletes gas surface density significantly while interacting with the gas disk \citep{pinilla12a, pinilla12b}. Small grains, on the other hand, trace the gas features because they are mostly coupled to the gas. They further study that such asymmetry possibly leads to azimuthal variations in dust opacity at the millimeter wavelength because large dust grains are trapped at pressure maxima due to the local gas density enhancement. They also argue that the variation of dust temperature in the azimuthal direction might be responsible for generating azimuthal asymmetry. The incident angle of the stellar radiation to the disk is not uniform if LkH$\alpha$ 330 disk is warped such as HD 142527 disk \citep{casassus15a,marino15}, the disk is differently flared in the azimuthal direction, or the shadowing variation behind the asymmetric outer gap wall in the inner disk, causing azimuthal variation in the dust temperature. Therefore, the $H$-band emission depends on internal disk structure, while sub-millimeter is not. All of the possibilities mentioned above can explain the origin of the anti-correlation observed between $H$-band and 0.87 mm continuum emissions. The key questions whether or not a massive planet can form the spirals in scattered light and the sub-millimeter asymmetry will arise. \citet{bae16} simultaneously reproduced the observed spiral arms in NIR and asymmetry in sub-millimeter continuum by a single planet-mass companion at the outer disk region in the case of SAO 206462 transitional disk, and thus their works support in explaining the LkHa330 structure similar to SAO 206462. If this is the case, a massive planet orbiting at the outer disk could be a more plausible explanation for current architecture seen in the LkHa330. They also predicted that brightest peak and the spiral arms will show observable displacement over the next few years. The follow-up observations in NIR and sub-millimeter continuum are demanded for revealing the origin of the structure. Details of the arm structures will be the subject of later studies in this series. Here, we simply stress that LkH$\alpha$ 330 has a spiral arm structure, a supportive evidence for the presence of a massive planet. In the next section, we discuss the possibility of grain growth due to a planet-induced pressure bump. \subsection{Implications of grain growth due to pressure bump} \label{growth} Grain growth can be found observationally from the slope of the SED, expressed as $F_\nu \propto \nu^\alpha$ where $\alpha = \beta + 2$, at mm/sub-mm wavelengths. We thus use this law to estimate the dust opacity in the LkH$\alpha$ 330 disk. $\beta$ is related to composition, shape, and size distribution of the dust that accounts for thermal emission at mm/sub-mm wavelengths. Generally, $\beta$ lies between 0 and 1 for disks around low-mass T Tauri stars and almost every disk falls in the range of $\beta < 2$ \citep{beckwith91}. Note that the $\beta$ of the interstellar medium (ISM) is about 1.7 \citep{li01}. It is widely accepted that the explanation for $\beta_{\rm disk} < \beta_{\rm ISM}$ is primarily a grain growth effect in the disk, and other contributing factors such as the geometry of dust particles or an optically thick disk are less effective and don't alter $\beta$ to the same extent \citep{draine06}. \citet{draine06} also shows that the dust grows from $\mu$m to mm sized grains when $\beta$ decreases below unity. Here we estimate $\beta$ with our 0.87 mm continuum observation and a previous result from a 1.3 mm continuum image taken by CARMA \citep{isella13}. The integrated intensity in the $\lambda = 0.87$ mm and $1.3$ mm continuum observations in west side of the disk is 101 $\pm$ 14.1 and 35 $\pm$ 1 mJy, respectively, but 87 $\pm$ 12.7 and 18 $\pm$ 1 mJy in the east side of the disk, respectively. As a result, the estimated $\beta$ becomes 0.7 $^{+0.5}_{-0.4}$ in the west side and 2.0 $\pm$ 0.5 in the east side, respectively. $\beta$ in the west side is below unity and indicates that grain growth probably takes place as suggested from the work of \citet{draine06}. It is theoretically inferred that the dust grains grow to around 500 $\mu$m or even larger, because such sizes are required for $\beta$ to be less than unity for a power-law distribution of grain sizes ($a$), $n(a) \propto a^{-p}$, where n(a) represents the abundance of grains with a particular size and $p$ is a power index of the distribution \citep{dalessio01,natta04a,natta04b}. $\beta$ in the east side, on the other hand, shows a typical number of $\sim$ 2, indicating that grain growth is not so active than the west side. Note that \citet{dalessio01} also show that $\beta$ will become approximately 1.7 when the maximum size of dust grains is less than 30 $\mu$m. A pressure bump is considered as one possible explanation to promote grain growth since it locally traps dust material in the azimuthal direction and as a result accelerates dust particles growth \citep{barge95,birnstiel13}. If this is the case, the larger grains subsequently stay near the pressure bump while the smaller grains get uniformly distributed in the entire disk as seen in Oph IRS 48 \citep{marel13,marel15}. We note that grains are likely growing everywhere in the disk, but transport is concentrating large grains at the pressure bump, resulting in asymmetric structures. LkH$\alpha$ 330 has at least one spiral arm feature and an asymmetric structure about east and west side of the disk. Although there are several explanations for these complex structures, a large unseen planet naturally explains simultaneously both of the structures. Observations at longer wavelengths available in the future with ALMA or VLA will be greatly helpful for further investigations of grain growth and the connection between large sized grains and rocky cores of planets. \section{Summary} \label{summary} We have conducted $H$-band (1.6 $\mu$m) linear polarimetric observations and 0.87 mm continuum observations toward a transitional disk around the intermediate-mass pre-main sequence star LkH$\alpha$ 330. The observations show that a spiral arm potentially exists in the disk interior where planets form. Although several possibilities for the origin of the spiral arm have been proposed, a massive unseen planet would be the most plausible candidate according to previous simulations \citep{zhu12,isella13,dong15}. The possibility of dust trapping and grain growth in the asymmetric disk component was investigated using the opacity index ($\beta$) through the observed SED slope between 0.87 mm and 1.3 mm continuum emission. As a result, a low opacity index, $\beta = 0.7^{+0.5}_{-0.4}$ is obtained in the west side of the disk, while the east side shows a typical value of $\beta = 2.0\pm0.5$ usually seen in the normal primordial disks, suggesting that grain growth probably takes place everywhere and only millimeter-sized grains selectively accumulates at the pressure maxima in the west side of the disk. Combining the results of near-infrared and submillimeter observations, we conjecture that the spiral arms were in the upper surface and an asymmetric (potentially arm-like) structure was in the mid-plane of the disk. Continuum observations at longer wavelengths such as centimeters may show how large dust grains connect to the formation of rocky cores of planets and near-infrared observations with so-called extreme AO are required for further exploring the dust distribution in the disk. \acknowledgments M.T. is supported by Ministry of Science and Technology (MoST) of Taiwan (grant No. 103-2112-M-001-029). Y.H is currently supported by JPL/Caltech. This work is supported by MEXT KAKENHI No. 23103004. \section{Appendix \\ $H$-band radial polarization} \label{appendix} In addition to the $H$-band PI image provided in Figure \ref{fig:fig1}, we additionally report a radial polarization (${\rm U}_\phi$) map as a check for the polarization fraction that is not originated from the source (mainly from instruments). The ${\rm U}_\phi$ image corresponding to the PI image is provided in Figure \ref{fig:fig4}. In principle, polarization vectors have to be orthogonal to the incident radiation from the central star. Since the polarization fraction of ${\rm U}_\phi$ component is approximately 10 times smaller than the polarization intensity in the entire disk as shown in Figure \ref{fig:fig1}, the contribution of the instrumental polarization effect is negligible and not affecting the discussed structures significantly. Thus, we conclude that the observed polarized emission is originated from the disk of interest.
ebfe31924472933bcb400733dfd7c7bde750df64
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Let $A$ and $B$ be nonempty subsets of an abelian group, and define the \emph{sumset} of $A$ and $B$ and the \emph{$h$-fold sumset} of $A$ as \begin{align*}A+B := \{ a + b: a \in A, b \in B\} \text{\hspace{3px} and \hspace{3px}} hA := \{ a_1 + \ldots + a_h: a_i \in A\}, \end{align*} respectively. When the set $A$ is implicitly understood, we will reserve the letter $K$ to denote the doubling constant of $A$; that is, $K := |A+A|/|A|$. A classical result of Pl\"unnecke bounds the cardinality of $hA$ in terms of $K$ and $|A|$. \begin{theorem}[Pl\"unnecke's inequality \cite{plu70}]\label{plunnecke} For any set $A$ and for any nonnegative integers $\ell$ and $m$, if $|A+A| = K |A|$, then \[ |\ell A - m A| \le K^{\ell + m} |A|. \] \end{theorem} \noindent See the survey of Ruzsa \cite{ruz09} for variations, generalizations, and a graph theoretic proof of Theorem \ref{plunnecke}; see Petridis \cite{pet12} for a new inductive proof. Given $\lambda \in \mathbb{Z}$, define a \emph{dilate} of $A$ as \begin{align*}\lambda \cdot A := \{ \lambda a: a \in A\}. \end{align*} Suppose $\lambda_1 , \ldots , \lambda_h$ are nonzero integers. Since $\lambda_i \cdot A \subseteq \lambda_i A$, one can apply Theorem \ref{plunnecke} to conclude that \begin{align*}|\lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{\sum_i |\lambda_i|}|A|. \end{align*} Bukh \cite{buk08} significantly improved this by considering the binary expansion of $\lambda_i$ and using Ruzsa's covering lemma and triangle inequality. \begin{theorem}[Bukh \cite{buk08}]\label{buk08} For any set $A$, if $\lambda_1 , \ldots , \lambda_h \in \mathbb{Z}\setminus \{0\}$ and $|A+A| = K|A|$, then \[ | \lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{7 + 12 \sum_{i=1}^h \log_2 (1+ |\lambda_i|)}|A|. \] \end{theorem} If $|\lambda_i| \le 2^r$ for all $i$, then Theorem \ref{buk08} yields that \begin{align}\label{logbound} | \lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{O(rh)}|A|. \end{align} In this paper we prove a bound that improves \eqref{logbound} when $r$ and $h$ are sufficiently large and comparable to each other. Throughout the paper $ln$ stands for the natural logarithm. \begin{theorem}\label{mainthm} Suppose $r, h \in \mathbb{Z}$ are sufficiently large and \begin{equation} \label{eq:mn} \min \{r+1, h\} \ge 10 \left(\ln \max\{r+1, h\} \right)^2. \end{equation} Given a set $A$ and nonzero integers $\lambda_1, \ldots, \lambda_h$ such that $|\lambda_i| \le 2^r$, if $|A+A| = K|A|$, then \begin{equation} \label{eq:main} | \lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{7 rh/ \ln (r+h)} |A|. \end{equation} \end{theorem} The proof of Theorem \ref{mainthm} relies on Theorem \ref{buk08} as well as a result of Tuza~\cite{Tuza84} on decomposing bipartite graphs into bicliques (complete bipartite subgraphs). The key idea is to connect Bukh's technique of considering the binary expansion of $\lambda_i$ to the graph decomposition problem that allows us to efficiently group certain powers of 2. We remark here that in all of the above theorems, the condition $|A+A| = K|A|$ can be replaced with $|A-A| = K|A|$ with no change to the conclusion. It is likely that Theorem \ref{mainthm} is not best possible -- we discuss this in the last section. \section{Basic Tools} We need the following analogue of Ruzsa's triangle inequality, see \cite[Theorem 1.8.7]{ruz09}. \begin{theorem}[Ruzsa \cite{ruz09}] \label{ruzsatriangle} For any sets $X, Y,$ and $Z$, \[ | X + Y| \le \frac{|X+Z||Z+Y|}{|Z|}. \] \end{theorem} A useful corollary of Theorem~\ref{ruzsatriangle} is as follows. \begin{corollary} \label{easyobs} For any sets $A$ and $B$, if $p_1$ and $p_2$ are nonnegative integers and $|A+A| \le K|A|$, then \[ |B + p_1 A - p_2 A| \le K^{p_1 + p_2 + 1}|B + A|. \] \end{corollary} \begin{proof} Apply Theorem \ref{ruzsatriangle} with $X = B$, $Y = p_1 A - p_2 A$, and $Z = A$, then apply Pl\"unnecke's inequality (Theorem \ref{plunnecke}). \end{proof} We can use Corollary~\ref{easyobs} to prove the following proposition that we will use in the proof of Theorem \ref{mainthm}. \begin{proposition}\label{easyobsgen} If $k_1, \ell_1, \ldots , k_q , \ell_q$ are nonnegative integers, $K> 0$, and $A_1, \ldots , A_q, C$ are sets such that $|A_i + A_i| \le K|A_i|$, then \begin{equation}\label{withc} |C + k_1 A_1 - \ell_1 A_1 + \ldots + k_q A_q - \ell_q A_q| \le |C + A_1 + \ldots + A_q| \cdot K^{q + \sum_{i=1}^q (k_i + \ell_i)}. \end{equation} In particular, \begin{equation}\label{withoutc} |k_1 A_1 - \ell_1 A_1 + \ldots + k_q A_q - \ell_q A_q| \le |A_1 + \ldots + A_q| \cdot K^{q + \sum_{i=1}^q (k_i + \ell_i)}. \end{equation} \end{proposition} \begin{proof} \eqref{withoutc} follows from \eqref{withc} by taking $C$ to be a set with a single element so it suffices to prove \eqref{withc}. We proceed by induction on $q$. The case $q = 1$ follows from Corollary~\ref{easyobs} immediately. When $q > 1$, suppose the statement holds for any positive integer less than $q$. Applying Corollary~\ref{easyobs} with $B = C + k_1 A_1 - \ell_1 A_1 + \ldots + k_{q-1} A_{q-1} - \ell_{q-1} A_{q-1}$ and $A = A_q$, we obtain that \begin{align}\label{stage1} & \hspace{5px} | C + k_1 A_1 - \ell_1 A_1 + \ldots + k_q A_q - \ell_q A_q| \nonumber \\ \le & \hspace{5px} K^{k_q + \ell_q + 1} |C + k_1 A_1 -\ell_1 A_1 + \ldots + k_{q-1} A_{q-1} - \ell_{q-1} A_{q-1} + A_q|. \end{align} Now, let $C' = C + A_q$ and apply the induction hypothesis to conclude that \begin{align}\label{stage2} & \hspace{5px} |C' + k_1 A_1 - \ell_1 A_1 + \ldots + k_{q-1} A_{q-1} - \ell_{q-1} A_{q-1}| \nonumber \\ \le & \hspace{5px}|C' + A_1 + \ldots + A_{q-1}| \cdot K^{q-1 + \sum_{i=1}^{q-1} (k_i + \ell_i)} \end{align} Combining \eqref{stage1} with \eqref{stage2} gives the desired inequality: \begin{equation} |C + k_1 A_1 - \ell_1 A_1 + \ldots + k_q A_q - \ell_q A_q| \le |C + A_1 + \ldots + A_q| \cdot K^{q + \sum_{i=1}^q (k_i + \ell_i)}. \qedhere \nonumber \end{equation} \end{proof} \section{Proof of Theorem \ref{mainthm}} Given $\lambda_1, \ldots , \lambda_h \in \mathbb{Z} \setminus \{0\}$, we define \begin{align}\label{rands} r := \max_i \lfloor \log_2 |\lambda_i| \rfloor \end{align} and write the binary expansion of $\lambda_i$ as \begin{align}\label{lambda}\lambda_i = \epsilon_i \sum_{j=0}^r \lambda_{i,j} 2^j \text{, where $\lambda_{i,j} \in \{0,1\}$ and $\epsilon_i \in \{-1, 1\}$}. \end{align} Bukh's proof of Theorem \ref{buk08} actually gives the following stronger statement. \begin{theorem}[\cite{buk08}] \label{binbound} If $\lambda_1, \ldots , \lambda_h \in \mathbb{Z} \setminus \{0\}$ and $|A+A| = K|A|$, then \begin{align*}| \lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{7 + 10r + 2 \sum_i \sum_j \lambda_{i,j} }|A|. \end{align*} \end{theorem} In his proof of Theorem \ref{buk08}, the first step is to observe that \[ \lambda_1 \cdot A + \ldots + \lambda_h \cdot A \subseteq \sum_{j=0}^r (\lambda_{1,j} 2^j) \cdot A + \ldots + \sum_{j=0}^r (\lambda_{h,j} 2^j) \cdot A. \] In our proof, we also consider the binary expansion of $\lambda_i$, but we do the above step more efficiently by first grouping together $\lambda_i$ that have shared binary digits. In order to do this systematically, we view the problem as a graph theoretic problem and apply the following result of Tuza~\cite{Tuza84}. \begin{theorem}[Tuza~\cite{Tuza84}]\label{ces} There exists $n_0$ such that the following holds for any integers $m\ge n \ge n_0$ such that $n\ge 10 (\ln m) ^2 $. Every bipartite graph $G$ on two parts of size $m$ and $n$ can be decomposed into edge-disjoint complete bipartite subgraphs $H_1, \ldots , H_q$ such that $E(G) = \cup_{i} E(H_i)$ and \begin{equation} \label{eq:Tuza} \sum_{i=1}^q |V(H_i)| \le \frac{3mn}{\ln m}. \end{equation} \end{theorem} Tuza stated this result \cite[Theorem 4]{Tuza84} for the \emph{covers} of $G$, where a \emph{cover} of $G$ is a collection of subgraphs of $G$ such that every edge of $G$ is contained in at least one of these subgraphs. However, the cover provided in his proof is indeed a decomposition. Furthermore, the assumption $n\ge 10 (\ln m) ^2 $ was not stated in \cite[Theorem 4]{Tuza84} but such kind of assumption is needed.\footnote{In his proof, copies of $K_{q, q}$ were repeatedly removed from $G$, where $q= \lfloor \ln m/ \ln j\rfloor$ for $2\le j\le (\ln m)\ln\ln m$. By a well-known bound on the Zarankiewicz problem, every bipartite graph $G$ with parts of size $m$ and $n$ contains a copy of $K_{q, q}$ if $|E(G| \ge (q-1)^{1/q} (n-q+1) m^{1-1/q} + (q-1)m$. A simplified bound $|E(G| \ge (1+ o(1)) n m^{1-1/q}$ was used in \cite{Tuza84} but it requires that $q m^{1/q} = o(n)$.} Indeed, \eqref{eq:Tuza} becomes false when $n= o(\ln m)$ because $\sum_{i=1}^q |V(H_i)| \ge m+ n$ for any cover $H_1, \ldots , H_q$ of $G$ if $G$ has no isolated vertices. Note that \eqref{eq:Tuza} is tight up to a constant factor. Indeed, Tuza~\cite{Tuza84} provided a bipartite graph $G$ with two parts of size $n\le m$ such that every biclique \emph{cover} $H_1, \ldots , H_q$ of $G$ satisfies $\sum_{i=1}^q |V(H_i)| \ge mn/ (e^2\ln m)$, where $e= 2.718...$. \begin{proof}[Proof of Theorem \ref{mainthm}] Let $r, h\in \mathbb{Z}$ be sufficiently large and satisfy \eqref{eq:mn}. Given nonzero integers $\lambda_1, \ldots , \lambda_h$, define $r$ and $\lambda_{i,j}$ as in \eqref{rands} and \eqref{lambda}. We define a bipartite graph $G = (X,Y,E)$ as follows: let $X = \{ \lambda_1, \ldots , \lambda_h\}$, $Y = \{ 2^0, \ldots , 2^r \}$, and $E = \{ (\lambda_i, 2^j): \lambda_{i,j} = 1 \}$. In other words, $\lambda_i$ is connected to the powers of 2 that are present in its binary expansion. We apply Theorem \ref{ces} to $G$ and obtain a biclique decomposition $H_1, \ldots , H_q$ of $G$. Assume $H_i := (X_i, Y_i, E_i)$ where $X_i \subseteq X$, $Y_i \subseteq Y$. We have $E_i = \{ (u,v): u\in X_i, v\in Y_i\}$ and \begin{align}\label{concl} \sum_{i=1}^q \left(|X_i| + |Y_i| \right)\le \frac{ 3 (r+1)h }{\ln \max\{r+1, h\} }. \end{align} Now, we connect this biclique decomposition to the sum of dilates $\lambda_1 \cdot A + \ldots + \lambda_h \cdot A$. Since the elements of $X$ and $Y$ are integers, we can perform arithmetic operations with them. For $j=1, \ldots , q$, let \[ \gamma_j := \sum_{y \in Y_j} y, \] and since $\mathcal{H}$ is a biclique decomposition, for $i = 1, \ldots , h$, we have \[ \lambda_i = \epsilon_i \sum_{j: \lambda_i \in X_j} \gamma_j .\] Applying the above to each $\lambda_i$ along with the fact that $B + (\alpha + \beta) \cdot A \subseteq B + \alpha \cdot A + \beta \cdot A$ results in \begin{align}\label{reversesummation} \lambda_1 \cdot A + \ldots + \lambda_h \cdot A \subseteq \epsilon_1 \sum_{j: \lambda_1 \in X_j} (\gamma_j \cdot A) + \ldots + \epsilon_h \sum_{j: \lambda_h \in X_j} (\gamma_j \cdot A). \end{align} Let $k_j := | \{ \lambda_i \in X_j : \lambda_i > 0 \}|$, $\ell_j := |\{ \lambda_i \in X_j: \lambda_i < 0\}|$, and note that $k_j + \ell_j = |X_j|$. By regrouping the terms in \eqref{reversesummation}, we have \begin{align*} & \hspace{5px}\epsilon_1 \sum_{j: \lambda_1 \in X_j} \gamma_j \cdot A + \ldots + \epsilon_h \sum_{j: \lambda_h \in X_j} \gamma_j \cdot A \\ = & \hspace{5px} k_1 ( \gamma_1 \cdot A ) - \ell_1 (\gamma_1 \cdot A) + \ldots + k_q ( \gamma_q \cdot A) - \ell_q (\gamma_q \cdot A) . \end{align*} Since $|\gamma_j \cdot A + \gamma_j \cdot A| = |A+A| \le K|A| = K| \gamma_j \cdot A |$, we can apply Proposition~\ref{easyobsgen} to conclude that \begin{align}\label{stp1} \nonumber & \hspace{5px}|k_1 ( \gamma_1 \cdot A ) - \ell_1 (\gamma_1 \cdot A) + \ldots + k_q ( \gamma_q \cdot A) - \ell_q (\gamma_q \cdot A)| \\ \le & \hspace{5px}|\gamma_1 \cdot A + \ldots + \gamma_q \cdot A| \cdot K^{q + \sum_{i=1}^q k_i + \ell_i} \le |\gamma_1 \cdot A + \ldots + \gamma_q \cdot A| \cdot K^{2\sum_{i=1}^q |X_i|}. \end{align} For $1\le i\le q$ and $0\le j\le r$, let $\gamma_{i,j} = 1$ if $2^j$ is in the binary expansion of $\gamma_i$ and $0$ otherwise. Observe that \[ \max_j \lfloor \log_2 \gamma_j \rfloor \le \max_i \lfloor \log_2 |\lambda_i| \rfloor = r \text{\hspace{5px} and \hspace{5px}} \sum_{j=0}^r \gamma_{i,j} = |Y_i|. \] Hence, by Theorem \ref{binbound}, \begin{align}\label{stp2} |\gamma_1 \cdot A + \ldots + \gamma_q \cdot A| \le K^{7 + 10r + 2 \sum_{i=1}^q \sum_{j=0}^r \gamma_{i,j}} |A| = K^{7 + 10r + 2 \sum_{i=1}^q |Y_i|} |A|. \end{align} Combining \eqref{stp1} and \eqref{stp2} with \eqref{concl} results in \[ | \lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{7 + 10r + 2 \sum_{i=1}^q (|X_i|+|Y_i|)} |A| \le K^{7 + 10r + \frac{6(r+1)h}{\ln \max\{ r+1, h\} } } |A|. \] We have $7+10r = o( (r+1)h/ \ln \max\{ r+1, h \} )$ because of \eqref{eq:mn} and the assumption that $r, h$ are sufficiently large. Together with \[ \ln \max\{r+1, h\}\ge \ln \frac{r+1+h}2 \ge (1- o(1)) \ln (r+h), \] this implies that $ | \lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{ 7rh / \ln (r+h) } |A|$, as desired. \end{proof} \section{Concluding Remarks} Instead of Theorem~\ref{ces}, in an earlier version of the paper we applied a result of Chung, Erd\H{o}s, and Spencer \cite{chuerdspe83}, which states that every graph on $n$ vertices has a biclique decomposition $H_1, \ldots , H_q$ such that $\sum_{i=1}^q |V(H_i)| \le (1+o(1)) n^2/ (2\ln n)$. Instead of \eqref{eq:main}, we obtained that \[ |\lambda_1 \cdot A + \ldots + \lambda_h \cdot A | \le K^{O((r+h)^2/\ln (r+h) ) } |A|. \] This bound is equivalent to \eqref{eq:main} when $r=\Theta(h)$ but weaker than Bukh's bound \eqref{logbound} when $r$ and $h$ are not close to each other. Although the assumption \eqref{eq:mn} may not be optimal, Theorem~\ref{mainthm} is not true without any assumption on $r$ and $h$. For example, when $r$ is large and $h=o(\ln r)$, \eqref{eq:main} becomes $|\lambda_1 \cdot A + \ldots + \lambda_h \cdot A | \le K^{ o(r ) } |A|$, which is false when $A= \{1, \dots, n\}$. If each of $\lambda_1, \dots, \lambda_r$ has $O(1)$ digits in its binary expansion, then Theorem~\ref{buk08} yields that $|\lambda_1 \cdot A + \ldots + \lambda_r \cdot A| \le K^{O(r)}|A|$. Bukh asked if this bound holds whenever $\lambda_i \le 2^r$: \begin{question}[Bukh \cite{buk08}]\label{bukhconj} For any set $A$ and for any $\lambda_1 , \ldots , \lambda_r \in \mathbb{Z} \setminus \{0\}$, if $|A+A| = K|A|$ and $0 < \lambda_i \le 2^r$, then \[ | \lambda_1 \cdot A + \ldots + \lambda_r \cdot A| \le K^{O(r)}|A|. \] \end{question} \noindent In light of Question~\ref{bukhconj}, one can view Theorem~\ref{mainthm} as providing modest progress by proving a subquadratic bound of quality $O(r^2/\ln r)$ whereas Theorem~\ref{buk08} shows that the exponent is $O(r^2)$. Generalized arithmetic progressions give supporting evidence for Question~\ref{bukhconj}. A \emph{generalized arithmetic progression} $P$ is a set of the form \[ P := \{ d + x_1 d_1 + \ldots + x_k d_k: 0 \le x_i < L_i \}. \] Moreover, $P$ is said to be \emph{proper} if $|P| = L_1 \cdot \ldots \cdot L_k$. One can calculate that if $P$ is proper, then \[ |P+P| \le (2L_i - 1)^k \le 2^k |P| =: K |P|. \] Additionally, one can calculate that for any $\lambda_1, \ldots , \lambda_h \in \mathbb{Z}^+$, if $\lambda_i \le 2^r$ then \[ |\lambda_1 \cdot P + \ldots + \lambda_h \cdot P| \le (\lambda_1 + \ldots + \lambda_h)^k |P| = 2^{k \log_2(\lambda_1 + \ldots + \lambda_h)}|P| = K^{r + \log_2{h}}|P|. \] Freiman's theorem \cite{fre73} says that, roughly speaking, sets with small doubling are contained in generalized arithmetic progressions with bounded dimension. Using this line of reasoning, Schoen and Shkredov \cite[Theorem 6.2]{schshk16} proved that \[ |\lambda_1 \cdot A + \ldots + \lambda_h \cdot A | \le e^{O( \log_2^6 (2K) \log_2\log_2 (4K)) (h + \log_2 \sum_i |\lambda_i| ) }|A|. \] This naturally leads us to ask a more precise version of Question~\ref{bukhconj}. \begin{question} If $|A+A| = K|A|$, then is \[ |\lambda_1 \cdot A + \ldots + \lambda_h \cdot A| \le K^{O(h+ \ln \sum_i |\lambda_i| ) } |A| ?\] \end{question} \section*{Acknowledgment} We thank Boris Bukh and Giorgis Petridis for reading an earlier version of this paper and giving valuable comments. We also thank an anonymous referee for suggesting \cite{Tuza84} as well as other helpful comments.
8ca8defd661a8504d7d26c8c9e32e2b85e08e1bc
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Photonic crystals\cite{Yablonovitch1987,Joannopoulos_2008,Sakoda2013} are systems with a spatially periodic structure of refractive medium. Electromagnetic (EM) fields in a photonic crystal are given by the Bloch states where frequency dispersion may have an energy gap as the electronic band dispersion of solids. This gives us an opportunity to simulate microscopic quantum states of electrons by the macroscopic EM waves governed by the classical Maxwell equations. Especially, simulating topologically nontrivial phases in quantum solids has been an extensively studied topic these days. A prototypical topological phase of matter is an integer quantum Hall state\cite{Klitzing1980}, where its topological nature is encoded in a topological invariant, the Chern number, calculated using the Bloch wave functions\cite{Thouless1982}. A physical consequence of the topological nature is quantization of the Hall conductance, since it is proportional to the Chern number. On the other hand, the quantization is also explained in terms of the chiral edge states\cite{Laughlin1981,Hatsugai1993a}. Generally, there is an intimate relation between bulk topology and edge states, as known as the bulk-edge correspondence\cite{Hatsugai1993b}. So, we can access topological properties of the bulk via edge states, or inversely, we can know about the edge/surface states by the bulk information alone. The idea of the bulk-edge correspondence is especially important in photonic systems \cite{Raghu_2008,Z.Wang_2008,Z.Wang_2009,Ochiai09,Hafezi11,Fang12,Hafezi13,Khanikaev13,Rechtsman2013,L.Ling2013,LuLing14,L.Ling2015,Wu15}, since there is no ``Hall conductance'' in photonic systems, while edge states are always physical observable. Since we have the Bloch states and the Brillouin zone in periodic photonic systems, the Chern number is well-defined as in electronic systems. Then, comparing the bulk Chern number and the edge states is a natural strategy to attack topological issues in photonic systems. In fact, quantum Hall state analogs in 2D photonic crystals have been discussed focusing on the edge states. Note that in such studies, the time reversal symmetry (TRS) breaking is mandatory to realize a quantum Hall analog. In this paper, a Gaussian type localized basis set based method is introduced for numerical evaluation of the Bloch states of the EM waves. With the localized basis set, momentum, which defines the Brillouin zone, is treated as a twisted boundary condition $e^{i\phi_\alpha}$, $\alpha=x,y$, and then, the basis functions naturally and strictly becomes periodic in momentum $\phi_\alpha$. Since we will explain later, this feature is advantageous in the Chern number calculation, and this is the reason for our choice of the localized basis set. The Gaussian expansion is not the only choice for the localized basis set, i.e., for instance, the finite element method \cite{Jin2014,Axmann1999,Dobson1999,Dobson2000,Burger2005} gives an alternative. However, the Gaussian basis element is convenient when derivatives of the basis functions are required, obviously due to its Gaussian nature. Such a situation really arises in 3D crystals where the decomposition of the EM field into TE or TM modes is impossible. We demonstrate the merit of the Gaussian expansion through the calculation of the ``section'' Chern numbre \cite{Avron1983,Halperlin1987,Kohmoto1992,Hatsugai04} of the 3D photonic crystals. As we explained, the introduced method is applied to 3D photonic crystals. First of all, note that it is possible to define and use the Chern number to characterize the given system even in 3D cases as follows. That is, let us fix one of the three components of the momentum, say $k_3$, and define the Chern number $C (k_3)$ using $(k_1,k_2)$ as a periodic parameter space. Since this parameter space is a 2D section of the 3D Brillouin zone, we call $C (k_3)$ the section Chern number. This section Chern number is well defined only when the band gap is always finite on the constant $k_3$ plane (section). Different from the usual Chern number in 2D systems, the section Chern number can be finite even in systems having TRS. This feature enables us to realize a topologically nontrivial state in photonic crystals without magneto-optical media. Specifically, $C (k_3)$ can take a nonzero value when the TRS \textit{or} the spatial inversion symmetry (SIS) is broken. Since the section Chern number $C (k_i)$ is a topological invariant, it can change only when the band gap vanishes at a certain point on the constant $k_i$ plane. Generally, this gap closing point is a Weyl point in the band structure, having a linear dispersion around it. Intriguing topological properties are expected in the system with the Weyl points\cite{Murakami07,X.Wan2011,Burkov2011,Burkov11b}. Also, it has been already shown that the Weyl points emerge in the double gyroid photonic crystal when TRS or SIS is broken \cite{L.Ling2013,L.Ling2015}. In this paper, we consider a photonic crystal of a simple structure with Weyl points to discuss the topological nature. In this paper, we limit ourselves to the cases with TRS but without SIS. The considered system is anisotropic and the Weyl points are clearly resolved only in a constant $k_z$ plane in the Brillouin zone. We also discuss the edge states associated with the nonzero section Chern number, which can be realized by making a wave packet. This paper is organized as follows. In Sec.~II, we introduce the Gaussian expansion method for the photonic crystals, and explain the importance of the localized basis set for the calculation of the Chern number. Section~III is devoted to demonstrate the validity and the usefulness of the introduced method by applying the method to simple models. In Sec.~IV, a specific photonic crystal having the Weyl points in its photonic band structure is proposed, and the section Chern number for that model is evaluated. The relation between the section Chern number and the edge modes is also discussed. Summary and discussion are given in Sec.~V. \section{Methods} \begin{figure}[tbp] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure1.pdf} \caption{(a) Gaussian (spatially localized) basis elements are put on the grid that divides unit cell uniformly in the directions of $\boldsymbol{a}_1$ and $\boldsymbol{a}_2$. (b) Each basis element is set on each unit cell with the factor {$\mathrm{e}^{\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}}$}.} \label{gaussian-picture} \end{center} \end{figure} The Maxwell equations for the normal modes introduced as $\boldsymbol{E} \negthinspace \left( \boldsymbol{r}, t \right) = {\mathrm{e}^{-\mathrm{i} \omega t}} \boldsymbol{E} \negthinspace \left( \boldsymbol{r} \right)$ and $\boldsymbol{H} \negthinspace \left( \boldsymbol{r}, t \right) = {\mathrm{e}^{-\mathrm{i}\omega t}} \boldsymbol{H} \negthinspace \left( \boldsymbol{r} \right)$ are written as \begin{equation} \begin{aligned} \nabla \times \boldsymbol{E} \negthinspace \left( \boldsymbol{r} \right) &= {\mathrm{i} \omega} \boldsymbol{B}(\bm{r}), & \nabla \times \boldsymbol{H} \negthinspace \left( \boldsymbol{r} \right) &= {-\mathrm{i} \omega } \boldsymbol{D}(\bm{r}), \\ \nabla \cdot \boldsymbol{D} \negthinspace \left( \boldsymbol{r} \right) &= 0, & \nabla \cdot \boldsymbol{B} \negthinspace \left( \boldsymbol{r} \right) &= 0. \label{Maxwell-eq} \end{aligned} \end{equation} Here, we assume that the permittivity and the permeability are linear and lossless, i.e., we assume \begin{equation} \boldsymbol{D} \negthinspace \left( \boldsymbol{r} \right) = \varepsilon_0 \hat{\varepsilon} \negthinspace \left( \boldsymbol{r} \right) \boldsymbol{E} \negthinspace \left( \boldsymbol{r} \right), \thickspace \boldsymbol{B} \negthinspace \left( \boldsymbol{r} \right) = \mu_0 \hat{\mu} \negthinspace \left( \boldsymbol{r} \right) \boldsymbol{E} \negthinspace \left( \boldsymbol{r} \right), \end{equation} where $\hat{\varepsilon} \negthinspace \left( \boldsymbol{r} \right)$ and $\hat{\mu} \negthinspace \left( \boldsymbol{r} \right)$ are $3 \times 3$ tensors satisfying the Hermitian conditions $\hat{\varepsilon}^\dagger = \hat{\varepsilon}$ and $\hat{\mu}^\dagger = \hat{\mu} $. With this assumption, the Maxwell equations are rewritten into Hermitian eigenequation forms as \begin{align} \nabla \times \hat{\varepsilon}^{-1}\negthinspace \left( \boldsymbol{r} \right) \nabla \times \boldsymbol{H} \negthinspace \left( \boldsymbol{r} \right) &= \left( \frac{\omega}{c} \right)^2 \hat{\mu} \negthinspace \left( \boldsymbol{r} \right) \boldsymbol{H} \negthinspace \left( \boldsymbol{r} \right), \label{eigenequation-H}\\ \nabla \times \hat{\mu}^{-1}(\bm{r}) \nabla \times \boldsymbol{E} \negthinspace \left( \boldsymbol{r} \right) &= \left( \frac{\omega}{c} \right)^2 \hat{\varepsilon} \negthinspace \left( \boldsymbol{r} \right) \boldsymbol{E} \negthinspace \left( \boldsymbol{r} \right). \label{eigenequation-E} \end{align} We only have to solve the either of these eigenequations with respect to $\boldsymbol{E}$ or $\boldsymbol{H}$. In a spatially periodic system, all the eigenstates are represented as Bloch states. Roughly speaking, there are two ways to expand a Bloch state. The first way is to use a basis set localized in wave space like plane waves \cite{Ho90,Plihal1991,Sozuer1993,Meade93,Johnson01}, and the second way is to use a basis set localized in real space. For the plane wave type basis sets , a Bloch state is usually represented as \begin{equation} \psi_{k} \negthinspace \left( \boldsymbol{r} \right) = \sum_G f_G {\mathrm{e}^{\mathrm{i} \left(\boldsymbol{G} + \boldsymbol{k} \right)\cdot \boldsymbol{r} }} , \label{plane-wave} \end{equation} with $\boldsymbol{G}$ denoting reciprocal lattice vectors. In this form, each basis element is not periodic for a parameter change $\boldsymbol{k} \rightarrow \boldsymbol{k} + \boldsymbol{G} $, since {$\mathrm{e}^{\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{r} }$} and {$\mathrm{e}^{\mathrm{i} \left( \boldsymbol{G} + \boldsymbol{k}\right) \cdot \boldsymbol{r} }$ } are not identical. On the other hand, for spatially localized basis sets, the matrix eigenequation can be naturally made into periodic in parameter $k$. A Bloch state is represented by localized basis set as \begin{equation} \psi_k \negthinspace \left( \boldsymbol{r} \right) = \frac{1}{\sqrt{N}} \sum_m {\mathrm{e}^{\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m }} w_m \negthinspace \left(\boldsymbol{r} \right), \label{localize-wavenumber} \end{equation} with $\boldsymbol{R}_m $ and $N$ representing a lattice vector in real space and the number of unit cells of the system under consideration respectively, and $w_m \negthinspace \left( \boldsymbol{r} \right) = w_0 \negthinspace \left( \boldsymbol{r} - \boldsymbol{R}_m \right)$. In this form, a parameter $k$ only appears as {$\boldsymbol{e}^{\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m }$} and is introduced in the eigenequation as a boundary condition. This periodicity is crucial in the numerical evaluation of the Chern number\cite{Fukui2005}. Representative localized functions are the maximally localized Wannier functions\cite{Marzari1997,Whittaker2003,Busch2003,Marzari2014}, however, it is known that the so-called \textit{composite bands}, which compose the maximally localized Wannier functions, must be distinguished from the bands below and above by gaps with vanishing Chern number \cite{Panati2007,Brouder2007}. However, photonic bands are generally tangling in higher frequencies and it is not common for photonic band structures to have an isolated set of bands enclosed by gaps. Then, we use a Gaussian basis set as a typical localized basis set in the following. Explicitly, the Gaussian basis element is written as \begin{equation} \boldsymbol{g}_{i,k}^{\lambda} \negthinspace \left( \boldsymbol{r} \right) = \frac{1}{\sqrt{N}} \sum_{m,\lambda} {\mathrm{e}^{\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m}} \boldsymbol{f}^{\lambda} \negthinspace \left( \boldsymbol{r}_{im} \right) \mathrm{e}^{- r_{im}^2 / \alpha^2}, \label{gaussian-wavenumber} \end{equation} with \begin{equation} \bm{f}^{ \lambda} (\bm{r}_{im})= \begin{pmatrix} f_x^{ \lambda} \negthinspace \left(x_{im},y_{im},z_{im} \right)\\ f_y^{ \lambda} \negthinspace \left(x_{im},y_{im},z_{im} \right)\\ f_z^{ \lambda} \negthinspace \left(x_{im},y_{im},z_{im} \right) \end{pmatrix}. \end{equation} Here, suffices $i$ and $ \lambda$ specify the grid position within the unit cell and the polarization degrees of freedom respectively, and $\boldsymbol{r}_{im} = \boldsymbol{r} - \left( \boldsymbol{R}_m + \boldsymbol{r}_i \right)$. The $f_i^{ \lambda}\negthinspace \left( x, y, z \right)$ is a polynomial of $\left(x, y, z\right)$. Using this basis set for the eigenfunction expansion in Eq.~\eqref{eigenequation-H}, we obtain a matrix eigenequation \begin{equation} \sum_{j,\lambda'} \Theta_{i\lambda,j \lambda'} \negthinspace \left( \boldsymbol{k} \right) c_{j}^{\lambda'} = \left( \frac{\omega}{c} \right)^2 \sum_{j,\lambda'} O_{i\lambda,j\lambda'} \negthinspace \left( \boldsymbol{k} \right) c_{j}^{\lambda'}. \label{eigenequation-matrix} \end{equation} Here, $\Theta_{i\lambda,j\lambda'} \negthinspace \left(\boldsymbol{k}\right)$ is the matrix element of the operator in the left hand side of Eq.~\eqref{eigenequation-H}, i.e., $\hat{\Theta} = \nabla \times \hat{\varepsilon}^{-1} \negthinspace \left( \boldsymbol{r} \right) \nabla \times$, which is explicitly written as \begin{eqnarray} \Theta_{i\lambda,j\lambda'} \negthinspace \left( \boldsymbol{k} \right) &&= \left< \boldsymbol{g}_{ik}^{ \lambda} \right| \nabla \times \hat{\varepsilon}^{-1} \negthinspace \left( \boldsymbol{r} \right) \nabla \times \left| \boldsymbol{g}_{jk}^{\lambda'} \right> \notag \\ &&= \sum_{m} {\mathrm{e}^{ -\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m }} \int \mathrm{d}^3 r \left( \nabla \times \boldsymbol{f}^{\lambda} \negthinspace \left( \boldsymbol{r}_{im} \right) \mathrm{e}^{- r_{im}^2/\alpha^2 }\right)^\ast \notag \\ && \qquad \qquad \cdot \hat{\varepsilon}^{-1} \negthinspace \left( \boldsymbol{r} \right) \left( \nabla \times \boldsymbol{f}^{\lambda} \negthinspace \left( \boldsymbol{r}_{j0} \right) \mathrm{e}^{- r_{j0}^2 / \alpha^2 }\right). \label{matrix-element-theta} \end{eqnarray} Here we used the Hermiticity of the rotation, $ \int \mathrm{d}^3 r \thickspace \boldsymbol{g}^\ast \negthinspace \left( \boldsymbol{r}\right) \cdot \left\{ \nabla \times \boldsymbol{f} \negthinspace \left( \boldsymbol{r}\right) \right\} = \int \mathrm{d}^3 r \thickspace \boldsymbol{f} \negthinspace \left( \boldsymbol{r}\right) \cdot \left\{ \nabla \times \boldsymbol{g} \negthinspace \left( \boldsymbol{r}\right) \right\}^\ast $ under the boundary condition, $\displaystyle \lim_{r \rightarrow \infty } \boldsymbol{g}_{i,k}^{ \lambda} \negthinspace \left( \boldsymbol{r} \right) = 0$. On the other hand, $O_{i\lambda,j\lambda'} \negthinspace \left( \boldsymbol{k} \right)$ is the overlap, which is obtained as, \begin{eqnarray} O_{i\lambda,j\lambda'} \negthinspace \left( \boldsymbol{k} \right) &&= \left< \boldsymbol{g}_{ik}^{\lambda} \right| \hat{\mu} \negthinspace \left( \boldsymbol{r} \right) \left| \boldsymbol{g}_{jk}^{\lambda'} \right> \notag \\ &&= \sum_{m} {\mathrm{e}^{ -\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m } } \int \mathrm{d}^3 r \left( \boldsymbol{f}^{\lambda} \negthinspace \left( \boldsymbol{r}_{im} \right) \mathrm{e}^{- r_{im}^2 / \alpha^2 } \right)^\ast \notag \\ && \qquad \qquad \cdot \hat{\mu} \negthinspace \left( \boldsymbol{r} \right) \left( \boldsymbol{f}^{\lambda'} \negthinspace \left( \boldsymbol{r}_{j0} \right) \mathrm{e}^{- r_{j0}^2 / \alpha^2} \right). \label{matrix-element-overlap} \end{eqnarray} The integration is performed over the infinite spatial region. In many cases, a photonic crystal is constituted by repetition of structures having different permittivity or permeability from the uniform media. Then, it is convenient to separate the inverse of permittivity $\hat{\varepsilon}^{-1} \negthinspace \left( \boldsymbol{r} \right)$ as \begin{equation} \hat{\varepsilon}^{-1} \negthinspace \left( \boldsymbol{r} \right) = \hat{\varepsilon}_c^{-1} + \left( \hat{\varepsilon}^{-1} \negthinspace \left( \boldsymbol{r} \right) - \hat{\varepsilon}_c^{-1} \right) \label{permittivity-separate}. \end{equation} The first term of the right hand side is constant corresponding to the uniform background media. The second term is nonzero only in the region where the structures exist. For the constant term, the integrals of the matrix elements are the standard Gaussian integral, which can be analytically evaluated. For the second term, the numerical integration is easy since the integrand is finite only in the limited region. The same argument also applies to the inverse of the permeability $\hat{\mu}^{-1} \negthinspace \left(\boldsymbol{r} \right)$. For a uniform system all the matrix elements are given without numerical integration. \subsection{Gaussian basis set (2D case)} First, we consider the 2D case. Here, 2D means that $\hat{\varepsilon} \negthinspace \left( \boldsymbol{r} \right)$ and $\hat{\mu} \negthinspace \left( \boldsymbol{r} \right)$ have no dependence on $z$. In such a system, a mirror plane perpendicular to the $z$ axis exists and the solutions of Eqs.~\eqref{eigenequation-H} and \eqref{eigenequation-E} are separated into TE modes ($H_z$ polarization) or TM modes ($E_z$ polarization). As a result, we only have to consider the $z$ component of either field $\boldsymbol{H}$ or $\boldsymbol{E}$, which indicates that the equation to be solved becomes a scalar equation. For the scalar eigenequation, we use a basis element \begin{equation} g_{ik} \negthinspace \left( \boldsymbol{r} \right) = \frac{1}{\sqrt{N}} \sum_m {\mathrm{e}^{\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m }} \mathrm{e}^{-r_{im}^2 / \alpha^2 }. \label{2d-gaussian-base} \end{equation} Each basis function is cylindrically symmetric and there are finite overlaps between the basis functions. Then, these basis elements are put on the grid dividing the unit cell regularly (Fig.\ref{gaussian-picture}). Later, we will see that the simple configuration of the basis elements gives sufficient accuracy for our purpose, and there arises no need for further attention on the distribution of the grids, which is often important in the standard finite element methods. For this basis set, the integral of the overlap $O_{ij} \negthinspace \left( \boldsymbol{k} \right)$ for the constant term (when $\thinspace \hat{\mu}_c = \hat{1}$) is obtained as \begin{equation} O_{ij}^\text{empty} \negthinspace \left( \boldsymbol{k} \right) = \sum_{m} \frac{\pi \alpha^2 }{2} {\mathrm{e}^{-\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m}} \mathrm{e}^{- \frac{{r_m'}^2 }{2 \alpha^2} }, \label{2d-overlap-value} \end{equation} where $\boldsymbol{r}_m' = \boldsymbol{r}_i + \boldsymbol{R}_m - \boldsymbol{r}_j$ . On the other hand, the integral of the matrix element $\Theta_{ij} \negthinspace \left( \boldsymbol{k} \right)$ for the constant term (when $\hat{\varepsilon}_c^{-1} = \hat{1} $) is \begin{equation} \Theta_{ij}^\text{empty} \negthinspace \left( \boldsymbol{k} \right) = \sum_{m} \frac{\pi \alpha^2}{2} {\mathrm{e}^{-\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m}} \mathrm{e}^{-\frac{{r_m'}^2}{2 \alpha^2} } \frac{1}{\alpha^2} \left( 2 - \frac{{r_m'}^2}{\alpha^2} \right).\label{2d-matrix-value} \end{equation} For the empty lattice, all the matrix elements of Eq.~\eqref{eigenequation-matrix} are given by using Eqs. \eqref{2d-overlap-value} and \eqref{2d-matrix-value}. When some structures are introduced, the additional work to obtain matrix elements is the numerical integration within the structures. \subsection{Gaussian basis set (3D case)} For a 3D system, it is generally impossible to decompose the EM fields into TE or TM modes, and we have to handle all three components of the vector eigenequation. A naive thought suggests us to use the scalar Gaussian functions [Eq.~\eqref{2d-gaussian-base}] for each component of the vector as \begin{equation} \boldsymbol{\psi}_k \negthinspace \left( \boldsymbol{r} \right) = \sum_{ \lambda = 1}^3 \boldsymbol{e}_{ \lambda} \sum_i c_{ik}^{\lambda} g_{ik} \negthinspace \left( \boldsymbol{r} \right) . \label{extensionAtFirst} \end{equation} Here, $\boldsymbol{e}_{ \lambda} $ represents unit vectors in three orthogonal directions. This naively introduced basis set, however, has a deficiency in the following sense. Firstly, note that the divergence of the Eq.~\eqref{eigenequation-H} vanishes because of $\nabla \cdot \left( \nabla \times \boldsymbol{f} \negthinspace \left( \boldsymbol{r} \right) \right) = 0$ for any vector field $\boldsymbol{f} \negthinspace \left( \boldsymbol{r} \right)$. Thus, when the permeability is constant and isotropic, the solution of Eq.~\eqref{eigenequation-H} satisfies the constraint, \begin{equation} \nabla \cdot \boldsymbol{H} \negthinspace \left( \boldsymbol{r} \right) = 0 \label{constraint}. \end{equation} On the other hand, the divergence of Eq.~\eqref{extensionAtFirst} is \begin{equation} \nabla \cdot \sum_{i, \lambda} \boldsymbol{e}_{ \lambda} c_{ik}^{\lambda} g_{ik} \negthinspace \left( \boldsymbol{r} \right) = \sum_{i,\lambda} \left( - 2 \alpha x_{\lambda} \right) c_{ik}^{\lambda} g_{ik} \negthinspace \left( \boldsymbol{r} \right), \end{equation} where the right-hand side does not vanish except in the trivial case with $c_{ik}^{ \lambda} = 0$ for all combinations of $i$ and ${\lambda}$. Namely, expansion Eq.~\eqref{extensionAtFirst} does not satisfy constraint Eq.\eqref{constraint}, and as a consequence, the spectrum of the eigenvalues includes unphysical spurious values. For each basis to satisfy constraint Eq.~\eqref{constraint}, we modify each basis by taking its rotation as \begin{equation} \begin{split} \boldsymbol{g}_{ik}^{\lambda} \negthinspace \left( \boldsymbol{r}\right) &= \nabla \times \boldsymbol{e}_{\lambda} g_{ik} \negthinspace \left( \boldsymbol{r} \right) \\ &= \sum_{ \mu, \nu = 1}^3 \mathrm{\varepsilon}_{\mu \lambda \nu} \left( - 2 \alpha x_{\mu} \right) g_{ik} \negthinspace \left( \boldsymbol{r} \right) \boldsymbol{e}_{\nu} , \label{the-base} \end{split} \end{equation} where $\varepsilon_{\lambda \mu \nu} $ is an anti-symmetric symbol $\left(\varepsilon_{123} = 1 \right)$. This is also obtained by choosing three independent basis elements of the lowest order from Eq.~\eqref{gaussian-wavenumber} under the divergence-free condition. The modified basis element Eq.~\eqref{the-base} has no divergence, and satisfies the constraint [Eq.~\eqref{constraint}]. We expand the solution using these basis elements as \begin{equation} \boldsymbol{\psi}_k \negthinspace \left( \boldsymbol{r} \right) = \sum_{{\lambda} = 1}^3 \sum_i c_{ik}^{\lambda} \boldsymbol{g}_{ik}^{\lambda} \negthinspace \left( \boldsymbol{r} \right) . \label{expansion-vec} \end{equation} Here again, we write down the integral values for this basis set. The integrals for the constant term of the overlaps $O_{iz,jz} \left( \boldsymbol{k} \right) $ and $O_{ix,jz} \left( \boldsymbol{k} \right)$ are \begin{eqnarray} O_{iz,jz}^\text{empty} \negthickspace \left( \boldsymbol{k} \right) && = \sum_{m} \sqrt{\left(\frac{\pi \alpha^2}{2} \right)^3} {\mathrm{e}^{-\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m }} \mathrm{e}^{-\frac{{r_m'}^2}{2 \alpha^2} } \notag \\ && \qquad \qquad \qquad \times \frac{1}{\alpha^2} \left\{ 2 - \frac{1}{\alpha^2} \left( {x_m'}^2 + {y_m'}^2 \right) \right\} \label{3d-overlap-value1} \quad \\ O_{ix,jz}^\text{empty} \negthickspace \left( \boldsymbol{k} \right) && = \sum_m \sqrt{\left(\frac{\pi \alpha^2}{2}\right)^3} {\mathrm{e}^{-\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m}} \mathrm{e}^{-\frac{{r_m'}^2}{2 \alpha^2} } \frac{1}{\alpha^4} x_m' z_m' \label{3d-overlap-value2} \end{eqnarray} respectively, while the integrals for the matrix elements $\Theta_{iz,jz} \left(\boldsymbol{k}\right)$ and $\Theta_{ix,jz} \left(\boldsymbol{k}\right)$ are \begin{eqnarray} && \Theta_{iz,jz}^\text{empty} \negthickspace \left(\boldsymbol{k}\right) = \sum_m \sqrt{\left(\frac{\pi \alpha^2}{2}\right)^3} {\mathrm{e}^{-\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m }} \mathrm{e}^{-\frac{{r_m'}^2}{2 \alpha^2}} \notag \\ && \qquad \times \frac{1}{\alpha^4} \Big[ \frac{1}{\alpha^4} \left( {x_m'}^2 + {y_m'}^2 \right)^2 - \frac{9}{\alpha^2} \left( {x_m'}^2 + {y_m'}^2 \right)\notag \\ && \qquad \qquad + \frac{{z_m'}^2}{\alpha^2} \Big\{ \frac{1}{\alpha^2} \left( {x_m'}^2 + {y_m'}^2 \right) -2 \Big\} + 10\Big] \label{3d-matrix-value1} \\ && \Theta_{ix,jz}^\text{empty} \negthickspace \left(\boldsymbol{k}\right) = \sum_m \sqrt{\left(\frac{\pi \alpha^2}{2}\right)^3} {\mathrm{e}^{-\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{R}_m } } \mathrm{e}^{-\frac{{r_m'}^2 }{2 \alpha^2} } \notag \\ && \qquad \qquad \qquad \qquad \times \frac{{x_m'} {z_m'}}{\alpha^6} \left( 7 - \frac{{r_m'}^2}{\alpha^2} \right) \label{3d-matrix-value2} \end{eqnarray} respectively. The integrals of the other components are obtained by permutations of the indices $x$, $y$, and $z$. In addition, the integrals are invariant against the simultaneous exchange of the position of the localization center and the direction of the polarization $\left( i \lambda' \leftrightarrow j \lambda \right)$. Thus all matrix components for an isotropic bulk dielectric are given by Eqs.~\eqref{3d-overlap-value2}-\eqref{3d-matrix-value2}. The grid points for these basis elements can be chosen to divide the unit cell of the Bravais lattice with a regular spacing like Fig.~\ref{gaussian-picture}. Here, we make some comments on the case of spatially varying $\hat{\mu}(\bm{r})$. In such a case, we have $\nabla\cdot\bm{H}(\bm{r})\neq{0}$, but the expansion (Eq.~\eqref{extensionAtFirst}) still produces a spurious spectrum. However, the basis set (Eq.~\eqref{the-base}) is also insufficient because it has no longitudinal component of $\boldsymbol{H} \negthinspace \left( \boldsymbol{r} \right)$, which should arise in this case. \begin{figure*}[htbp] \begin{center} \includegraphics[width=2\columnwidth]{figures2/figure2.pdf} \caption{Band structures of empty lattices (homogeneous systems $\hat{\varepsilon} = \hat{1}$ and $\hat{\mu} = \hat{1}$). (a) Dispersion of 2D square empty lattice calculated by using $64$ ($8 \times 8$) scalar Gaussian basis elements. The basis elements are located on grid points that regularly the divide square lattice. Results of different settings are compared between $\alpha = d$ (solid lines) and $\alpha = 0.83 d$ (broken lines). For the basis set with smaller overlap (broken line),the lowest eigenvalue at $\Gamma$ point does not converge to $0$. (b) Dispersion of 3D simple cubic empty lattice calculated by using 1,536 basis elements with $\alpha = 1.07 d$ (solid lines) and $\alpha = d$ (broken line). The fictitious divergence of the lowest eigenvalue near $\Gamma$ suppressed by taking larger overlap (solid line). The lowest eigenstate at $\Gamma$ point cannot be restored by the vector Gaussian basis set.}\label{empty-lattice} \end{center} \end{figure*} \subsection{Chern number calculation} The most sophisticated method for the calculation of the Chern number seems to be the one devised by Fukui, Hatsugai and Suzuki\cite{Fukui2005}. Here, let us review the method briefly. In this method, the Chern number is derived on a discretized mesh of the Brillouin zone. An $U \negthinspace \left( 1 \right)$ link variable is defined as \begin{equation} U_\mu \negthinspace \left( \boldsymbol{k}_l \right) \equiv \frac{ \langle n \negthinspace \left( \boldsymbol{k}_l \right) | n \negthinspace \left( \boldsymbol{k}_l + \boldsymbol{\mathrm{e}}_\mu \right) \rangle} {\left| \langle n \negthinspace \left( \boldsymbol{k}_l \right) | n \negthinspace \left( \boldsymbol{k}_l + \boldsymbol{\mathrm{e}}_\mu \right) \rangle \right|} \label{link-variable}, \end{equation} where $\left| n \negthinspace \left( \boldsymbol{k}_l \right) \right>$ is the $n$th eigenvector obtained by diagonalizing an eigenequation, $\boldsymbol{k}_l $ represents a lattice point in the discretized, Brillouin zone and $\boldsymbol{\mathrm{e}}_\mu$ represents one lattice displacement in the direction $\mu\left( = 1, 2\right)$. The $\boldsymbol{k}_l$ is invariant under the displacement of periodic length $\boldsymbol{k}_l + N_\mu \boldsymbol{\mathrm{e}}_\mu = \boldsymbol{k}_l$. From the link variable, a lattice field strength is taken as \begin{eqnarray} \tilde{F}_{12} \negthinspace \left( \boldsymbol{k}_l \right) &&\equiv \ln U_1 \negthickspace \left( \boldsymbol{k}_l \right) U_2 \negthickspace \left( \boldsymbol{k}_l \negthickspace + \negthickspace \boldsymbol{\mathrm{e}}_1 \right) U_1^{\scriptscriptstyle -1} \negthickspace \left( \boldsymbol{k}_l \negthickspace + \negthickspace \boldsymbol{\mathrm{e}}_2 \right) U_2^{\scriptscriptstyle -1} \negthickspace \left( \boldsymbol{k}_l \right) \notag \\ &&- \pi < \frac{1}{\mathrm{i}} \tilde{F}_{12} \negthinspace \left( \boldsymbol{k}_l \right) \le \pi. \label{field-strength} \end{eqnarray} Here the field strength is defined as the principal branch of the logarithm. From this field strength, the Chern number is computed as \begin{equation} \tilde{c} = \frac{1}{2 \pi \mathrm{i} } \sum_l \tilde{F}_{12} \negthinspace \left(\boldsymbol{k}_l \right), \end{equation} where the sum is taken over all the lattice points in Brillouin zone. It is straightforward to extend this method to multiband cases. Equation~\eqref{field-strength} naturally reflects gauge invariance. The result is always given as an integer because of the periodicity of the parameters ($| n\negthinspace \left(\boldsymbol{k}_l + N_\mu \boldsymbol{\mathrm{e}}_\mu \right) \rangle = | n\negthinspace \left(\boldsymbol{k}_l \right) \rangle $) and it can be seen by fixing the gauge over the entire lattice points. A gauge potential is defined as \begin{equation} \widetilde{A}_\mu \negthinspace \left(\boldsymbol{k}_l\right) = \ln U_\mu \negthinspace \left( \boldsymbol{k}_l\right), \quad - \pi < \frac{1}{\mathrm{i}} \widetilde{A}_\mu \negthinspace \left( \boldsymbol{k}_l \right) \le \pi, \end{equation} which is periodic on the lattice $\tilde{A}_\mu \negthinspace \left( \boldsymbol{k}_l + N_\mu \boldsymbol{\mathrm{e}}_\mu \right) = \tilde{A}_\mu \negthinspace \left( \boldsymbol{k}_l \right)$. The field strength (Eq.~\eqref{field-strength}) is related to this gauge potential by \begin{equation} \tilde{F}_{12} \negthinspace \left(\boldsymbol{k}_l \right) = \triangle_1 \widetilde{A}_2 - \triangle_2 \widetilde{A}_1 + 2 \pi \mathrm{i} n_{12} \negthinspace \left( \boldsymbol{k}_l \right), \label{field-strength-gauge-potential} \end{equation} where $\triangle_\mu$ represents a forward difference operation $\triangle_\mu \widetilde{\boldsymbol{A}} \negthinspace \left( \boldsymbol{k}_l \right) = \widetilde{\boldsymbol{A}} \negthinspace \left( \boldsymbol{k}_l + \boldsymbol{e}_\mu \right) - \widetilde{\boldsymbol{A}} \negthinspace \left( \boldsymbol{k}_l \right) $ and $n_{12} \negthinspace \left( \boldsymbol{k}_l \right)$ is an integer-valued field that makes $ - \pi < \tilde{F}_{12}/\mathrm{i} \leq \pi$. Taking the sum of Eq.~\eqref{field-strength-gauge-potential}, the first and second terms of the right hand side cancel between adjacent links. Then the total sum of those terms vanishes because of the periodicity, and the result is given by \begin{equation} \tilde{c} = \sum_l n_{12} \negthinspace \left( \boldsymbol{k}_l \right). \end{equation} For the Gaussian(spatially localized) basis sets, $\left| n \negthinspace \left(\boldsymbol{k} \right) \right>$ is given by the eigenvector of the matrix eigenequation (Eq.~\eqref{eigenequation-matrix}). Since Eq.~\eqref{eigenequation-matrix} is invariant under the displacement $\boldsymbol{k} \rightarrow \boldsymbol{k} + \boldsymbol{G}$, it is guaranteed that the computed Chern number is an integer. \section{Test for the method} \subsection{Empty lattice} In this subsection, we give appropriate settings for the Gaussian basis sets to produce sufficiently accurate results for empty lattices in 2D and 3D cases. Further, using the determined setting, we calculate band structures of typical photonic crystals of both 2D and 3D as examples. Let us begin with the 2D empty lattice, in which the equation to be solved becomes a scalar equation. The results for a square lattice are shown in Fig.~\ref{empty-lattice}(a) for two different settings represented as solid and broken lines. For both of the settings, we use $64$ basis elements on regularly aligned grid points, i.e., eight basis elements per each direction. The broken line is obtained by setting the localization factor in the $\mathrm{e}^{- r^2 / \alpha^2}$ to $\alpha = 0.87 d$ ($d$ being the spacing between two gird points) and the solid line is obtained by setting it to $ \alpha = d$. A good convergence for these settings requires matrix elements up to the fifth-nearest-neighbor pairs of the grid points. The necessary furthest pairs for matrix elements are determined by the overlap length of the matrix $\Theta$ since the convergence of the matrix elements of $\Theta$ is slower than those of $O$. Within the shown frequency scale, the solid and broken lines overlap well and the solid lines match perfectly with the analytically derived dispersion, namely, $\omega = c k$ folded in the first Brillouin zone. However, at the $\Gamma$-point, the lowest band does not converge to $\omega \rightarrow 0$, in $k\rightarrow 0$ limit. This mismatch comes from the fact that the Gaussian basis elements should have sufficiently large overlaps between them in order to compose a spatially homogeneous state, which is expected to be realized in the $\omega\rightarrow 0$ and $k\rightarrow 0$ limit. On the other hand, the accuracy for higher frequency modes is improved by increasing the number of basis elements since higher modes vary more rapidly in space. The failure in the $\omega\rightarrow{0}$ limit is usually irrelevant in the discussion of topological states, since we usually focus on photonic gaps at finite $\omega$ to see topological phenomena. For the 3D case, the vector equations should be treated and the basis set $\left\{ | \boldsymbol{\tilde{g}}_{ik}^{\lambda} \rangle \right\}$ is used. Also in this case, the exact band structure of the empty lattice is well restored by using $1,536$ basis elements, i.e., 512 grid points with eight grid points per each direction, and three basis elements $\left( \lambda = 1, 2, 3\right)$ on each grid point. The results are shown in Fig.~\ref{empty-lattice}(b) for two different settings. The broken line is obtained by setting the localization factor in the $ | \boldsymbol{\tilde{g}}_{ik}^{\lambda} \rangle $[Eq.~\eqref{the-base}] to $\alpha = d$ and the solid line is obtained by setting it to $\alpha = 1.07 d$. However, since the $|\boldsymbol{\tilde{g}}_{ik}^{\lambda} \rangle$ is the Gaussian function multiplied by the first-order polynomial, it extends in a broader region than the genuine Gaussian function does. \begin{figure}[htb] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure3.pdf} \caption{Calculated band structure of dielectric columns arranged in a square lattice in air with radius $r = 0.378 a$. We set the relative permittivity $\varepsilon = 8.9$ in columns ({$\varepsilon = 1$} in air). Solid lines (dashed lines) represents TE (TM) mode dispersion. The calculation is performed using 64 basis elements with $\alpha = d$. } \label{square-columns} \end{center} \end{figure} The difference of the first eigenvalues between $\alpha = d$ and $\alpha = 1.07 d$ near $\Gamma$-point is more prominent than the 2D case. We have confirmed that the results are well converged and match perfectly to the dispersion, $\omega = c k$, folded in the first Brillouin zone in the shown frequency scale when we set $\alpha$ more than $1.07 d$, which takes the matrix elements up to the $8$th nearest neighbor points. Yet the lowest eigenmode at $\Gamma$ point is never restored with the basis set $\{|\bm{\tilde{g}}_{ik}^{ \lambda}\rangle\}$, due to the difficulty in representing a uniform field by $|\bm{\tilde{g}}_{ik}^{\lambda}\rangle$, since $|\bm{\tilde{g}}_{ij}^{ \lambda}\rangle$ is an odd function with respect to its center. \subsection{The case of spatially modulated permittivity} To begin with, we choose an array of dielectric columns arranged on a square lattice as a sample system for the 2D Gaussian basis set. We assume the radius of the column to be $0.378 a$ with $a$ representing the lattice spacing, and the relative permittivity to be $\varepsilon = 8.9$ within the columns ($\varepsilon = 1.0$ out of the columns)\cite{Dobson1999}. The calculated dispersions for TE modes (solid lines) and TM modes (dashed lines) are shown in Fig.~\ref{square-columns}. The calculation is done with the same setting as the empty lattice [the solid line in Fig.~\ref{empty-lattice}(a)]. The numerical integration between Gaussian basis elements converges by fining the integral mesh up to $\alpha / 5$. The precision of the results increases with the number of basis elements in one direction $n_x$. In this system, the most of eigenvalues converge in the shown scale before $n_x = 8$, with which the result in Fig.~\ref{square-columns} is obtained. \begin{figure}[thbp] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure4.pdf} \caption{Calculated band structure of the three-dimensional square rod structure with the width of rod being $0.26a$ and the relative permittivity $\varepsilon$ = $13.0$. The calculation is performed using 1,536 basis elements with $\alpha = 1.07 d$. } \label{cubic-scaffold} \end{center} \end{figure} In this case, the radius of the column $r = 0.378 a$ is relatively large compared to the size of the unit cell. Therefore the structure is smooth and well restored with the integral mesh $\alpha / 5$ for $n_x = 8$. However, when the radius of the column is smaller, the basis set has to detect a finer structure in integration and we need a larger number of basis elements, which implies slower convergence. Another potential source of bad convergence is a sharp structure such as a corner of a rectangular rod, and we should be careful when such a structure is treated in the present method. Next, we choose a simple scaffold structure, namely a three-dimensional square rod structure studied by S\"oz\"uer and Haus\cite{Sozuer1993}, as an example system for the 3D Gaussian basis set. We assume the width of the rod as $0.26 a $ and the relative permittivity $\varepsilon = 13.0$, which is the same situation as in the study of Dobson \textit{et al.} \cite{Dobson2000}. The result shown in Fig.~\ref{cubic-scaffold} is obtained with the same setting as in the empty lattice case [the solid line in Fig.~\ref{empty-lattice}(b)]. As known, there exists full gap between the fifth and sixth bands. As in the case of the empty lattice, there is difficulty in the $\omega\rightarrow 0$ and $k\rightarrow 0$ limit, however, except that point, the global profile of the calculated band structure, such as the existence of the full gap, is consistent with that of the previous works. \section{Section Chern numbers and edge states in a 3D photonic crystal} In this section, we demonstrate the topologically protected edge states of EM waves associated with the finite section Chern number in a 3D photonic crystal with broken SIS. In a 3D periodic system, the wave vector has three components, $k_1$, $k_2$, and $k_3$, and the section Chern number is defined using two of them, regarding the remaining one as a free parameter. If we fix $k_3$, the section Chern number $C_{n} (k_3)$ is defined as \begin{equation} \begin{aligned} C_{n} (k_3) &= \frac{1}{2\pi \mathrm{i}} \int \mathrm{d}k_1 \mathrm{d}k_2 \thickspace B_{n,12} (\bm{k}), \\ B_{n,12} (\bm{k}) &= \partial_{k_1} \mathcal{A}_{n,2} \left( \bm{k} \right) - \partial_{k_2} \mathcal{A}_{n,1} \left( \bm{k} \right), \label{secChern_def} \\ \mathcal{A}_{n,i} (\bm{k}) &= \langle \psi_n (\bm{k}) | \partial_{{k}_i} | \psi_n (\bm{k}) \rangle, \end{aligned} \end{equation} and those defined by fixing $k_1$ or $k_2$ are similarly given by permutations of the components $(k_1,k_2,k_3)\rightarrow (k_2,k_3,k_1) \rightarrow (k_3,k_1,k_2)$. In order to make $C_{n} (k_3)$ well-defined, the gap should remain finite over the entire $k_3$ constant plane. In general, the section Chern number takes any value, but the symmetry of a given system induces restriction. If a system has TRS, the eigenstate for $\bm{k}$ is related to that for $-\bm{k}$ by the bosonic time-reversal operation $\mathcal{T}$, as $\mathcal{T}|\psi_n(\bm{k})\rangle = |\psi_n(-\bm{k})\rangle^*$, which leads to $\bm{B}_n (-\bm{k})=-\bm{B}_n (\bm{k})$. Then, we have $C_{n} (k_3)=0$ for $k_3=0$ and $\pi$, because of $B_n (-k_1,-k_2,0)=-B_n (k_1,k_2,0)$ and $B_n (-k_1,-k_2,-\pi(=\negthickspace \pi))=-B_n (k_1,k_2,\pi)$, and the definition Eq.~\eqref{secChern_def}. If there is SIS in addition to TRS, an additional restriction may arise, that is, SIS gives $|\psi_n(-\bm{k})\rangle=|\psi_n(\bm{k})\rangle$, which indicates $\bm{B}_n (-\bm{k})=\bm{B}_n (\bm{k})$, and if this relation is combined with $\bm{B}_n (-\bm{k})=-\bm{B}_n (\bm{k})$ required by TRS, we finally obtain $\bm{B}_n (\bm{k})=-\bm{B}_n (\bm{k})=0$. Therefore, SIS has to be broken for a nonzero section Chern number if the system has TRS. Owing to the topological nature of $C_{n} (k_3)$, when $C_{n} (k_3)$ changes as a function of $k_3$, there should be a gap closing point somewhere on the corresponding $k_3$ constant plane. On the other hand, as we have noted, the section Chern number is always zero at $k_3=0$ and $\pi$ with TRS. These facts indicate that the finite section Chern number requires existence of a gap closing point. Typically, gap closing occurs on isolated points in the Brillouin zone and the linear dispersion appears around those degeneracy points. A degeneracy point with linear dispersion is named as a Weyl point. Note that $k_z$ appearing in the section Chern number $C_{n} \left(k_z\right)$ is related to the fixed $k_z$ of an incident wave on the photonic crystal. \subsection{Honeycomb array of air hole columns} \begin{figure}[htb] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure5.pdf} \caption{(Color online) (a) Honeycomb array of hollow columns in a dielectric material. (b) Two-dimensional dispersion of the photonic crystal. We use permittivity $\varepsilon = 20.0$, the radius of hollow column $r = \sqrt{3}a / 6$. The band structure of TE modes (red line) has a distinctive degeneracy point, which is clear (indicated by black lines) in the $k_z=0$ plane among both of TE and TM modes.} \label{hollow-columns} \end{center} \end{figure} Now, we consider to realize Weyl points in photonic crystals. In specific, we need a system in which the Weyl point related physics is readily accessible. More specifically, in order to observe chiral edge modes associated with the finite section Chern number $C_{n}(k_3)$, it is required to have a clear gap on the entire two-dimensional plane with fixed $k_3$. Since a clear gap is required only on a fixed $k_3$ plane, a possible strategy to achieve our goal is to start with a two-dimensional system having ``pseudo-gap,'' a frequency region filled with only a few bands, and then to apply appropriate three-dimensional modifications to the system. Because we are handling photonic crystals, the band structure can be modified by controlling the background medium. However, our experience tells us that careless modifications prone to fail, namely, even if Weyl points are successively generated, they are often masked by the other dispersive bands. Hence, the careful design is important and this is what we discuss in the following. In order to realize easily accessible Weyl points, we consider a photonic crystal of hollow columns aligned on the honeycomb lattice as depicted in Fig.~\ref{hollow-columns}. The band structure of this system has some similarity to that of the triangular lattice consisting of dielectric columns\cite{Plihal1991}. This is because the complementary region of the hollow honeycomb lattice composes a triangular lattice. Despite this similarity, we choose to use the hollow honeycomb lattice since it turns out that the existence of the sublattice structure of the honeycomb lattice is advantageous in the following discussion. Without three-dimensional modulation, the second and third TE modes of the hollow honeycomb lattice form 2D Dirac cones at K and K' points, and TM modes do not mask these TE Dirac cones. Owing to SIS of the hollow honeycomb lattice, these TE Dirac cones are actually line degeneracies extending in the $k_z$ direction in the three-dimensional Brillouin zone. Then, SIS should be broken to transform the line degeneracy into Weyl points, and therefore, we reshape each hollow column into a hollow helix. Note that it is possible to modulate columns differently on sublattices A and B of the honeycomb lattice. \subsection{Weyl points and change in section Chern number} \begin{figure}[tbp] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure6.pdf} \caption{(Color online) (a) Columns are modulated into helices with the same twists for sublattices A and B. The radius of the twisting circle is taken to be $r_h = 0.05 a$ and it circles one time with translation in the $z$ direction by a unit length $a$. The system does not have SIS but is invariant under the twofold rotation around the $z$ axis and the axis perpendicular to the $z$ axis. (b) The value of the total section Chern number summed up to the 9th band for each $k_z$ $\left( - \pi/4 \le k_z \le \pi/4 \right)$. (c) Bulk band dispersions for several values of $k_z$. The 1-9th bands are represented by purple lines.} \label{bulk-equal} \end{center} \end{figure} Now, we consider to apply the same twist on the both hollows at sublattices A and B (Fig.~\ref{bulk-equal}). By transforming columns into helices, the SIS is broken and the line degeneracy is lifted except K and K' points, which implies emergence of the Weyl points. This emergence is explained in terms of the point group symmetry as follows. Before the twist is applied, the system has $D_{6h}$ symmetry, and the group of $k$ is $C_{3v}$ on the line parallel to the $k_z$ axis passing through K and K' points, whereas it is $D_{3h}$ for K and K' points. Since both of $C_{3v}$ and $D_{3h}$ contain two-dimensional representations, twofold degeneracy is allowed on the whole line, which supports existence of the line degeneracy without twist. Then, after applying the twist, the reflection symmetry with respect to a plane including the $z$ axis is broken and the group of $k$ on the line parallel to the $k_z$ axis passing through K and K' points turns to $C_3$, except K and K' points. This implies the line degeneracy dissolves except K and K' points since $C_3$ has no multidimensional representation. For K and K' points, the group of $k$ changes from $D_{3h}$ to $D_3$. By the compatibility relation, 2D representations of $D_{3h}$ are connected to 2D representations of $D_3$, and therefore, the degeneracies at K and K' points survive even with the twist. The surviving degeneracy points are two Weyl points with the same chirality since these points are related by two-fold rotation around the $z$ axis. Then, the section Chern number $C_{n}(k_z)$ changes by $\pm 2$ when $k_z$ crosses $k_z=0$, because of \textit{two} Weyl points on the $k_z=0$ plane [Figs.~\ref{bulk-equal}(b) and \ref{bulk-equal}(c)]. The exactly same argument also applies for the $k_z=\pi$ plane, but it is not possible to resolve Weyl points on the $k_z=\pi$ plane, since the bands on which we focus merge into the other bands as we increase $k_z$ from $0$ to $\pi$. Note that the observation of $\pm 2$ jump in $C_{n}(k_z)$ at $k_z=0$ does not contradict with the former statement that $C_{n}(0)$ should be $0$, since at that time we had assumed that the gap is finite on entire $k_z=0$ plane, and this assumption is invalid in this case. Next, we consider to change the twist of helix only for the hollow on the sublattice B [Fig.~\ref{bulk-differentr}(a)]. With this modification, the twofold rotational symmetry around the $k_z$ axis and the axis perpendicular to $k_z$ axis are broken and then, there is no reason to have degeneracy on the $k_z=0$ plane. However, the Weyl points found in Fig.~\ref{bulk-equal}(c) do not simply disappear because Weyl points are topologically stable and only disappear by annihilating as a pair of Weyl points with opposite chiralities \cite{Nielsen81}. Our numerical calculation shows that the Weyl points move to $\pm k_z$ directions as we increase the difference of the radius of the helices. When one of the Weyl points moves in $+k_z$ direction, the other one moves in $-k_z$ direction due to TRS. In this case, the dispersion is fully gapped on the $k_z=0$ plane, and we have $C_{n}(0)=0$, which is consistent with the former statement. The section Chern number becomes finite for large enough $k_z$ as expected from the existence of the Weyl points. \begin{figure}[tbp] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure7.pdf} \caption{(Color online) (a) Different twists for sublattices A and B, with the twisting circle for helix A being $r_{h,A} = 0.05 a$ and for helix B being $r_{h,B} = 0.03 a$. (b) The value of the total section Chern number summed up to the ninth band for each $k_z$ $\left( -\pi/4 \le k_z \le \pi/4 \right)$. (c) Bulk band dispersions for several values of $k_z$. The first to ninth bands are represented by purple lines.}\label{bulk-differentr} \end{center} \end{figure} \subsection{Edge modes} \begin{figure}[tbp] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure8.pdf} \caption{(Color online) (a) The dispersion curve (color spectrum) for a finite width system (eight unit cells) that corresponds to Fig.~\ref{bulk-equal} in bulk, plotted with projected bulk dispersion (yellow lines). The color spectrum indicates on which side the eigenstate inclines by red (left) - yellow (middle) - green (right). (b) The profile of the $n$th eigenstate for $k_z = \pi/4$, $k_2 = (4/15) 2 \pi$ and $k_2 = (11/15) 2 \pi$, which are indicated by the points in the right most panel of (a).} \label{edge-mode-same} \end{center} \end{figure} In order to investigate edge modes, we perform calculations using the developed Gaussian basis elements on the photonic crystal truncated in the $\vec{a}_1$ direction [see Fig.~\ref{edge-mode-same}(b) for the definition of $\vec{a}_1$] with finite width (eight unit cells in specific). At the interface, we place a material with a smaller dielectric constant so as to prevent the light from evading to outside. This is necessary because the momentum/frequency region we are focusing on lies above the light cone (for the same dielectric constant with the hollow region), which means that the light is able to escape from the system. In Figs.~\ref{edge-mode-same} and \ref{edge-mode-differentr}, the dispersion relations of the finite width system are compared with the bulk dispersions projected on the surface for several values of $k_z$. We find several bands apart from the bulk contribution, which signals the existence of edge modes. In fact, it is confirmed that the non-bulk bands are localized at the interfaces by examining the eigenvector for each mode. The eigenvectors also tell us on which side, left or right, each mode localizes. In Figs.~\ref{edge-mode-same} and \ref{edge-mode-differentr}, the red color corresponds to the states on the left edge, while the green color to the states on the right edge. The color spectrum is determined by the moment of the eigenfield $\langle r_1 \rangle_{\scriptscriptstyle \boldsymbol{H}_{n,k}} = \int \mathrm{d}^3 r \thickspace r_1 \vec{H}_{n,k}^2 \negthinspace \left(\boldsymbol{r}\right)$. As we change $k_z$, the way that the edge modes connect bulk bands changes in accordance with the change of the sum $\sum_{n=1}^9 C_{n}(k_z)$, which confirms the bulk-edge correspondence. The same argument also applies to gaps other than the 9th gap with nonzero total section Chern number. When the fixed momentum is conserved for $k_z$ with $\sum_{n=1}^{i} C_{n} \negthinspace (k_z)$ being nonzero, the wave propagates unidirectionally along the interface. \begin{figure}[tbp] \begin{center} \includegraphics[width=\columnwidth]{figures2/figure9.pdf} \caption{(Color online) The dispersion curve (with color spectrum) for a finite width system (eight unit cells) that corresponds to Fig.~\ref{bulk-differentr} in bulk, plotted with projected bulk dispersion (yellow lines). See the caption of Fig.~\ref{edge-mode-same} for the description of the color spectrum.} \label{edge-mode-differentr} \end{center} \end{figure} This photonic crystal has the structure similar to the honeycomb array of twisted waveguides studied by Rechtsman \textit{et al.}\cite{Rechtsman2013}, where evanescently coupled modes are considered. In the system of evanescently coupled waveguides, the $z$ axis, which is the propagation direction of a waveguide mode, is regarded as the \textit{temporal} axis and the system is spatially two-dimensional. There, the twist of a waveguide becomes a temporal periodic modulation for waveguide modes and the appearance of the chiral edge modes is attributed to the Chern number for the \textit{Floquet} band structure. On the other hand, we consider, in this paper, TE-like modes in the spatially three-dimensional photonic crystal and relate the unidirectional propagation of them to the finite value of section Chern numbers. Depending on the dielectric constant of the material, there might exist some wave guide modes near the frequency of the Weyl point and it leads to coupling between edge modes and wave guide modes that extend into bulk. In such a case, it is possible to suppress the mixing between the edge modes and the wave guide modes by shifting the relative positions of them in the frequency space. The relative shift can be achieved by changing the pitch of the helix while keeping its twist unchanged because the dispersion of wave guide modes is easily affected by the periodic length in the $z$ direction, whereas that of TE-like modes is not affected so much. \section{Summary} We proposed the Gaussian basis sets for the calculation of EM fields in 2D and 3D photonic crystals. In the formulation with spatially localized basis elements, the wave number appears as a boundary condition in the eigenequation and the eigenequation strictly becomes periodic in the wave number, which is advantageous for the Chern number calculation. In addition, the localized property of the Gaussian basis element becomes effective for the consideration of finite size or interface effects, which usually requires larger systems, by utilizing iterative algorithms. Besides, in three-dimensional cases, the Gaussian basis element can easily be accommodated to the divergence-free constraint of the Maxwell equations due to its simplicity for differentiation operation. For further improvement, it would be expected to optimize the mesh alignment or the localization factor of the Gaussian basis elements for each structure. We demonstrated the bulk-edge relation between the section Chern number and chiral edge modes in the SIS broken 3D photonic crystal with TRS. It was confirmed that chiral edge modes in the $n$th gap at each $k_z$ reflects the total of the section Chern number below the $n$th gap $C_{n} \left(k_z\right)$. A system with a finite section Chern number is expected to lead to some applications for wave-packet dynamics. For simplicity, we take a situation that the Weyl points exist on the $k_z=0$ plane. When a wave packet is injected into the finite width system of the SIS broken 3D photonic crystal, a wave packet composed of the Bloch states with positive $k_z$ does propagate on the one side, but does not on the other side. If we use the Bloch states with negative $k_z$, the side that allows propagation changes. In this way, the system is useful to filter wave packets with fixed sign of $k_z$. \begin{acknowledgments} We thank S.~Takahashi and S.~Iwamoto for fruitful discussions. This work is partly supported by Grants-in-Aid for Scientific Research, Nos. 26247064, 25107005, and 16K13845 from JSPS. The computation in this work has been done using the facilities of the Supercomputer Center, the institute for Solid State Physics, the University of Tokyo. \end{acknowledgments} \bibliographystyle{apsrev4-1}
69e71ae3c0943105f9da2bc9835a08ef2f8c8a2d
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The DC conductivity formula obtained by the Fermi-liquid theory is shown to be equivalent to the one obtained by the linearized Boltzmann equation in the textbook~\cite{AGD}. This proof is done for the case of spherical Fermi surface. Its extension to the case of anisotropic Fermi surface is done by Maebashi and Fukuyama~\cite{MF1}. The Maebashi-Fukuyama formula is formal but exact. On the other hand, Taylor~\cite{Taylor} derived a formal but exact solution of the linearized Boltzmann equation. These two works employ opposite strategies. The Maebashi and Fukuyama formula is on the basis of the Fermi-surface harmonics which is the polynomial of the velocity. Then the interaction is expanded in terms of the velocity. The Taylor formula is on the basis of the scattering eigenfunction. In this formula the interaction is diagonalized but the velocity is expanded in terms of the scattering eigenfunction. By these two formulae we know the exact forms of the DC conductivity of anisotropic Fermi liquids. Consequently it becomes evident that there is no room for the current-vertex-correction discussed in the fluctuation-exchange (FLEX) approximation~\cite{FLEX}. \section{Maebashi-Fukuyama formula} \noindent {\large\bf 2-1 Fermi-surface harmonics expansion} \vskip 10pt We start from the Fermi-liquid formula~\cite{MF1} for the DC conductivity \begin{equation} \sigma_{xx} = 2 e^2 \sum_{\bf p} \sum_{\bf p'} v_x({\bf p}) \big(K''\big)^{-1}_{{\bf p}{\bf p'}} \ v_x({\bf p'}) \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big), \label{FL} \end{equation} which is (2.34) of \cite{MF1} and equivalent to the result of the linearized Boltzmann equation, (58) of \cite{rev1}. Using the Fermi-surface harmonics~\cite{Allen,EG}, $\psi_L({\bf p})$, (\ref{FL}) is written as \begin{equation} \sigma_{xx} = 2 e^2 \sum_{\bf p} \sum_{\bf p'} \psi_1({\bf p}) \sum_L \sum_{L'} \big(K''\big)^{-1}_{LL'} \psi_L({\bf p}) \psi_{L'}({\bf p'}) \psi_1({\bf p'}) \langle v_x^2 \rangle \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big), \label{anisoFL} \end{equation} where $\psi_1({\bf p}) = v_x({\bf p}) / \sqrt{\langle v_x^2 \rangle}$ and \begin{equation} \big(K''\big)^{-1}_{{\bf p}{\bf p'}} = \sum_L \sum_{L'} \big(K''\big)^{-1}_{LL'} \psi_L({\bf p}) \psi_{L'}({\bf p'}). \end{equation} The normalization factor is defined as \begin{equation} \langle v_x^2 \rangle = {1 \over N(\epsilon)} \sum_{\bf p} v_x({\bf p})^2 \ \delta\big( \epsilon - \epsilon_{\bf p} \big), \end{equation} with \begin{equation} N(\epsilon) = \sum_{\bf p} \delta\big( \epsilon - \epsilon_{\bf p} \big). \end{equation} By the orthonormal condition \begin{equation} {1 \over N(\epsilon)} \sum_{\bf p} \psi_L({\bf p}) \psi_{L'}({\bf p}) \delta\big( \epsilon - \epsilon_{\bf p} \big) = \delta_{LL'}, \label{orthonormal} \end{equation} (\ref{anisoFL}) is written as \begin{equation} \sigma_{xx} = 2 e^2 \int d\epsilon N(\epsilon) \int d\epsilon' N(\epsilon') \big(K''\big)^{-1}_{11} \langle v_x^2 \rangle \Big( - { \partial f \over \partial \epsilon } \Big). \end{equation} Here we have used the relation \begin{equation} \sum_{\bf p} = \int d\epsilon \sum_{\bf p} \delta\big( \epsilon - \epsilon_{\bf p} \big). \end{equation} Introducing the memory-function matrix \begin{equation} \big(M''\big)^{-1}_{11} = \int d\epsilon' N(\epsilon') \big(K''\big)^{-1}_{11}, \end{equation} we obtain \begin{equation} \sigma_{xx} = 2 e^2 \int d\epsilon N(\epsilon) \big(M''\big)^{-1}_{11} \langle v_x^2 \rangle \Big( - { \partial f \over \partial \epsilon } \Big). \label{M} \end{equation} Finally (\ref{M}) is written as \begin{equation} \sigma_{xx} = 2 e^2 \sum_{\bf p} v_x({\bf p}) \big(M''\big)^{-1}_{11} v_x({\bf p}) \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big). \end{equation} In conclusion the DC conductivity is determined by the memory-function matrix~\cite{MF1}. The microscopic calculation of the memory-function matrix based on the Fermi-liquid theory is carried out by Maebashi and Fukuyama~\cite{MF1,MF2}. To obtain a finite DC conductivity we have to take the contribution of the Umklapp process into the calculation of $M''$. Next we have to perform the matrix inversion from $M''$ to $(M'')^{-1}$. \vskip 20pt \noindent {\large\bf 2-2 Collision term} \vskip 10pt In the previous subsection it is clarified that the DC conductivity is determined by the memory-function matrix. In the Boltzmann equation it appears in the collision term. Thus we investigate the collision term after Allen~\cite{Allen} to obtain the feeling of the matrix structure. The collision term $I_{\bf p}$ for $n_{\bf p}$ is given by \begin{equation} I_{\bf p} = \sum_{\bf p'} K''_{{\bf p}{\bf p'}} \Big\{ n_{\bf p'} - n_{\bf p} \Big\}, \label{C_p} \end{equation} as (47) of \cite{rev1}. Using the Fermi-surface harmonics (\ref{C_p}) is written as \begin{equation} I_{\bf p} = \sum_{\bf p'} \sum_L \sum_{L'} K''_{LL'} \psi_L({\bf p}) \psi_{L'}({\bf p'}) \sum_{L''} n_{L''} \Big\{ \psi_{L''}({\bf p'}) - \psi_{L''}({\bf p}) \Big\}. \label{C_L} \end{equation} By the orthonormal condition (\ref{orthonormal}) and $ \psi_0({\bf p}) = 1$, (\ref{C_L}) is written as \begin{equation} I_{\bf p} = \int d\epsilon' N(\epsilon') \sum_L \sum_{L'} \sum_{L''} K''_{LL'} \ n_{L''} \ \psi_L({\bf p}) \Big\{ \delta_{L'L''} - \delta_{L'0} \psi_{L''}({\bf p}) \Big\}. \label{C_L'} \end{equation} Introducing the memory-function matrix \begin{equation} M''_{LL'} = \int d\epsilon' N(\epsilon') K''_{LL'}, \end{equation} (\ref{C_L'}) is written as \begin{equation} I_{\bf p} = \sum_L \psi_L({\bf p}) \Big\{ \sum_{L'} M''_{LL'} \ n_{L'} - \sum_{L''}\sum_{L'} C_{L''L'L} \ M''_{L''0} \ n_{L'} \Big\}, \label{C_L''} \end{equation} where $ C_{LL'L''} $ is the Clebsch-Gordan coefficient defined by \begin{equation} \psi_{L}({\bf p}) \psi_{L'}({\bf p}) = \sum_{L''} C_{LL'L''} \ \psi_{L''}({\bf p}). \label{CG} \end{equation} Thus we obtain the collision term $I_L$ for $n_L$ as \begin{equation} I_L = \sum_{L'} M''_{LL'} \ n_{L'} - \sum_{L''}\sum_{L'} C_{L''L'L} \ M''_{L''0} \ n_{L'}. \end{equation} For anisotropic Fermi surfaces $n_L$ couples to the other $n_{L'}$s via the memory-function matrix so that we have to perform the matrix inversion, $ M'' \rightarrow (M'')^{-1}$, to obtain the conductivity. The case of the spherical Fermi surface is special and we do not have to perform the matrix inversion. By the symmetry $ K''_{{\bf p}{\bf p'}}$ only depends on the angle between ${\bf p}$ and ${\bf p'}$ so that the memory-function matrix becomes a diagonal matrix. Since $ C_{0L'L''} = \delta_{L'L''} $ which is obtained from (\ref{CG}) by using $ \psi_0({\bf p}) = 1$, we obtain \begin{equation} I_L = 4\pi \Big\{ w_L - w_0 \Big\} n_L, \end{equation} where we put $ M''_{LL'} = 4\pi w_L \delta_{LL'} $. This is the result shown in \cite{Peierls}. The DC conductivity is determined by $ n_1 $ and the transport life-time $\tau$ is given by \begin{equation} {1 \over \tau} = 4\pi \Big\{ w_0 - w_1 \Big\}. \label{Peierls} \end{equation} \vskip 20pt \noindent {\large\bf 2-3 Vector mean free path} \vskip 10pt The use of the vector mean free path ${\bf \Lambda}$ is convenient~\cite{Taylor,Ziman} for the discussion of the symmetry property of the conductivity tensor. In the following all the vectors are defined on the Fermi surface. The conductivity tensor ${\boldsymbol \sigma}$ is expressed in the dyadic form as~\cite{MF1,MF2,Taylor,Ziman} \begin{equation} {\boldsymbol \sigma} = 2 e^2 \sum_{\bf p} {\bf v}_{\bf p} \ {\bf \Lambda}_{\bf p} \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big), \label{dyadic} \end{equation} where \begin{equation} {\bf \Lambda}_{\bf p} = \sum_{\bf p'} \big(K''\big)^{-1}_{{\bf p}{\bf p'}} \ {\bf v}_{\bf p'}, \label{Lambda} \end{equation} with $\displaystyle {\bf v}_{\bf p} = \big( v_x({\bf p}), v_y({\bf p}), v_z({\bf p}) \big)$. For systems with cubic symmetry the conductivity tensor reduces to a scalar $\sigma$~\cite{Taylor}. This means the integrand in (\ref{dyadic}) is proportional to the scalar product $\displaystyle {\bf v}_{\bf p} \cdot {\bf v}_{\bf p}$. Thus introducing the anisotropic life time $\tau_{\bf p}$, the vector mean free path is written as \begin{equation} {\bf \Lambda}_{\bf p} = \tau_{\bf p} \ {\bf v}_{\bf p}. \label{cubic} \end{equation} Namely, the direction of ${\bf \Lambda}_{\bf p}$ is the same as that of ${\bf v}_{\bf p}$. Substituting (\ref{cubic}) into (\ref{Lambda}) we obtain \begin{equation} \tau_{\bf p} = \sum_{\bf p'} \big(K''\big)^{-1}_{{\bf p}{\bf p'}} \ { {\bf v}_{\bf p} \cdot {\bf v}_{\bf p'} \over {\bf v}_{\bf p} \cdot {\bf v}_{\bf p} }. \label{aniso-tau} \end{equation} This relation (\ref{aniso-tau}) is a mere rewriting and has no practical use. Using this anisotropic life time the conductivity is expressed as \begin{equation} \sigma_{xx} = 2 e^2 \sum_{\bf p} v_x({\bf p}) \ \tau_{\bf p} \ v_x({\bf p}) \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big). \label{simple} \end{equation} The Fermi-liquid formula \cite{Yamada} by Yamada and Yosida corresponds to this case. The appearance of (\ref{simple}) is simple but it is none other than (\ref{FL}). Actually \begin{equation} \sigma_{xx} = 2 e^2 \sum_{\bf p} { {\bf v}_{\bf p} \cdot {\bf v}_{\bf p} \over 3 } \ \tau_{\bf p} \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big) = 2 e^2 \sum_{\bf p} \sum_{\bf p'} \big(K''\big)^{-1}_{{\bf p}{\bf p'}} \ { {\bf v}_{\bf p} \cdot {\bf v}_{\bf p'} \over 3 } \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big), \end{equation} in the case of cubic symmetry. For systems without cubic symmetry the proportional relation (\ref{cubic}) does not hold~\cite{Taylor,Ziman} so that we have to use the Maebashi-Fukuyama formula discussed above. For systems with spherical symmetry $\tau_{\bf p}$ reduces to a constant $\tau$. In this case the inverse relation of (\ref{Lambda}) \begin{equation} {\bf v}_{\bf p} = \sum_{\bf p'} K''_{{\bf p}{\bf p'}} \ {\bf \Lambda}_{\bf p'}, \end{equation} reduces to \begin{equation} {\bf v}_{\bf p} = \tau \sum_{\bf p'} K''_{{\bf p}{\bf p'}} \ {\bf v}_{\bf p'}. \end{equation} Thus we obtain \begin{equation} {1 \over \tau} = \sum_{\bf p'} K''_{{\bf p}{\bf p'}} \ \big( {\bf {\hat p}} \cdot {\bf {\hat p'}} \big), \label{life-time} \end{equation} with $ {\bf v}_{\bf p} = {\bf p} / m $ and $ {\bf {\hat p}} = {\bf p}/|{\bf p}| $. Since \begin{equation} K''_{{\bf p}{\bf p'}} = {1 \over \tau_0} \delta_{{\bf p}{\bf p'}} - W_{{\bf p}{\bf p'}}, \end{equation} and \begin{equation} {1 \over \tau_0} = \sum_{\bf p'} W_{{\bf p}{\bf p'}}, \end{equation} (\ref{life-time}) is written as \begin{equation} {1 \over \tau} = \sum_{\bf p'} W_{{\bf p}{\bf p'}} \big( 1 - \cos\theta \big), \label{constant} \end{equation} with $ \cos\theta = {\bf {\hat p}} \cdot {\bf {\hat p'}} $. This expression (\ref{constant}) is none other than (\ref{Peierls}). \section{Taylor formula} \noindent {\large\bf 3-1 Setup} \vskip 10pt Taylor~\cite{Taylor} introduced the vector mean free path ${\bf \Lambda}$ and wrote down the DC conductivity tensor ${\boldsymbol \sigma}$ in the dyadic form as \begin{equation} {\boldsymbol \sigma} = 2 e^2 \sum_{\bf p} {\bf v}({\bf p}) \ {\bf \Lambda}({\bf p}) \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big). \end{equation} In the following all the vectors are defined on the Fermi surface. The vector mean free path is determined by \begin{equation} {\bf v}({\bf p}) = \sum_{\bf p'} Q({\bf p},{\bf p'}) \big[ {\bf \Lambda}({\bf p}) - {\bf \Lambda}({\bf p'}) \big]. \label{vector} \end{equation} The Fermi-liquid theory works~\cite{MF1} on the calculation of the collision operator $Q$. Hereafter we only consider the $x$-component of (\ref{vector}) \begin{equation} v_x({\bf p}) = \sum_{\bf p'} Q({\bf p},{\bf p'}) \big[ \Lambda_x({\bf p}) - \Lambda_x({\bf p'}) \big]. \label{x-com} \end{equation} Introducing the inner product \begin{equation} (a,b) \equiv \sum_{\bf p} \Big( - { \partial f_{\bf p} \over \partial \epsilon_{\bf p} } \Big) a({\bf p})b({\bf p}), \end{equation} the DC conductivity is compactly written as\footnote{ See, for example, Ashcroft and Mermin: {\it Solid State Physics} (1976) Chap. 16 - PROBLEM 4. } \begin{equation} \sigma_{xx} = 2 e^2 \big( v_x, \Lambda_x \big). \label{DC} \end{equation} \vskip 20pt \noindent {\large\bf 3-2 Loose explanation} \vskip 10pt Loosely (\ref{x-com}) is written as \begin{equation} v_x = \Big( {1 \over \tau} - {\hat Q} \Big) \Lambda_x, \label{matrix-x} \end{equation} where \begin{equation} {\hat Q} \Lambda_x \equiv \sum_{\bf p'} Q({\bf p},{\bf p'}) \Lambda_x({\bf p'}), \end{equation} and \begin{equation} {1 \over \tau} \Lambda_x \equiv \sum_{\bf p'} {1 \over \tau({\bf p})} \delta_{{\bf p}{\bf p'}} \Lambda_x({\bf p'}), \end{equation} with \begin{equation} {1 \over \tau({\bf p})} \equiv \sum_{\bf p'} Q({\bf p},{\bf p'}). \end{equation} The inverse of (\ref{matrix-x}) is \begin{equation} \Lambda_x = \Big( {1 \over \tau} - {\hat Q} \Big)^{-1} v_x = \tau v_x + \tau {\hat Q} \tau v_x + \tau {\hat Q} \tau {\hat Q} \tau v_x + \cdot\cdot\cdot. \label{inversion} \end{equation} The perturbative expansion for the DC conductivity is shown in Fig. 1. \vskip 8mm \begin{figure}[htbp] \begin{center} \includegraphics[width=12cm]{Ladder.eps} \caption{ Diagrammatic representation of the first three inner products for the DC conductivity: $ (v_x,\tau v_x) $, $ (v_x,\tau {\hat Q} \tau v_x) $ and $ (v_x,\tau {\hat Q} \tau {\hat Q} \tau v_x) $. } \label{fig:1} \end{center} \end{figure} If we can find the scattering eigenfunction $a_s({\bf p})$ which satisfies \begin{equation} \tau({\bf p}) \sum_{\bf p'} Q({\bf p},{\bf p'}) a_s({\bf p'}) = \alpha_s a_s({\bf p}), \label{Q-alpha} \end{equation} with the eigenvalue $\alpha_s$, the inversion (\ref{inversion}) is readily accomplished. To proceed we expand the velocity in terms of the scattering eigenfunction as \begin{equation} \tau({\bf p}) v_x({\bf p}) = \sum_s \beta_s a_s({\bf p}). \label{v-beta} \end{equation} Employing (\ref{Q-alpha}) and (\ref{v-beta}) in (\ref{inversion}) we obtain\footnote{ \begin{equation} \tau {\hat Q} \tau v_x = \tau({\bf p}) \sum_{\bf p'} Q({\bf p},{\bf p'}) \sum_s \beta_s a_s({\bf p'}) = \sum_s \beta_s \alpha_s a_s({\bf p}), \nonumber \end{equation} and so on. } \begin{equation} \Lambda_x({\bf p}) = \sum_s \big[ 1 + \alpha_s + \alpha_s^2 + \cdot\cdot\cdot \big] \beta_s a_s({\bf p}), \label{Lambda-ab} \end{equation} To proceed further we have to establish the orthonormality of the scattering eigenfunction. \vskip 20pt \noindent {\large\bf 3-3 Tight explanation} \vskip 10pt First we rewrite (\ref{Q-alpha}) as \begin{equation} \sum_{\bf p'} \sqrt{\tau({\bf p})} Q({\bf p},{\bf p'})\sqrt{\tau({\bf p'})}\ { a_s({\bf p'}) \over \sqrt{\tau({\bf p'})} } = \alpha_s \ { a_s({\bf p}) \over \sqrt{\tau({\bf p})} }, \end{equation} Since $ \sqrt{\tau({\bf p})}Q({\bf p},{\bf p'})\sqrt{\tau({\bf p'})} $ is a Hermitian matrix\footnote{ $Q({\bf p},{\bf p'})$ itself is a Hermitian matrix.}, the function $a_s({\bf p}) / \sqrt{\tau({\bf p})}$ constitutes an orthogonal basis. Thus the orthonormal condition is written as \begin{equation} \Big( {a_s \over \sqrt{\tau}}, {a_{s'} \over \sqrt{\tau}} \Big) = \Big( a_s, {a_{s'} \over \tau} \Big) = \delta_{ss'}. \label{orthonormal-a} \end{equation} Employing (\ref{v-beta}) and (\ref{Lambda-ab}) in (\ref{DC}) we obtain \begin{equation} \sigma_{xx} = 2 e^2 \sum_{s'} \sum_s \beta_{s'} {1 \over 1 - \alpha_s } \beta_s \Big( {a_{s'} \over \tau}, a_s \Big). \end{equation} Finally making use of the orthonormal condition (\ref{orthonormal-a}) we obtain the Taylor formula for the DC conductivity \begin{equation} \sigma_{xx} = 2 e^2 \sum_s \beta_{s} {1 \over 1 - \alpha_s } \beta_s. \end{equation} The repeated scatterings shown in Fig. 1 result in the renormalization of the life time $\tau$. \section{Current-vertex-correction in FLEX approximation} The purpose of this note is to clarify the exact expression for the DC conductivity of anisotropic Fermi liquids. Consequently it becomes evident that there is no room for the current-vertex-correction discussed in the FLEX approximation~\cite{FLEX}. Other deficiencies of the FLEX approximation have been also discussed previously~\cite{noFLEX1,noFLEX2}.
93072f03769fed1b3db948312dede77e18ba62a8
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Our main aim in this article is to analyze and then understand the structure of flat epimorphisms of rings more deeply. Indeed, we have successfully applied the Gabriel localization theory and thereby we have found a new and simple proof to the fact that ``a ring map is a flat epimorphism if and only if it corresponds to a type of the G-localizations", see Corollary \ref{G-coro 2} and Theorem \ref{remark 200}. We should mention that, according to the knowledge of the author, there are some different proofs of this theorem or a part of it in the literature. E.g. see \cite{Akiba}, \cite{Nastasescu-Popescu}, \cite{Olivier}, \cite{Popescu-Spircu} and \cite[Chap. XI, Theorem 2.1]{Stenstrom}. These proofs are based upon other enormous results and therefore it is practically very hard to follow and fully understand the proof. In fact, they use the theory of Grothendieck categories and Giraud subcategories and also they use the fact that ``the set of idempotent topologizing systems on the ring $R$ is bijectively corresponding to the set of Giraud subcategories of the category of $R-$modules", see \cite[Chap. X, Theorem 2.1]{Stenstrom}. While the latter in turn is a huge result. Our approach, unlike their methods, is based upon some simple observations. More precisely, after developing the Gabriel localization theory for commutative rings we then obtain more general and new results, namely Theorems \ref{G-lemma 10}, \ref{G-th 5} and \ref{G-lemma 1}. These results pave the way in order to reach to a natural and simple proof of the fact. Using these results then we are also able to characterize the exactness of the G-localization functor, see Theorem \ref{th621890}. Note that the G-localization functor, in general, is left exact. Moreover, as a by-product of this study, a characterization for the flatness of the quotient rings is given, Corollary \ref{coro 87}. \\ The localization theory, in particular the local rings, play a major role in commutative algebra, algebraic and arithmetic geometry, number and valuation theories. The notion of the G-localization with respect to an idempotent topologizing system, which is in turn a natural generalization of the usual localization theory, first appeared in the Gabriel thesis \cite[Chap V, \S 2]{Gabriel} which was conducted under the supervision of Grothendieck. In the literature, an idempotent topologizing system is also called a Gabriel topology. The Gabriel localization theory provides a very general method of localization which is even applicable in noncommutative situations. In commutative algebra a large number of important constructions are special cases of the G-localizations. The idea of the G-localization is essentially due to Grothendieck and it appeared in more general setting in SGA 4, tomme 1, expos\'{e} II. In this article, our presentation of the theory will closely follow the Gabriel's approach in Bourbaki \cite[Les exercices 16 \`{a} 25]{Bourbaki}.\\ The article is organized as follows. In Section 2, as mentioned in the above, the Gabriel localization theory is developed for commutative rings. The content of the third section was completely described in the first paragraph. Regarding to this section, we should mention that this expression of flat epimorphisms (Corollary \ref{G-coro 2}) has some important applications. For example, using this result then one can show that every injective flat epimorphism of rings which is also of finite type then it is of finite presentation. The latter is also a highly non-trivial result in commutative algebra. In the final section, in addition to the characterizing the exactness of the G-localization functor, the structure of prime ideals in the G-localization rings are also studied. Theorems \ref{th621890} and \ref{theorem 33} are the main results of this section.\\ \section{G-localizations} Throughout the article, all of the rings which we consider are commutative.\\ \begin{definition} A \emph{topologizing system} on the ring $R$ is a non-empty family $\mathscr{F}$ of ideals of $R$ satisfying in the following conditions.\\ $\textbf{(a)}$ Every ideal of $R$ containing an ideal $I\in\mathscr{F}$ belongs to $\mathscr{F}$.\\ $\textbf{(b)}$ The family $\mathscr{F}$ is stable under finite intersections.\\ \end{definition} \begin{example}\label{exam 1} Let $R$ be a ring, let $S$ be a multiplicative subset of $R$ and let $\mathcal{P}$ be a family of prime ideals of $R$. Examples of topologizing systems are the single-point set $\{R\}$, the set of ideals of $R$ containing a fixed ideal, the set of ideals $I$ of $R$ such that $I+J=R$ where $J$ is a fixed ideal of $R$, the set of ideals of $R$ which meeting $S$, the set of ideals $I$ of $R$ such that $\mathcal{P}\cap V(I)=\emptyset$, the set of ideals $I$ of $R$ such that $V(I)$ is contained in $V(J)$ where $J$ is a fixed ideal of $R$. See also Theorem \ref{G-lemma 1} and \cite[Chap. X, Theorem 2.1]{Stenstrom}. By $V(I)$ we mean the set of prime ideals $\mathfrak{p}$ of $R$ such that $I\subseteq\mathfrak{p}$. \\ \end{example} \begin{definition} Let $\mathscr{F}$ be a topologizing system on the ring $R$. An $R-$module $M$ is called $\mathscr{F}-$\emph{negligible} if the annihilator of every element of $M$ belongs to $\mathscr{F}$. Clearly submodules, quotients, localizations and finite direct sums of $\mathscr{F}-$negligible modules are $\mathscr{F}-$negligible.\\ \end{definition} Let $\mathscr{F}$ be a topologizing system on the ring $R$. For every $R-$module $M$, the set $\mathscr{F}(M)=\{m\in M : \Ann_{R}(m)\in\mathscr{F}\}$ is a $R-$submodule of $M$ because for every elements $m,m'\in\mathscr{F}(M)$ and for each $r\in R$ we have $\Ann(m)\cap\Ann(m')\subseteq\Ann(m+m')$ and $\Ann(m)\subseteq\Ann(rm)$. Clearly it is the greatest $\mathscr{F}-$negligible submodule of $M$. Each $R-$linear map $u:M\rightarrow N$ induces a map $\mathscr{F}(M)\rightarrow\mathscr{F}(N)$ given by $m\rightsquigarrow u(m)$ which we denote it by $\mathscr{F}(u)$. In fact $\mathscr{F}(-)$ is a left exact functor from the category of $R-$modules to itself. \\ Let $\mathscr{F}$ be a topologizing system on the ring $R$. The family $\mathscr{F}$ with the relation $I<J$ if $J$ is a proper subset of $I$, is a directed poset. If $I\leq J$ then for each $R-$module $M$ the canonical injection $J\subseteq I$ induces the $R-$linear map $u_{I,J}:\Hom_{R}(I,M)\rightarrow\Hom(J,M)$ given by $f\rightsquigarrow f|_{J}$. Clearly $\big(\Hom_{R}(I,M), u_{I,J}\big)$ is an inductive system of $R-$modules and $R-$homomorphisms over the directed poset $(\mathscr{F},<)$. We shall denote by $M_{(\mathscr{F})}$ the inductive limit (colimit) of the system. Therefore $M_{(\mathscr{F})}=\colimit_{I\in\mathscr{F}}\Hom_{R}(I,M)$. For each $I\in\mathscr{F}$ the canonical map $\Hom_{R}(I,M)\rightarrow M_{(\mathscr{F})}$, if there is no confusion, is denoted by $[\:]$. The composed map $\xymatrix{M\ar[r]^{\simeq\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:}& \Hom_{R}(R,M)\ar[r]&M_{(\mathscr{F})}}$ is also denoted by $\delta_{M}$. Clearly $\Ker(\delta_{M})=\mathscr{F}(M)$. Moreover $\Coker(\delta_{M})$ is $\mathscr{F}-$negligible. Because let $f:I\rightarrow M$ be a $R-$linear map where $I\in\mathscr{F}$. For each $a\in I$, $(\delta_{m})|_{I}=a.f$ where $m=f(a)$ and $\delta_{m}: R\rightarrow M$ is given by $r\rightsquigarrow rm$. This means that $I\subseteq\Ann_{R}\big([f]+\Ima(\delta_{M})\big)$.\\ Every $R-$linear map $u:M\rightarrow N$, by the universal property of the colimits, induces a unique $R-$linear map $u_{(\mathscr{F})}:M_{(\mathscr{F})}\rightarrow N_{(\mathscr{F})}$ such that for each $I\in\mathscr{F}$ the following diagram is commutative $$\xymatrix{ \Hom_{R}(I,M)\ar[r]^{} \ar[d]^{} & \Hom_{R}(I,N)\ar[d]^{} \\ M_{(\mathscr{F})}\ar[r]^{\:\:\:\:u_{(\mathscr{F})}} & N_{(\mathscr{F})}}$$ where the columns are the canonical maps and the top row map is given by $f\rightsquigarrow u\circ f$. Therefore $u_{(\mathscr{F})}$ is defined as $[f]\rightsquigarrow[u\circ f]$. \\ \begin{lemma}\label{G-lemma 4} Let $\mathscr{F}$ be a topologizing system on the ring $R$. If \\ $\xymatrix{0\ar[r]&M'\ar[r]^{u}&M\ar[r]^{v}&M''}$ is an exact sequence of $R-$modules then the sequence $\xymatrix{0\ar[r]&M'_{(\mathscr{F})}\ar[r]^{u_{(\mathscr{F})}}& M_{(\mathscr{F})}\ar[r]^{v_{(\mathscr{F})}}&M''_{(\mathscr{F})}}$ is exact.\\ \end{lemma} {\bf Proof.} It is an easy exercise. $\Box$ \\ Let $\mathscr{F}$ and $\mathscr{G}$ be two topologizing systems on the ring $R$. We denote by $\mathscr{F}.\mathscr{G}$ the set of ideals $I$ of $R$ such that there exists some $J\in\mathscr{G}$ containing $I$ in which $J/I$ is $\mathscr{F}-$negligible. Clearly $\mathscr{F}.\mathscr{G}$ is a topologizing system on the ring $R$. Also $IJ\in\mathscr{F}.\mathscr{G}$ for all $I\in\mathscr{F}$ and all $J\in\mathscr{G}$. In particular, $\mathscr{F}$ and $\mathscr{G}$ are contained in $\mathscr{F}.\mathscr{G}$. A topologizing system $\mathscr{F}$ is called \emph{idempotent} if $\mathscr{F}.\mathscr{F}=\mathscr{F}$. All of the topologizing systems in Example \ref{exam 1}, except the second one, are idempotent.\\ \begin{proposition} Let $\mathscr{F}$ and $\mathscr{G}$ be two topologizing systems on the ring $R$ and let $M$ be a $R-$module. Then $M$ is $\mathscr{F}.\mathscr{G}-$negligible if and only if there exists a $\mathscr{F}-$negligible submodule $M'$ of $M$ such that $M/M'$ is $\mathscr{G}-$negligible. In particular, if $\mathscr{H}$ is a third topologizing system on the ring $R$ then we have $(\mathscr{F}.\mathscr{G}).\mathscr{H}= \mathscr{F}.(\mathscr{G}.\mathscr{H})$.\\ \end{proposition} {\bf Proof.} If $M$ is $\mathscr{F}.\mathscr{G}-$negligible then take $M'=\mathscr{F}(M)$ and the remaining assertions are straightforward. The converse is also routine. $\Box$ \\ \begin{lemma}\label{G-lemma 5} Let $R$ be a ring and let $\mathscr{F}$ be a non-empty family of ideals of $R$. Then $\mathscr{F}$ is an idempotent topologizing system on the ring $R$ if and only if it satisfies in the following conditions.\\ $\textbf{(i)}$ Every ideal of $R$ containing an ideal $I\in\mathscr{F}$ belongs to $\mathscr{F}$.\\ $\textbf{(ii)}$ If $I\in\mathscr{F}$ and $J$ is an ideal of $R$ such that $J:a\in\mathscr{F}$ for all $a\in I$, then $J\in\mathscr{F}$.\\ \end{lemma} {\bf Proof.} Suppose $\mathscr{F}$ is an idempotent topologizing system on the ring $R$. Let $I\in\mathscr{F}$ and let $J$ be an ideal of $R$ such that $J:a\in\mathscr{F}$ for all $a\in I$. It suffices to show that $J\in\mathscr{F}.\mathscr{F}$. But it is clear since $J'=I+J\in\mathscr{F}$ and for each $a\in I$, $\Ann_{R}(a+J)=J:a\in\mathscr{F}$. \\Conversely, suppose $I,J\in\mathscr{F}$. We show that $I\cap J\in\mathscr{F}$. But $J\subseteq IJ:a$ for all $a\in I$. Therefore $IJ\in\mathscr{F}$ and so $I\cap J\in\mathscr{F}$. Hence, $\mathscr{F}$ is a topologizing system. Take $J\in\mathscr{F}.\mathscr{F}$. Thus there exists some $I\in\mathscr{F}$ containing $J$ such that $I/J$ is $\mathscr{F}-$negligible. Therefore $J:a\in\mathscr{F}$ for all $a\in I$. Thus $J\in\mathscr{F}$. $\Box$ \\ From now onwards, if it is not stated, $\mathscr{F}$ is always an idempotent topologizing system on the ring $R$.\\ Here some new rings and modules are introduced. The basic set-up is as follows. Let $M$ be a $R-$module. For each $I,J\in\mathscr{F}$ and for each $f\in\Hom_{R}(I,R)$ we have $f^{-1}(J)\in\mathscr{F}$. Because for each $a\in I$, $J\subseteq\Ann_{R}\big(a+ f^{-1}(J)\big)$. Therefore $I/f^{-1}(J)$ is $\mathscr{F}-$negligible and so $f^{-1}(J)\in\mathscr{F}.\mathscr{F}=\mathscr{F}$. The map $f$ induces a map $f^{-1}(J)\rightarrow J$ which we denote it as usual by $f|_{f^{-1}(J)}$. Now we define the pairing $R_{(\mathscr{F})}\times M_{(\mathscr{F})}\rightarrow M_{(\mathscr{F})}$ as $[f].[g]=[g\circ\big(f|_{f^{-1}(J)}\big)]$ where $J$ is the domain of $g$. Note that it is well-defined. Because let $[f_{1}]=[f_{2}]$ and $[g_{1}]=[g_{2}]$ where $f_{k}\in\Hom_{R}(I_{k},R)$ and $g_{k}\in\Hom_{R}(J_{k},M)$ for some $I_{k},J_{k}\in\mathscr{F}$ with $k=1,2$. There exist $L',L''\in\mathscr{F}$ such that $I_{k}\leq L'$, $J_{k}\leq L''$, $(f_{1})|_{L'}=(f_{2})|_{L'}$ and $(g_{1})|_{L''}=(g_{2})|_{L''}$. Take $L=f_{1}^{-1}(L'')\cap f_{2}^{-1}(L'')\cap L'$ which belongs to $\mathscr{F}$ and we have $\big(g_{1}\circ (f_{1})|_{f_{1}^{-1}(J_{1})}\big)|_{L}=\big(g_{2}\circ (f_{2})|_{f_{_{2}}^{-1}(J_{2})}\big)|_{L}$. The pairing is also $R-$bilinear (details omitted). \\ In particular, the binary operation $R_{(\mathscr{F})}\times R_{(\mathscr{F})}\rightarrow R_{(\mathscr{F})}$, as multiplicative, puts a commutative ring structure on the $R-$module $R_{(\mathscr{F})}$. Its commutativity implies from this simple observation that for every $R-$linear maps $f,g:I\rightarrow R$ where $I$ is an ideal of $R$ and for every elements $a,a'\in I$ then $g\big(f(aa')\big)=f\big(g(aa')\big)$. The canonical map $\delta_{R}:R\rightarrow R_{(\mathscr{F})}$ is a ring homomorphism. Moreover, the map $R_{(\mathscr{F})}\times M_{(\mathscr{F})}\rightarrow M_{(\mathscr{F})}$ puts a $R_{(\mathscr{F})}-$module structure over $M_{(\mathscr{F})}$. For every $R-$linear map $u:M\rightarrow N$ then $u_{(\mathscr{F})}:M_{(\mathscr{F})}\rightarrow N_{(\mathscr{F})}$ is $R_{(\mathscr{F})}-$linear. Indeed, $M\rightsquigarrow M_{(\mathscr{F})}$ is a left exact functor from the category of $R-$modules to the category of $R_{(\mathscr{F})}-$modules.\\ \begin{definition} Let $M$ be a $R-$module. The $R_{(\mathscr{F})}-$module $$\big(M/\mathscr{F}(M)\big)_{(\mathscr{F})}$$ is called the \emph{Gabriel localization} of $M$ with respect to the system $\mathscr{F}$ and it is denoted by $M_{\mathscr{F}}$. Therefore $M_{\mathscr{F}}= \colimit_{I\in\mathscr{F}}\Hom_{R}\big(I,M/\mathscr{F}(M)\big)$. The composed map $\xymatrix{M\ar[r]^{\pi\:\:\:\:\:\:\:\:\:\:\:}& M/\mathscr{F}(M)\ar[r]^{\:\:\:\:\:\:\:\delta} &M_{\mathscr{F}}}$ is denoted by $j_{M}$. Therefore for each $m\in M$, $j_{M}(m)=[\delta_{\overline{m}}]$ where $\overline{m}=m+\mathscr{F}(M)$ and $\delta_{\overline{m}}:R\rightarrow M/\mathscr{F}(M)$ is defined as $r\rightsquigarrow rm+\mathscr{F}(M)$.\\ \end{definition} \begin{lemma}\label{G-lemma 3} Let $M$ be a $R-$module. Then $\Ker(j_{M})=\mathscr{F}(M)$ and $\Coker(j_{M})$ is $\mathscr{F}-$negligible.\\ \end{lemma} {\bf Proof.} First note that $\mathscr{F}\big(M/\mathscr{F}(M)\big)=0$ because if $\overline{m}=m+\mathscr{F}(M)\in\mathscr{F}\big(M/\mathscr{F}(M)\big)$ then $\Ann(m):a=\Ann(am)\in\mathscr{F}$ for all $a\in\Ann_{R}(\overline{m})$. Therefore, by Lemma \ref{G-lemma 5}, $\Ann(m)\in\mathscr{F}$ and so $\overline{m}=0$. We have $\Ker(j_{M})=j_{M}^{-1}(0)=\pi^{-1}\big(\Ker(\delta)\big)= \pi^{-1}\Big(\mathscr{F}\big(M/\mathscr{F}(M)\big)\Big)= \Ker(\pi)=\mathscr{F}(M)$. We also have $\Coker(j_{M})=\Coker(\delta)$ therefore $\Coker(j_{M})$ is $\mathscr{F}-$negligible. $\Box$ \\ \begin{lemma}\label{G-lemma 7} Let $M$ be a $R-$module. Then $M$ is $\mathscr{F}-$negligible if and only if $M_{\mathscr{F}}=0$.\\ \end{lemma} {\bf Proof.} If $M$ is $\mathscr{F}-$negligible then clearly $M_{\mathscr{F}}=0$ because the inductive limit of the zero system is zero. Conversely, if $M_{\mathscr{F}}=0$ then $M=\Ker(j_{M})=\mathscr{F}(M)$. $\Box$ \\ Now we prove a very useful result:\\ \begin{proposition}\label{G-lemma 2} Let $f:M\rightarrow N$ and $g:M\rightarrow P$ be $R-$linear maps such that $\Ker(g)$ and $\Coker(g)$ are $\mathscr{F}-$negligible. Then there is a unique $R-$linear map $h:P\rightarrow N_{\mathscr{F}}$ which making commutative the following diagram $$\xymatrix{ M\ar[r]^{f} \ar[d]^{g} & N\ar[d]^{j_{N}\:\:\:\:} \\P\ar[r]^{h\:\:\:\: } & N_{\mathscr{F}}.}$$ \\ \end{proposition} {\bf Proof.} For each $x\in P$ let $J=\Ann_{R}\big(x+\Ima(g)\big)$ which belongs to $\mathscr{F}$ since $\Coker(g)$ is $\mathscr{F}-$negligible. Then consider the map $h_{x}:J\rightarrow N/\mathscr{F}(N)$ which maps each $a\in J$ into $f(m)+\mathscr{F}(N)$ where $m$ is an element of $M$ such that $g(m)=ax$. The map $h_{x}$ is well-defined since $\Ker(g)$ is $\mathscr{F}-$negligible. It is also $R-$linear. Therefore $h_{x}\in\Hom_{R}\big(J, N/\mathscr{F}(N)\big)$. We define $h:P\rightarrow N_{\mathscr{F}}$ as $x\rightsquigarrow [h_{x}]$. Clearly the map $h$ is $R-$linear and the completed diagram is commutative. Suppose $\psi:P\rightarrow N_{\mathscr{F}}$ is another $R-$linear map which making commutative the foregoing diagram. Take $x\in P$, and let $\psi(x)=[h']$ where $h':I\rightarrow N/\mathscr{F}(N)$ is a $R-$linear map for some $I\in\mathscr{F}$. It suffices to show that $h'|_{I\cap J}=(h_{x})|_{I\cap J}$ where $J=\Ann_{R}\big(x+\Ima(g)\big)$. For each $b\in I\cap J$ there is some $m\in M$ such that $g(m)=bx$. But $\psi\big(g(m)\big)=j_{N}\big(f(m)\big)$. Therefore there is an ideal $L\in\mathscr{F}$ contained in $I$ such that for each $c\in L$, $ch'(b)=cf(m)+\mathscr{F}(N)$. Let $h'(b)=n+\mathscr{F}(N)$ then we observe that $L\subseteq\Ann_{R}\big(f(m)-n+\mathscr{F}(N)\big)$. Thus $f(m)-n+\mathscr{F}(N)\in\mathscr{F}\big(N/\mathscr{F}(N)\big)=0$. $\Box$ \\ \begin{corollary}\label{G-coro 1} For each $R-$linear map $u:M\rightarrow N$ then there is a unique $R-$linear map $u_{\mathscr{F}}:M_{\mathscr{F}}\rightarrow N_{\mathscr{F}}$ such that the following diagram is commutative $$\xymatrix{ M\ar[r]^{u} \ar[d]^{j_{M}} & N\ar[d]^{j_{N}\:\:\:\:} \\M_{\mathscr{F}}\ar[r]^{u_{\mathscr{F}}} & N_{\mathscr{F}}.}$$ In particular, $(j_{M})_{\mathscr{F}}=j_{M_{\mathscr{F}}}$.\\ \end{corollary} {\bf Proof.} By Lemma \ref{G-lemma 3}, $\Ker(j_{M})$ and $\Coker(j_{M})$ are $\mathscr{F}-$negligible therefore the first part of the assertion is an immediate consequence of Proposition \ref{G-lemma 2}. Take $N=M_{\mathscr{F}}$ and $u=j_{M}$ then the second part implies. $\Box$ \\ For every $R-$linear map $u:M\rightarrow N$ and for each $[f]\in M_{\mathscr{F}}$, we have $u_{\mathscr{F}}([f])=[\overline{u}\circ f]$ where $\overline{u}:M/\mathscr{F}(M)\rightarrow N/\mathscr{F}(N)$ is induced by $u$.\\ \begin{lemma}\label{G-lemma 6} If $\xymatrix{0\ar[r]&M'\ar[r]^{u}&M\ar[r]^{v}&M''}$ is an exact sequence of $R-$modules then the sequence $\xymatrix{0\ar[r]&M'_{\mathscr{F}}\ar[r]^{u_{\mathscr{F}}}& M_{\mathscr{F}}\ar[r]^{v_{\mathscr{F}}}&M''_{\mathscr{F}}}$ is exact. \\ \end{lemma} {\bf Proof.} Let $[f]\in\Ker(v_{\mathscr{F}})$ where $f:I\rightarrow M/\mathscr{F}(M)$ is a $R-$linear map for some $I\in\mathscr{F}$. Thus there is an ideal $J\in\mathscr{F}$ contained in $I$ such that $\overline{v}\circ f|_{J}=0$. We have $L=J\cap f^{-1}\big(\Ima(\overline{u})\big)\in\mathscr{F}$. Because for each $b\in J$, $\Ann_{R}\big(v(m)\big)\in\mathscr{F}$ where $f(b)=m+\mathscr{F}(M)$. But $\Ann_{R}\big(v(m)\big)\subseteq\Ann_{R}\big(b+L\big)$ and so $J/L$ is $\mathscr{F}-$negligible. Hence $L\in\mathscr{F}.\mathscr{F}=\mathscr{F}$. Now consider the map $g:L\rightarrow M'/\mathscr{F}(M')$ given by $b\rightsquigarrow m'+\mathscr{F}(M')$ where $f(b)=\overline{u}\big(m'+\mathscr{F}(M')\big)$. The map $g$ is well-defined since $\overline{u}$ is injective. It is also $R-$linear. Thus $[g]\in M'_{\mathscr{F}}$. Clearly $f|_{L}=\overline{u}\circ g$ and so $[f]\in\Ima(u_{\mathscr{F}})$. $\Box$ \\ \begin{lemma}\label{G-prop 1} Let $M$ be a $R-$module. Then the canonical map $(j_{M})_{\mathscr{F}}=j_{M_{\mathscr{F}}}:M_{\mathscr{F}} \rightarrow(M_{\mathscr{F}})_{\mathscr{F}}$ is bijective. \\ \end{lemma} {\bf Proof.} By Corollary \ref{G-coro 1}, $(j_{M})_{\mathscr{F}}=j_{M_{\mathscr{F}}}$. By Lemma \ref{G-lemma 6}, from the exact sequence $\xymatrix{0\ar[r]&\mathscr{F}(M)\ar[r]^{i}&M\ar[r]^{j_{M}} &M_{\mathscr{F}}}$ we obtain the following exact sequence $\xymatrix{0\ar[r]& (\mathscr{F}(M))_{\mathscr{F}}\ar[r]^{\:\:\:\:\:\:\:\:\:i_{\mathscr{F}}} &M_{\mathscr{F}}\ar[r]^{(j_{M})_{\mathscr{F}}\:\:\:\:} &(M_{\mathscr{F}})_{\mathscr{F}}}$. By Lemma \ref{G-lemma 7}, $\big(\mathscr{F}(M)\big)_{\mathscr{F}}=0$ hence $j_{M_{\mathscr{F}}}$ is injective thus $\mathscr{F}(M_{\mathscr{F}})=0$. Therefore $(M_{\mathscr{F}})_{\mathscr{F}}= \colimit_{I\in\mathscr{F}}\Hom_{R}(I,M_\mathscr{F})$. Take $[f]\in(M_{\mathscr{F}})_{\mathscr{F}}$ where $f:I\rightarrow M_\mathscr{F}$ is a $R-$linear map and $I\in\mathscr{F}$. We claim that $L=f^{-1}\big(\Ima j_{M}\big)\in\mathscr{F}$. For each $a\in I$ there is an ideal $J\in\mathscr{F}$ and also there is a $R-$linear map $h:J\rightarrow M/\mathscr{F}(M)$ such that $f(a)=[h]$. To prove the claim, first we show that $J\subseteq\Ann_{R}(a+L)$. Let $b\in J$ and let $h(b)=m+\mathscr{F}(M)$ where $m\in M$. We have $f(ab)=bf(a)=[b.h]$. But $b.h=(\delta_{\overline{m}})|_{J}$ therefore $f(ab)=[\delta_{\overline{m}}]\in\Ima j_{M}$. This implies that $I/L$ is $\mathscr{F}-$negligible and so $L\in\mathscr{F}.\mathscr{F}=\mathscr{F}$. This establishes the claim. Now consider the map $g:L\rightarrow M/\mathscr{F}(M)$ which maps each $a\in L$ into $m+\mathscr{F}(M)$ where $f(a)=j_{M}(m)$. The map $g$ is well-defined since $\Ker(j_{M})=\mathscr{F}(M)$. It is also $R-$linear. Hence $[g]\in M_{\mathscr{F}}$. Clearly $f|_{L}=\overline{j_{M}}\circ g$ thus $[f]=j_{M_{\mathscr{F}}}([g])$ and so $j_{M_{\mathscr{F}}}$ is surjective. $\Box$ \\ Now we are ready to prove the first main result of this article:\\ \begin{theorem}\label{G-lemma 10} Let $\phi:M\rightarrow N$ be a $R-$linear map such that $\Ker\phi$ and $\Coker\phi$ are $\mathscr{F}-$negligible. Then $\phi_{\mathscr{F}}:M_{\mathscr{F}}\rightarrow N_{\mathscr{F}}$ is bijective.\\ \end{theorem} {\bf Proof.} By Proposition \ref{G-lemma 2}, there is a (unique) $R-$linear map $h:N\rightarrow M_{\mathscr{F}}$ such that $h\circ\phi=j_{M}$. By Lemma \ref{G-prop 1}, $j_{M_{\mathscr{F}}}$ is bijective. We shall prove that $j^{-1}_{M_{\mathscr{F}}}\circ h_{\mathscr{F}}$ is the inverse of $\phi_{\mathscr{F}}$. We have $j^{-1}_{M_{\mathscr{F}}}\circ h_{\mathscr{F}}\circ\phi_{\mathscr{F}}=j^{-1}_{M_{\mathscr{F}}} \circ(h\circ\phi)_{\mathscr{F}}=\Identity_{M_{\mathscr{F}}}$. To conclude the proof, by Corollary \ref{G-coro 1}, it suffices to show that $\phi_{\mathscr{F}}\circ j^{-1}_{M_{\mathscr{F}}}\circ h_{\mathscr{F}}\circ j_{N}=j_{N}$. But $\phi_{\mathscr{F}}\circ j^{-1}_{M_{\mathscr{F}}}\circ h_{\mathscr{F}}\circ j_{N}=\phi_{\mathscr{F}}\circ j^{-1}_{M_{\mathscr{F}}}\circ j_{M_{\mathscr{F}}}\circ h=\phi_{\mathscr{F}}\circ h$. For each $x\in N$, from the proof of Proposition \ref{G-lemma 2}, we know that $h(x)=[h_{x}]$ where $h_{x}:J=\Ann_{R}(x+\Ima\phi)\rightarrow M/{\mathscr{F}}(M)$ is a $R-$linear map which maps each $a\in J$ into $m+\mathscr{F}(M)$ such that $\phi(m)=ax$. Thus $\overline{\phi}\circ h_{x}=(\delta_{\overline{x}})|_{J}$ where $\overline{x}=x+\mathscr{F}(N)$. This means that $\phi_{\mathscr{F}}\circ h=j_{N}$. $\Box$ \\ Now the G-localization rings are introduced. Let $M$ be a $R-$module. For each $[f]\in R_{\mathscr{F}}$ and for each $[g]\in M_{\mathscr{F}}$ where $f:I\rightarrow R/\mathscr{F}(R)$ and $g:J\rightarrow M/\mathscr{F}(M)$ are $R-$linear maps and $I,J\in\mathscr{F}$, we have $f^{-1}(\overline{J})\in\mathscr{F}$ where $\overline{J}=J+\mathscr{F}(R)/\mathscr{F}(R)$. Moreover $g$ induces a map $\overline{J}\rightarrow R/\mathscr{F}(R)$ given by $b+\mathscr{F}(R)\rightsquigarrow g(b)$ which we denote it by $\overline{g}$. The map $\overline{g}$ is clearly well-defined. Now we define the pairing $R_{\mathscr{F}}\times M_{\mathscr{F}}\rightarrow M_{\mathscr{F}}$ as $[f].[g]=[\overline{g}\circ\big(f|_{f^{-1}(\overline{J})}\big)]$. It is easy to see that it is $R-$bilinear (details omitted). \\ In particular, the binary operation $R_{\mathscr{F}}\times R_{\mathscr{F}}\rightarrow R_{\mathscr{F}}$, as multiplicative, puts a commutative ring structure on the $R-$module $R_{\mathscr{F}}$. Its commutativity implies from the fact that for every $R-$linear maps $f,g:I\rightarrow R/\mathscr{F}(R)$ where $I$ is an ideal of $R$ and for every elements $a,b\in I$ then $\overline{g}\big(f(ab)\big)=\overline{f}\big(g(ab)\big)$. The unit element of the ring $R_{\mathscr{F}}$ is $[\pi]$ where $\pi:R\rightarrow R/\mathscr{F}(R)$ is the canonical map. The map $j_{R}:R\rightarrow R_{\mathscr{F}}$ is a ring homomorphism. Moreover, the pairing $R_{\mathscr{F}}\times M_{\mathscr{F}}\rightarrow M_{\mathscr{F}}$ puts a $R_{\mathscr{F}}-$module structure on $M_{\mathscr{F}}$. For every $R-$linear map $u:M\rightarrow N$ then $u_{\mathscr{F}}:M_{\mathscr{F}}\rightarrow N_{\mathscr{F}}$ is $R_{\mathscr{F}}-$linear. In fact $M\rightsquigarrow M_{\mathscr{F}}$ is a left exact functor from the category of $R-$modules into the category of $R_{\mathscr{F}}-$modules, see Corollary \ref{G-coro 1} and Lemma \ref{G-lemma 6}. It is called the Gabriel localization (G-localization) functor with respect to the system $\mathscr{F}$.\\ \begin{proposition}\label{lemma 189} Let $J$ be an ideal of $R$ such that $JR_{\mathscr{F}}=R_{\mathscr{F}}$. Then $J\in\mathscr{F}$.\\ \end{proposition} {\bf Proof.} We may write $1=\sum\limits_{i=1}^{n}z_{i}j_{R}(a_{i})$ where $a_{i}\in J$ and $z_{i}\in R_{\mathscr{F}}$ for all $i$. Let $f:I\rightarrow N/\mathscr{F}(N)$ be a $R-$linear map where $N=R/J$ and $I\in\mathscr{F}$. Then we have $[f]=\sum\limits_{i=1}^{n}z_{i}.[a_{i}.f]=0$ since $a.f=0$ for all $a\in J$. Therefore $R/J$ is $\mathscr{F}-$negligible and so $J\in\mathscr{F}.\mathscr{F}=\mathscr{F}$. $\Box$ \\ Note that the converse of Proposition \ref{lemma 189} does not necessarily hold. In fact, the condition ``$IR_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$'', as we shall observe in the article, is a crucial point of the G-localization rings. Many interesting and major facts are equivalent or imply from this condition, see for example, Corollary \ref{coro 6533} and Theorem \ref{th621890}.\\ The second main result of this article is the following.\\ \begin{theorem}\label{G-th 5} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$ and let $\phi:R\rightarrow S$ be a ring map such that $IS=S$ for all $I\in\mathscr{F}$. Then the following conditions hold.\\ $\textbf{(i)}$ For each $S-$module $M$, the canonical map $j_{M}:M\rightarrow M_{\mathscr{F}}$ is bijective.\\ $\textbf{(ii)}$ The map $\psi=j^{-1}_{S}\circ\phi_{\mathscr{F}}$ is the only ring homomorphism from $R_{\mathscr{F}}$ into $S$ such that $\phi=\psi\circ j_{R}$.\\ \end{theorem} {\bf Proof.} $\textbf{(i)}:$ If $m\in\mathscr{F}(M)$ then $I=\Ann_{R}(m)\in\mathscr{F}$. Thus we may write $1=\sum\limits_{i}s'_{i}\phi(a_{i})$ where $a_{i}\in I$ and $s'_{i}\in S$. Therefore $m=\sum\limits_{i}s'_{i}\phi(a_{i})m=0$ hence $j_{S}$ is injective. Let $f:J\rightarrow M$ be a $R-$linear map where $J\in\mathscr{F}$. We may write $1=\sum\limits_{j=1}^{n}s_{j}\phi(b_{j})$ where $b_{j}\in J$ and $s_{j}\in S$ for all $j$. For each $c\in J$ we have $f(c)= \sum\limits_{j=1}^{n}s_{j}\phi(b_{j})f(c)=\sum\limits_{j=1}^{n}s_{j}f(b_{j}c)= \big(\sum\limits_{j=1}^{n}s_{j}f(b_{j})\big)\phi(c)$. Therefore $(\delta_{m})|_{J}=f$ where $m=\sum\limits_{j=1}^{n}s_{j}f(b_{j})$ and $\delta_{m}:R\rightarrow M$ which maps each $r\in R$ into $r.m=\phi(r)m$. This means that $j_{M}(m)=[f]$ and so $j_{M}$ is surjective. \\ $\textbf{(ii)}:$ First we show that the map $\psi=j_{S}^{-1}\circ\phi_{\mathscr{F}}$ is actually a ring homomorphism. Clearly it is additive and transforms the unit element of $R_{\mathscr{F}}$ to the unit of $S$. To prove that it is multiplicative take two elements $z,z'\in R_{\mathscr{F}}$ with representations $f$ and $g$, i.e., $z=[f]$ and $z'=[g]$. By passing to the restrictions, if it is necessary, therefore we may assume that the $R-$linear maps $f$ and $g$ have the same domain. Namely $f,g:J\rightarrow R/\mathscr{F}(R)$ where $J\in\mathscr{F}$. We know that $f^{-1}(\overline{J})\in\mathscr{F}$ where $\overline{J}=J+\mathscr{F}(R)/\mathscr{F}(R)$. Therefore we may write $1=\sum\limits_{j}\phi(b_{j})s_{j}$ where $b_{j}\in f^{-1}(\overline{J})\subseteq J$ and $s_{j}\in S$. We have then $\psi(z.z')= \sum\limits_{j}\phi(c'_{j})s_{j}$ where for each $j$, $f(b_{j})=c_{j}+\mathscr{F}(R)$, $g(c_{j})=c'_{j}+\mathscr{F}(R)$ and $c_{j}\in J$. Similarly $\psi(z)=\sum\limits_{j}\phi(c_{j})s_{j}$ and $\psi(z')=\sum\limits_{i}\phi(e_{i})s_{i}$ where for each $i$, $g(b_{i})=e_{i}+\mathscr{F}(R)$. Thus $\psi(z) \psi(z')= \sum\limits_{j}\big(\sum\limits_{i}\phi(e_{i}c_{j})s_{i}\big)s_{j}$. But for each $j$ we have $\phi(c'_{j})=\sum\limits_{i}\phi(c'_{j}b_{i})s_{i}= \sum\limits_{i}\overline{\phi}\big(c'_{j}b_{i}+\mathscr{F}(R)\big)s_{i}= \sum\limits_{i}\overline{\phi}\big(b_{i}g(c_{j})\big)s_{i}= \sum\limits_{i}\overline{\phi}\big(c_{j}g(b_{i})\big)s_{i}= \sum\limits_{i}\overline{\phi}\big(e_{i}c_{j}+\mathscr{F}(R)\big)s_{i}= \sum\limits_{i}\phi(e_{i}c_{j})s_{i}$. Therefore $\psi(z.z')=\psi(z) \psi(z')$. Finally, suppose $\psi':R_{\mathscr{F}}\rightarrow S$ is another ring map such that $\psi'\circ j_{R}=\phi$. Clearly $j_{S}\circ\psi'$ is $R-$linear and $(j_{S}\circ\psi')\circ j_{R}=j_{S}\circ\phi$. Therefore, by Corollary \ref{G-coro 1}, $j_{S}\circ\psi'=\phi_{\mathscr{F}}$ and so $\psi'=j_{S}^{-1}\circ\phi_{\mathscr{F}}$. $\Box$ \\ \section{Flat epimorphisms as G-localizations} By an epimorphism $\phi:R\rightarrow S$ we mean it is an epimorphism in the category of commutative rings. We should mention that the surjective ring maps are just special cases of the epimorphisms. For example, the canonical ring map $\mathbb{Z}\rightarrow\mathbb{Q}$ is an epimorphism while it is not surjective. A ring map which is both flat and an epimorphism is called a flat epimorphism. The canonical map $R\rightarrow S^{-1}R$ where $S$ is a multiplicative subset of a ring $R$ is a typical example of flat epimorphisms.\\ The following lemma is a well-known result but we have provided a proof for the sake of completeness. \\ \begin{lemma}\label{G-lemma 9} Let $\phi:R\rightarrow S$ be a ring homomorphism. Then the following conditions are equivalent.\\ $\mathbf{(i)}$ $\phi$ is an epimorphism.\\ $\mathbf{(ii)}$ For each $s\in S$, $s\otimes1=1\otimes s$.\\ $\mathbf{(iii)}$ The map $p: S\otimes_{R}S\rightarrow S$ defined by $s\otimes s'\rightsquigarrow ss'$ is bijective.\\ $\mathbf{(iv)}$ The map $j: S\rightarrow S\otimes_{R}S$ defined by $s\rightsquigarrow 1\otimes s$ is bijective.\\ \end{lemma} {\bf Proof.} $\mathbf{(i)}\Rightarrow\mathbf{(ii):}$ We have $i\circ\phi=j\circ\phi$ where $i,j:S\rightarrow S\otimes_{R}S$ are the canonical ring maps which map each $s\in S$ into $s\otimes1$ and $1\otimes s$, respectively. \\ $\mathbf{(ii)}\Rightarrow\mathbf{(iii):}$ We have $\Ker(p)=\langle s\otimes1-1\otimes s : s\in S\rangle$ because if $ss'=0$ then we may write $s\otimes s'=1\otimes s'(s\otimes1- 1\otimes s)$.\\ $\mathbf{(iii)}\Rightarrow\mathbf{(i):}$ Let $f,g:S\rightarrow T$ be ring maps such that $f\circ\phi=g\circ\phi$. By the universal property of the pushouts, there is a (unique) ring map $\psi:S\otimes_{R}S\rightarrow T$ such that $f=\psi\circ i$ and $g=\psi\circ j$. But $i=j$ since $\Ker(p)=0$. Thus $f=g$.\\ $\mathbf{(ii)}\Rightarrow\mathbf{(iv):}$ We have $s\otimes s'=(s\otimes1)(1\otimes s')=1\otimes ss'=j(ss')$.\\ $\mathbf{(iv)}\Rightarrow\mathbf{(iii):}$ We have $p\circ j=\Identity$. Therefore $p$ is bijective. $\Box$ \\ The third main result of this article is the following.\\ \begin{theorem}\label{G-lemma 1} Let $\phi:R\rightarrow S$ be a ring map and let $\mathscr{F}$ be the set of ideals of $R$ whose extensions under $\phi$ are equal to $S$. Then the following conditions hold.\\ $\textbf{(i)}$ The family $\mathscr{F}$ is an idempotent toplogizing system on the ring $R$. \\ $\textbf{(ii)}$ If $\phi$ is a flat epimorphism then for each $R-$module $M$, the map $\eta_{\mathscr{F}}:M_{\mathscr{F}} \rightarrow(S\otimes_{R}M)_{\mathscr{F}}$ induced by the canonical map $\eta:M\rightarrow S\otimes_{R}M$ given by $m\rightsquigarrow1\otimes m$ is an isomorphism. In particular the map $\phi_{\mathscr{F}}$ is bijective.\\ \end{theorem} {\bf Proof.} $\textbf{(i)}:$ Clearly $\mathscr{F}$ is non-empty since $R\in\mathscr{F}$. Suppose $I\in\mathscr{F}$ and $J$ is an ideal of $R$ such that $J:a\in\mathscr{F}$ for all $a\in I$. We may write $1=\sum\limits_{j=1}^{n}s_{j}\phi(a_{j})$ where $a_{j}\in I$ and $s_{j}\in S$. For each $j$, we may also write $1=\sum\limits_{i_{j}=1}^{k_{j}}s_{i_{j},j}\phi(b_{i_{j},j})$ where $b_{i_{j},j}\in J:a_{j}$ and $s_{i_{j},j}\in S$. We have then $1=\sum\limits_{i_{1}=1}^{k_{1}}...\sum\limits_{i_{n}=1}^{k_{n}} \sum\limits_{j=1}^{n}s_{i_{1},1}...s_{i_{n},n}s_{j} \phi(b_{i_{1},1}...b_{i_{n},n}a_{j})$ and all the elements $b_{i_{1},1}...b_{i_{n},n}a_{j}$ belong to $J$ and so $J\in\mathscr{F}$. Therefore, by Lemma \ref{G-lemma 5}, $\mathscr{F}$ is an idempotent topologizing system. \\ $\textbf{(ii)}:$ To prove the assertion, by Theorem \ref{G-lemma 10}, it suffices to show that $\Ker\eta$ and $\Coker\eta$ are $\mathscr{F}-$negligible. Let $m\in\Ker\eta$ then the map $\lambda:S\otimes_{R}N\rightarrow S\otimes_{R}M$ induced by the canonical injection $N=Rm\rightarrow M$ is injective since $S$ is $R-$flat. But $\Ima\lambda=0$ because $1\otimes m=0$. This implies that $\Ann_{R}(m)S=S$ and so $\Ann_{R}(m)\in\mathscr{F}$. Thus $\Ker\eta$ is $\mathscr{F}-$negligible. \\ By applying the right exact functor $S\otimes_{R}-$ to the exact sequence $\xymatrix{M\ar[r]^{\eta\:\:\:\:\:\:\:\:}&S\otimes_{R} M\ar[r]&\Coker\eta\ar[r]&0}$ we then obtain the following exact sequence $\xymatrix{S\otimes_{R}M\ar[r]^{1\otimes\eta\:\:\:\:\:\:\:\:\:\:\:\:}&S\otimes_{R} (S\otimes_{R}M)\ar[r]&S\otimes_{R}\Coker\eta\ar[r]&0.}$ \\ The map $1\otimes\eta$ factors as $\xymatrix{S\otimes_{R}M\ar[r]^{j\otimes 1\:\:\:\:\:\:\:\:\:\:\:\:}& (S\otimes_{R}S)\otimes_{R}M\ar[r]^{\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\:\simeq}&}$\\ $\xymatrix{S\otimes_{R}(S\otimes_{R}M).}$ By Lemma \ref{G-lemma 9}, $j\otimes1$ is bijective, hence so is $1\otimes\eta$. In particular it is surjective and so $S\otimes_{R}\Coker\eta=0$. For each $x\in\Coker\eta$, the map $S\otimes_{R}Rx\rightarrow S\otimes_{R}\Coker\eta$ induced by the canonical injection $Rx\rightarrow\Coker\eta$ is injective since $S$ is $R-$flat. This implies that $\Ann_{R}(x)S=S$ and so $\Ann_{R}(x)\in\mathscr{F}$. Therefore $\Coker\eta$ is also $\mathscr{F}-$negligible. Finally, the map $\phi$ factors as $\xymatrix{R\ar[r]^{\eta\:\:\:\:\:\:\:}&R\otimes_{R}S\ar[r]^{\:\:\:\:\:\simeq}&S}$ hence $\phi_{\mathscr{F}}$ is bijective. $\Box$ \\ \begin{corollary}\label{G-coro 2} Let $\phi:R\rightarrow S$ be a ring map and let $\mathscr{F}$ be the set of ideals $I$ of $R$ such that $IS=S$. If $\phi$ is a flat epimorphism then there is a unique isomorphism of rings $\psi:R_{\mathscr{F}}\rightarrow S$ such that $\phi=\psi\circ j_{R}$. \\ \end{corollary} {\bf Proof.} By Theorem \ref{G-th 5}, the map $\psi=j^{-1}_{S}\circ\phi_{\mathscr{F}}$ is the only ring homomorphism such that $\phi=\psi\circ j_{R}$. By Theorem \ref{G-lemma 1}, $\phi_{\mathscr{F}}$ is bijective, therefore $\psi$ is an isomorphism. $\Box$ \\ \begin{corollary}\label{coro 87} Let $J$ be an ideal of a ring $R$. Then $R/J$ is $R-$flat if and only if $\Ann_{R}(a)+J=R$ for all $a\in J$.\\ \end{corollary} {\bf Proof.} Suppose $R/J$ is $R-$flat. Let $\mathscr{F}$ be the set of ideals $I$ of $R$ such that $I+J=R$. By Corollary \ref{G-coro 2}, there is a (unique) isomorphism of rings $\psi: R/J\rightarrow R_{\mathscr{F}}$ such that $j_{R}=\psi\circ\pi$ where $\pi:R\rightarrow R/J$ is the canonical map. We have $\mathscr{F}(R)=\Ker(j_{R})=j_{R}^{-1}(0)=\pi^{-1}\big(\psi^{-1}(0)\big) =\Ker(\pi)=J$. Conversely, let $f:M\rightarrow N$ be an injective $R-$linear map. To prove the assertion it suffices to show that the induced map $M/JM\rightarrow N/JN$ given by $m+JM\rightsquigarrow f(m)+JN$ is injective. If $f(m)\in JN$ then we may write $f(m)=\sum\limits_{i=1}^{s}a_{i}n_{i}$ where $a_{i}\in J$ and $n_{i}\in N$ for all $i$. By the hypothesis, there are elements $b_{i}\in\Ann_{R}(a_{i})$ and $c_{i}\in J$ such that $1=b_{i}+c_{i}$. It follows that $1=(b_{1}+c_{1})(b_{2}+c_{2})...(b_{s}+c_{s})=b+c$ where $b=b_{1}b_{2}...b_{s}$ and $c\in J$. Thus $f(m)=bf(m)+cf(m)=f(cm)$. Therefore $m=cm\in JM$. $\Box$ \\ Corollary \ref{coro 87}, in particular, tells us that if the ideal $J$ has a generating set $S$ such that each $a\in S$ can be written as $a=a^{2}b$ for some $b\in R$ then $R/J$ is $R-$flat. As another application, if $J$ is an ideal of a domain $R$ such that $R/J$ is $R-$flat then we have either $J=0$ or $J=R$.\\ \begin{lemma}\label{coro 652} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$ such that $IR_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$. Then $j_{R}$ is an epimorphism.\\ \end{lemma} {\bf Proof.} Let $f,g:R_{\mathscr{F}}\rightarrow S$ be two ring maps such that $f\circ j_{R}=g\circ j_{R}$. If we consider the ring map $\phi=f\circ j_{R}:R\rightarrow S$ then $IS=S$ for all $I\in\mathscr{F}$. Therefore, by Theorem \ref{G-th 5}, there is a unique ring map $\psi:R_{\mathscr{F}}\rightarrow S$ such that $\phi=\psi\circ j_{R}$. Thus $f=g$. $\Box$ \\ The converse of Corollary \ref{G-coro 2} also holds even under a mild hypothesis which is another main result of this article:\\ \begin{theorem}\label{remark 200} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$ and let $\phi:R\rightarrow S$ be a ring map such that $IS=S$ for all $I\in\mathscr{F}$. If there is an isomorphism of rings $\psi:R_{\mathscr{F}}\rightarrow S$ such that $\phi=\psi\circ j_{R}$ then $\phi:R\rightarrow S$ is a flat epimorphism and $\psi=j_{S}^{-1}\circ\phi_{\mathscr{F}}$.\\ \end{theorem} {\bf Proof.} We have $IR_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$. Therefore, by Lemma \ref{coro 652}, $j_{R}$ and so $\phi$ are epimorphisms. By Theorem \ref{G-th 5}, $\psi=j_{S}^{-1}\circ\phi_{\mathscr{F}}$. Therefore $\phi_{\mathscr{F}}$ is bijective. This, in particular, implies that $j_{S}(s)\in\Ima\phi_{\mathscr{F}}$ for all $s\in S$. Thus there is an ideal $I\in\mathscr{F}$ such that $I\subseteq\Ann_{R}(s+\Ima\phi)$. Therefore $\Coker\phi$ and so any finite direct sum of it are $\mathscr{F}-$negligible. To prove that $S$ is $R-$flat, by \cite[Theorem 7.7]{Matsumura}, it suffices to show that for each ideal $J$ of $R$ then the canonical map $J\otimes_{R}S\rightarrow S$ given by $\sum\limits_{i=1}^{n}a_{i}\otimes s_{i}\rightsquigarrow\sum\limits_{i=1}^{n}\phi(a_{i})s_{i}$ is injective. Suppose $\sum\limits_{i=1}^{n}\phi(a_{i})s_{i}=0$ where $a_{i}\in J$ and $s_{i}\in S$ for all $i$. We have $\Ann_{R}(x)\in\mathscr{F}$ where $x=(s_{i}+\Ima\phi)^{n}_{i=1}\in(\Coker\phi)^{n}$ since $(\Coker\phi)^{n}$ is $\mathscr{F}-$negligible. Thus there are elements $b_{1},...,b_{m}\in\Ann_{R}(x)$ and also elements $s'_{1},...,s'_{m}\in S$ such that $1=\sum\limits_{j=1}^{m}\phi(b_{j})s'_{j}$. Moreover there are elements $r_{i,j}\in R$ such that $s_{i}\phi(b_{j})=\phi(r_{i,j})$ for all $i,j$. Thus $c_{j}=\sum\limits_{i=1}^{n}a_{i}r_{i,j}\in\Ker\phi=\mathscr{F}(R)$ for all $j$. Hence for each $j$ there are elements $r'_{j,1},...,r'_{j,N}\in\Ann_{R}(c_{j})$ and also elements $s''_{j,1},...,s''_{j,N}\in S$ such that $1=\sum\limits_{k=1}^{N}\phi(r'_{j,k})s''_{j,k}$. Now we have $\sum\limits_{i=1}^{n}a_{i}\otimes s_{i}=\sum\limits_{i=1}^{n}a_{i}\otimes\big( \sum\limits_{j=1}^{m}s_{i}\phi(b_{j})s'_{j}\big)= \sum\limits_{j=1}^{m}c_{j}\otimes s'_{j}$. For each $j$, $c_{j}\otimes s'_{j}=c_{j}\otimes \big(\sum\limits_{k=1}^{N}\phi(r'_{j,k})s'_{j}s''_{j,k}\big)= \sum\limits_{k=1}^{N}r'_{j,k}c_{j}\otimes s'_{j}s''_{j,k}=0$. $\Box$ \\ \begin{corollary}\label{coro 6533} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$ such that $IR_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$. Then $j_{R}$ is a flat epimorphism.\\ \end{corollary} {\bf Proof.} It is an immediate consequence of Theorem \ref{remark 200}. $\Box$ \\ It is natural to ask whether the converse of Corollary \ref{coro 6533} holds.\\ \section{G-localizations of finite type systems} \begin{definition} An $R-$module $M$ is said to be $\mathscr{F}-$closed if the canonical map $j_{M}$ is bijective. It is called strongly $\mathscr{F}-$closed if the canonical map $M\rightarrow\Hom_{R}(I,M)$ given by $m\rightsquigarrow(\delta_{m})|_{I}$ is bijective for all $I\in\mathscr{F}$. \\ \end{definition} For each $R-$module $M$ then $M_{\mathscr{F}}$, by Lemma \ref{G-prop 1}, is $\mathscr{F}-$closed. Clearly each strongly $\mathscr{F}-$closed module is $\mathscr{F}-$closed. By the category of $\mathscr{F}-$closed modules we mean a full subcategory of the category of $R-$modules whose objects are the $\mathscr{F}-$closed modules. \\ \begin{lemma}\label{prop 431} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$ such that $IR_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$. Then every $\mathscr{F}-$closed module is strongly $\mathscr{F}-$closed and the G-localization functor with respect to $\mathscr{F}$ is an equivalence between the category of $\mathscr{F}-$closed modules and the category of $R_{\mathscr{F}}-$modules. \\ \end{lemma} {\bf Proof.} Let $N$ be a $\mathscr{F}-$closed module and let $I\in\mathscr{F}$. Then the canonical map $N\rightarrow\Hom_{R}(I,N)$ is injective since $\mathscr{F}(N)=0$. Let $f:I\rightarrow N$ be a $R-$linear map. We may write $1=\sum\limits_{i=1}^{n}j_{R}(a_{i})z_{i}$ where $a_{i}\in I$ and $z_{i}\in R_{\mathscr{F}}$ for all $i$. There exists an element $x\in N$ such that $j_{N}(x)=\sum\limits_{i=1}^{n}z_{i}j_{N}\big(f(a_{i})\big)$. Now, for each $a\in I$, we have $j_{N}(ax)=j_{N}\big(f(a)\big)$ and so $ax=f(a)$. Thus $(\delta_{x})|_{I}=f$. Therefore $N$ is strongly $\mathscr{F}-$closed. Moreover, for each $R-$module $M$ then the map $\Hom_{R}(M,N)\rightarrow\Hom_{R_{\mathscr{F}}} (M_{\mathscr{F}},N_{\mathscr{F}})$ given by $u\rightsquigarrow u_{\mathscr{F}}$ is bijective. Because suppose $u_{\mathscr{F}}=u'_{\mathscr{F}}$ where $u,u':M\rightarrow N$ are $R-$linear maps. We have $u=j_{N}^{-1}\circ u_{\mathscr{F}}\circ j_{M}=u'$. Let $\phi:M_{\mathscr{F}}\rightarrow N_{\mathscr{F}}$ be a $R-$linear map. Then, by Corollary \ref{G-coro 1}, $u_{\mathscr{F}}=\phi$ where $u=j_{N}^{-1}\circ\phi\circ j_{M}$. Finally, we show that the G-localization functor is essentially surjective. Let $L$ be a $R_{\mathscr{F}}-$module. By Theorem \ref{G-th 5}, $j_{L}$ is bijective. Thus $L$ as $R-$module is $\mathscr{F}-$closed. It remains to show that $j_{L}$ is $R_{\mathscr{F}}-$linear. We have $j_{L}=\sigma_{L}\circ\eta$ where $\eta:L\rightarrow R_{\mathscr{F}}\otimes_{R}L$ is the canonical map. The map $\sigma_{L}$ is already $R_{\mathscr{F}}-$linear. By Lemma \ref{coro 652}, $j_{R}$ is an epimorphism. Note that for any epimorphism of rings $\phi:R\rightarrow S$ and for any $S-$modules $M$ and $N$ then the two $S-$module structures on $M\otimes_{R}N$ defined on pure tensors by $s.(m\otimes n)=sm\otimes n$ and $s\ast(m\otimes n)=m\otimes sn$ are the same since $s\otimes1=1\otimes s$ for all $s\in S$ and so the canonical map $\eta:N\rightarrow S\otimes_{R}N$ is $S-$linear. Therefore $j_{L}$ is $R_{\mathscr{F}}-$linear. $\Box$ \\ \begin{definition} An idempotent topologizing system is called \emph{of finite type} if every element of it containing a finitely generated ideal belonging to the system. For example, if $\phi:R\rightarrow S$ is a ring map then the system $\{I\subseteq R : IS=S\}$ is of finite type.\\ \end{definition} In the next result we shall use, for each $R-$module $M$, the canonical $R_{\mathscr{F}}-$linear map $\sigma_{M}:R_{\mathscr{F}}\otimes_{R}M\rightarrow M_{\mathscr{F}}$ which maps each pure tensor $a\otimes m$ into $a.j_{M}(m)$ for all $a\in R_{\mathscr{F}}$ and all $m\in M$. Note that $\sigma_{R}$ is bijective and for each $R-$linear map $\phi:M\rightarrow N$ then the following diagram is commutative $$\xymatrix{ R_{\mathscr{F}}\otimes_{R}M\ar[r]^{1\otimes\phi} \ar[d]^{\sigma_{M}} & R_{\mathscr{F}}\otimes_{R}N\ar[d]^{\sigma_{N}} \\ M_{\mathscr{F}}\ar[r]^{\:\:\:\:\phi_{\mathscr{F}}} & N_{\mathscr{F}}.}$$ \\ The following is another main result of this article:\\ \begin{theorem}\label{th621890} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$. Then the following conditions are equivalent.\\ $\textbf{(i)}$ For each $R-$module $M$, the canonical map $\sigma_{M}:R_{\mathscr{F}}\otimes_{R}M\rightarrow M_{\mathscr{F}}$ is bijective.\\ $\textbf{(ii)}$ For every $R-$module $M$, $\mathscr{F}(M)$ is the kernel of the canonical map $\eta:M\rightarrow R_{\mathscr{F}}\otimes_{R}M$.\\ $\textbf{(iii)}$ For each $I\in\mathscr{F}$, $IR_{\mathscr{F}}=R_{\mathscr{F}}$. \\ $\textbf{(iv)}$ The G-localization functor with respect to $\mathscr{F}$ is exact and preserves direct sums.\\ $\textbf{(v)}$ The G-localization functor with respect to $\mathscr{F}$ is exact and the system $\mathscr{F}$ is of finite type.\\ $\textbf{(vi)}$ The G-localization functor with respect to $\mathscr{F}$ is essentially surjective.\\ $\textbf{(vii)}$ For each $R_{\mathscr{F}}-$module $N$, $\mathscr{F}(N)=0$.\\ \end{theorem} {\bf Proof.} $\textbf{(i)}\Rightarrow\textbf{(ii)}:$ We have $j_{M}=\sigma_{M}\circ\eta$. Thus $\mathscr{F}(M)=\Ker j_{M}=j_{M}^{-1}(0)=\eta^{-1}\big(\sigma_{M}^{-1}(0)\big)=\Ker\eta$.\\ $\textbf{(ii)}\Rightarrow\textbf{(iii)}:$ If $I\in\mathscr{F}$ then $R/I$ is $\mathscr{F}-$negligible. Thus $\Ker\eta=R/I$ where $\eta:R/I\rightarrow R_{\mathscr{F}}\otimes_{R}R/I$ is the canonical map. Therefore $\Ima\eta=0$. This implies that $R_{\mathscr{F}}\otimes_{R}R/I=0$ since $z\otimes(r+I)=\big(z\otimes(1+I)\big)\big(1\otimes(r+I)\big)=0$. It follows that $IR_{\mathscr{F}}=R_{\mathscr{F}}$. \\ $\textbf{(iii)}\Rightarrow\textbf{(i)}:$ By Corollary \ref{coro 6533}, $j_{R}$ is a flat epimorphism. Moreover, by Proposition \ref{lemma 189}, $\mathscr{F}=\{I\subseteq R : IR_{\mathscr{F}}=R_{\mathscr{F}}\}$. Therefore, by the proof of Theorem \ref{G-lemma 1}, $\Ker\eta$ and $\Coker\eta$ are $\mathscr{F}-$negligible where $\eta:M\rightarrow R_{\mathscr{F}}\otimes_{R}M$ is the canonical map. Thus, by Proposition \ref{G-lemma 2}, there is a unique $R-$linear map $h: R_{\mathscr{F}}\otimes_{R}M\rightarrow M_{\mathscr{F}}$ such that $j_{M}=h\circ\eta$. By Theorem \ref{G-th 5}, $j_{N}$ is bijective where $N=R_{\mathscr{F}}\otimes_{R}M$. Moreover, by Theorem \ref{G-lemma 1}, $\eta_{\mathscr{F}}$ is bijective. We also have $j_{M}=(\eta_{\mathscr{F}}^{-1}\circ j_{N})\circ\eta$. Therefore $\sigma_{M}=\eta_{\mathscr{F}}^{-1}\circ j_{N}$ and so $\sigma_{M}$ is bijective.\\ $\textbf{(i)}\Rightarrow\textbf{(iv)}$ and $\textbf{(v)}:$ Easy.\\ $\textbf{(iv)}\Rightarrow\textbf{(i)}:$ Let $L$ be a free $R-$module and let $B=\{x_{i} : i\in I\}$ be a basis of $L$. Then there is a bijective map $\psi:L\rightarrow\bigoplus\limits_{i\in I}R$ which maps each $x\in L$ into $(r_{i})_{i\in I}$ where $x=\sum\limits_{i\in I}r_{i}x_{i}$. We denote the converse of $\psi$ by $\theta$ and for a given $R-$module $M$, $\bigoplus\limits_{i\in I}M$ is denoted by $M^{\oplus I}$. We claim that the map $\sigma_{L}$ factors as $$\xymatrix{ R_{\mathscr{F}}\otimes_{R}L\ar[r]^{\simeq\:\:\:\:\:\:} &(R_{\mathscr{F}}\otimes_{R}R)^{\oplus I} \ar[r]^{\:\:\:\:\:\:\:\:\:(\sigma_{R})_{I}}& (R_{\mathscr{F}})^{\oplus I}\ar[r]^{\tau^{(I)}}&(R^{\oplus I})_{\mathscr{F}}\ar[r]^{\:\:\:\:\:\:\theta_{\mathscr{F}}}&L_{\mathscr{F}}}$$ where $\tau^{(I)}\big((z_{i})_{i\in I}\big)=\sum\limits_{i\in I}(\tau_{i})_{\mathscr{F}}(z_{i})$ and for each $i$, $\tau_{i}:R\rightarrow R^{\oplus I}=\bigoplus\limits_{j\in I}R$ is the canonical map which is defined as $r\rightsquigarrow(r\delta_{i,j})_{j\in I}$. In fact, for each pure tensor $z\otimes x$ of $R_{\mathscr{F}}\otimes_{R}L$ we have $z\otimes x\rightsquigarrow(z\otimes r_{i})_{i\in I}\rightsquigarrow(r_{i}.z)_{i\in I}\rightsquigarrow\xi\rightsquigarrow\theta_{\mathscr{F}}(\xi)$ where $\xi=\sum\limits_{i\in I}(\tau_{i})_{\mathscr{F}}(r_{i}.z)$. But $\theta_{\mathscr{F}}(\xi)=z\theta_{\mathscr{F}}\big(\sum\limits_{i\in I}(\tau_{i})_{\mathscr{F}}\circ j_{R}(r_{i})\big)=z\sum\limits_{i\in I}\theta_{\mathscr{F}}\circ j_{(R^{\oplus I})}\circ\tau_{i}(r_{i})=z\sum\limits_{i\in I}j_{L}\circ\theta\circ\tau_{i}(r_{i}) =zj_{L}\Big(\theta\big(\sum\limits_{i\in I}\tau_{i}(r_{i})\big)\Big)=zj_{L}(x)=\sigma_{L}(z\otimes x)$. This establishes the claim. The map $\tau^{(I)}$ is bijective since the G-localization functor preserves direct sums. Therefore $\sigma_{L}$ is bijective. Now, by applying the tensor functor $R_{\mathscr{F}}\otimes_{R}-$ and the G-localization functor which is, by the hypotheses, right exact to the exact sequence $\xymatrix{L'\ar[r]&L\ar[r]&M\ar[r]&0}$ where $L$ and $L'$ are free $R-$modules, and also by using a special case of the five lemma we conclude that $\sigma_{M}$ is bijective: \begin{displaymath} \xymatrix{ R_{\mathscr{F}}\otimes_{R}L'\ar[r]^{} \ar[d]_{\sigma_{L'}}&R_{\mathscr{F}}\otimes_{R}L \ar[r]\ar[d]_{\sigma_{L}}& R_{\mathscr{F}}\otimes_{R}M\ar[r]\ar[d]_{\sigma_{M}}&0\\ L'_{\mathscr{F}}\ar[r]&L_{\mathscr{F}}\ar[r]& M_{\mathscr{F}}\ar[r]&0.}\\ \end{displaymath} $\textbf{(v)}\Rightarrow\textbf{(iii)}:$ The G-localization functor, by Corollary \ref{G-coro 1}, is additive. In every abelian category, each split and exact sequence is left split and exact by an additive functor. This, in particular, implies that the G-localization functor preserves finite direct sums. Specially, $\tau^{(n)}:(R_{\mathscr{F}})^{n}\rightarrow(R^{n})_{\mathscr{F}}$ is an isomorphism as $R_{\mathscr{F}}-$modules for all natural numbers $n$. Therefore all of the factors in the decomposition of $\sigma_{R^{n}}$ are bijective. Let $I$ be a f.g. ideal of $R$. Now, by applying the tensor functor $R_{\mathscr{F}}\otimes_{R}-$ and the G-localization functor which is, by the hypotheses, right exact to the exact sequence $\xymatrix{R^{n}\ar[r]&R\ar[r]&R/I\ar[r]&0}$ and also by using a special case of the five lemma we obtain that $\sigma_{R/I}$ is bijective: \begin{displaymath} \xymatrix{ R_{\mathscr{F}}\otimes_{R}R^{n}\ar[r]^{} \ar[d]_{\sigma_{R^{n}}}&R_{\mathscr{F}}\otimes_{R}R \ar[r]\ar[d]_{\sigma_{R}}& R_{\mathscr{F}}\otimes_{R}R/I\ar[r]\ar[d]_{\sigma_{R/I}}&0\\ (R^{n})_{\mathscr{F}}\ar[r]&R_{\mathscr{F}}\ar[r]& (R/I)_{\mathscr{F}}\ar[r]&0.} \end{displaymath} If moreover $I\in\mathscr{F}$ then $0=(R/I)_{\mathscr{F}}\simeq R_{\mathscr{F}}\otimes_{R}R/I\simeq R_{\mathscr{F}}/IR_{\mathscr{F}}$. Therefore $IR_{\mathscr{F}}=R_{\mathscr{F}}$. In fact the latter holds for each element of the system $\mathscr{F}$ since it is of finite type.\\ $\textbf{(iii)}\Rightarrow\textbf{(vi)}:$ See Lemma \ref{prop 431}.\\ $\textbf{(vi)}\Rightarrow\textbf{(vii)}:$ Let $N$ be a $R_{\mathscr{F}}-$module. By the hypothesis, there exists a $R-$module $M$ such that $N\simeq M_{\mathscr{F}}$. Therefore $\mathscr{F}(N)\simeq\mathscr{F}(M_{\mathscr{F}})=0$.\\ $\textbf{(vii)}\Rightarrow\textbf{(iii)}:$ If $I\in\mathscr{F}$ then $R_{\mathscr{F}}\otimes_{R}R/I=0$ since $I\subseteq\Ann_{R}\big(z\otimes(r+I)\big)$ thus $z\otimes(r+I)\in\mathscr{F}(R_{\mathscr{F}}\otimes_{R}R/I)=0$. Therefore $IR_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$. $\Box$ \\ \begin{proposition}\label{G-prop 2} Let $\phi:R\rightarrow S$ be a ring map, let $\mathscr{F}$ be an of finite type system on the ring $R$ and let $\mathscr{G}$ be the set of ideals $J$ of $S$ such that $S/J$ as $R-$module is $\mathscr{F}-$negligible. Then the following conditions hold.\\ $\textbf{(i)}$ An ideal $J$ of $S$ belongs to $\mathscr{G}$ if and only if $IS\subseteq J$ for some $I\in\mathscr{F}$.\\ $\textbf{(ii)}$ The family $\mathscr{G}$ is an idempotent topologizing system on the ring $S$ which is also of finite type. \\ $\textbf{(iii)}$ For each $S-$module $M$, $\mathscr{F}(M)=\mathscr{G}(M)$ and there is a canonical isomorphism of $R-$modules $\eta:M_{\mathscr{G}}\rightarrow M_{\mathscr{F}}$ such that $j_{M}=\eta\circ j'_{M}$ where $j'_{M}:M\rightarrow M_{\mathscr{G}}$ is the canonical map.\\ $\textbf{(iv)}$ The map $\phi_{\mathscr{F}}:R_{\mathscr{F}}\rightarrow S_{\mathscr{F}}$ is a ring homomorphism when we identify $S_{\mathscr{F}}$ with $S_{\mathscr{G}}$ via $\eta$.\\ \end{proposition} {\bf Proof.} $\textbf{(i)}:$ Suppose $J$ is an ideal of $S$ and $IS\subseteq J$ for some $I\in\mathscr{F}$. Let $s\in S$ we have then $I\subseteq\Ann_{R}(s+J)$. Therefore $S/J$ is $\mathscr{F}-$negligible. Conversely, suppose $J\in\mathscr{G}$. Then $I=\Ann_{R}(1_{S}+J)\in\mathscr{F}$ and clearly $IS\subseteq J$.\\ $\textbf{(ii)}:$ Clearly $\mathscr{G}$ is a topologizing system on the ring $S$, it is also of finite type once we have verified that it is idempotent. Let $J\in\mathscr{G}.\mathscr{G}$ then there exists an ideal $J'\in\mathscr{G}$ containing $J$ such that $J'/J$ is $\mathscr{G}-$negligible. By the hypothesis, there is a finitely generated ideal $I=\langle a_{1},...,a_{n}\rangle$ of $R$ such that $I\in\mathscr{F}$ and $IS\subseteq J'$. For each $j$, there is also an ideal $I_{j}\in\mathscr{F}$ such that $(I_{j}a_{j})S\subseteq J$. Let $I'=I_{1}I_{2}...I_{n}$ then clearly $II'\in\mathscr{F}$ and $II'S\subseteq J$. This means that $J\in\mathscr{G}$.\\ $\textbf{(iii)}:$ If $m\in\mathscr{F}(M)$ then $I=\Ann_{R}(m)\in\mathscr{F}$ and $IS\subseteq\Ann_{S}(m)$ therefore $m\in\mathscr{G}(M)$. Conversely, if $m\in\mathscr{G}(M)$ then there exists an ideal $I\in\mathscr{F}$ such that $IS\subseteq\Ann_{S}(m)$. Thus $I\subseteq\Ann_{R}(m)$ and so $m\in\mathscr{F}(M)$. Therefore $\mathscr{F}(M)=\mathscr{G}(M)$. \\ Let $f:J\rightarrow M/\mathscr{G}(M)$ be a $S-$linear map where $J\in\mathscr{G}$; there exists some $I\in\mathscr{F}$ such that $IS\subseteq J$. Consider the map $\eta:M_{\mathscr{G}}\rightarrow M_{\mathscr{F}}$ given by $[f]\rightsquigarrow [f\circ\phi_{|I}]$. Note that this definition is independent of choosing such $I$. We shall prove that $\eta$ is bijective. Clearly it is injective. Let $f:I\rightarrow M/\mathscr{F}(M)$ be an $R-$linear map for some $I\in\mathscr{F}$. Then consider the map $f^{\ast}:IS\rightarrow M/\mathscr{G}(M)$ which maps each $\sum\limits_{j=1}^{n}s_{j}\phi(a_{j})$ into $\sum\limits_{j=1}^{n}s_{j}f(a_{j})$. The map $f^{\ast}$ is well-defined. Because, suppose $\sum\limits_{j=1}^{n}s_{j}\phi(a_{j})=\sum\limits_{i=1}^{m}s'_{i}\phi(b_{i})$ where $a_{j},b_{i}\in I$ and $s_{j},s'_{i}\in S$. Let $f(a_{j})=m_{j}+\mathscr{F}(M)$ and $f(b_{i})=m'_{i}+\mathscr{F}(M)$ for all $i$ and $j$. Set $m=\sum\limits_{j=1}^{n}s_{j}m_{j}$ and $m'=\sum\limits_{i=1}^{m}s'_{i}m'_{i}$. We have $IS\subseteq\Ann_{S}\big(m-m'+\mathscr{G}(M)\big)$. Because for each $c\in I$, $\phi(c)\big(m-m'+\mathscr{G}(M)\big)= \sum\limits_{j=1}^{n}s_{j}\phi(c)f(a_{j})- \sum\limits_{i=1}^{m}s'_{i}\phi(c)f(b_{i})= \sum\limits_{j=1}^{n}s_{j}f(a_{j}c)-\sum\limits_{i=1}^{m}s'_{i}f(b_{i}c)= \big(\sum\limits_{j=1}^{n}s_{j}\phi(a_{j})- \sum\limits_{i=1}^{m}s'_{i}\phi(b_{i})\big)f(c)=0$. Therefore $m-m'+\mathscr{G}(M)\in\mathscr{G}\big(M/\mathscr{G}(M)\big)=0$. Thus $f^{\ast}$ is well-defined. Clearly it is $S-$linear and $\eta([f^{\ast}])=[f]$ and so $\eta$ is surjective.\\ $\textbf{(iv)}$: One can easily verify that $\eta^{-1}\circ\phi_{\mathscr{F}}:R_{\mathscr{F}}\rightarrow S_{\mathscr{G}}$ is actually a ring homomorphism. $\Box$ \\ \begin{lemma}\label{G-lemma 12} Let $\mathscr{F}$ be an idempotent topologizing system on the ring $R$ such that the zero ideal of $R$ does not belong to $\mathscr{F}$. If $R$ is an integral domain, then so is $R_{\mathscr{F}}$.\\ \end{lemma} {\bf Proof.} Clearly $R_{\mathscr{F}}$ is a non-trivial ring if and only if $0\notin\mathscr{F}$. We also have $\mathscr{F}(R)=0$. Let $z,z'\in R_{\mathscr{F}}$ with representations $f$ and $g$, i.e., $z=[f]$ and $z'=[g]$ such that $z.z'=0$. Suppose $z,z'\neq0$. We may assume that the $R-$linear maps $f$ and $g$ have the same domain $I\in\mathscr{F}$, i.e., $f,g:I\rightarrow R$. There exists an ideal $J\in\mathscr{F}$ such that $J\subseteq f^{-1}(I)$ and $g\big(f(a)\big)=0$ for all $a\in J$. There are also elements $b,c\in J$ such that $f(b), g(c)\neq0$ since $z,z'\neq0$. But we have $0=g\big(f(bc)\big)=g\big(cf(b)\big)=f(b)g(c)$ which is a contradiction. Therefore we have either $z=0$ or that $z'=0$. $\Box$ \\ \begin{remark} If $M'$ is a $R-$submodule of $M$ then $i_{\mathscr{F}}:M'_{\mathscr{F}}\rightarrow M_{\mathscr{F}}$ is injective where $i:M'\rightarrow M$ is the canonical injection. By abuse of the notation, the image of $i_{\mathscr{F}}$ is also denoted by $M'_{\mathscr{F}}$. \\ \end{remark} \begin{remark}\label{remark 762} For each ideal $I$ of $R$ we have $IR_{\mathscr{F}}\subseteq I_{\mathscr{F}}$. Moreover, $I_{\mathscr{F}}=R_{\mathscr{F}}$ for all $I\in\mathscr{F}$ because $R/I$ is $\mathscr{F}-$negligible then apply Lemma \ref{G-lemma 6}. As a second proof, we observe that $\pi|_{I}=\overline{i}\circ f$ where $\pi:R\rightarrow R/\mathscr{F}(R)$ is the canonical map, $f:I\rightarrow I/\mathscr{F}(I)$ which maps each $a\in I$ into $a+\mathscr{F}(I)$ and $i:I\rightarrow R$ is the canonical injection. Thus $1_{R_{\mathscr{F}}}=[\pi]=[\pi|_{I}]\in I_{\mathscr{F}}$. \\ \end{remark} \begin{lemma}\label{G-lemma 13} Let $M$ be a $R-$module and let $N$ be a $R-$submodule of $M_{\mathscr{F}}$. Then $\big(j^{-1}_{M}(N)\big)_{\mathscr{F}}= j^{-1}_{M_{\mathscr{F}}}(N_{\mathscr{F}})$. If moreover $\mathscr{F}\big(M_{\mathscr{F}}/N\big)=0$, then $N=j^{-1}_{M_{\mathscr{F}}}(N_{\mathscr{F}})$. \\ \end{lemma} {\bf Proof.} Let $N'=j^{-1}_{M}(N)$. First we show that $N'_{\mathscr{F}}\subseteq j^{-1}_{M_{\mathscr{F}}}(N_{\mathscr{F}})$. Let $x=[\overline{i}\circ f]\in N'_{\mathscr{F}}$ where $f:I\rightarrow N'/\mathscr{F}(N')$ is an $R-$linear map and $I\in\mathscr{F}$. Consider the map $g:I\rightarrow N$ which maps each $a\in I$ into $j_{M}(z)$ where $f(a)=z+\mathscr{F}(N')$. Note that the map $g$ is well-defined because if there exists another $z'\in N'$ such that $f(a)=z'+\mathscr{F}(N')$ then $j_{M}(z-z')\in\mathscr{F}(N)\subseteq\mathscr{F}(M_{\mathscr{F}})=0$. Clearly it is also $R-$linear. Therefore $[g]\in N_{\mathscr{F}}$. We have $(\delta_{x})|_{I}=i\circ g$ where $i:N\rightarrow M_{\mathscr{F}}$ is the canonical injection. This implies that $j_{M_{\mathscr{F}}}(x)\in N_{\mathscr{F}}$. To prove the converse inclusion we act as follows. Let $x\in j^{-1}_{M_{\mathscr{F}}}(N_{\mathscr{F}})$. Thus there exists an element $y\in N_{\mathscr{F}}$ such that $j_{M_{\mathscr{F}}}(x)= i_{\mathscr{F}}(y)$. We claim that $\Coker(\psi)=N/j_{M}(N')$ is $\mathscr{F}-$negligible where $\psi=(j_{M})|_{N'}:N'\rightarrow N$. For each $x=[f]\in N$, we show that $I\subseteq\Ann_{R}\big(x+j_{M}(N')\big)$ where $f:I\rightarrow M/\mathscr{F}(M)$ is an $R-$linear map and $I\in\mathscr{F}$. For each $a\in I$ there is some $m\in M$ such that $f(a)=\overline{m}=m+\mathscr{F}(M)$. Clearly $(\delta_{\overline{m}})|_{I}=a.f$ thus $j_{M}(m)=a.x\in N$ and so $m\in N'$. This establishes the claim. We also have $\Ker(\psi)=\mathscr{F}(N')$. Therefore, by Theorem \ref{G-lemma 10}, $\psi_{\mathscr{F}}$ is bijective. Thus there is an element $z\in N'_{\mathscr{F}}$ such that $\psi_{\mathscr{F}}(z)=y$. Therefore $x=i'_{\mathscr{F}}(z)\in\Ima i'_{\mathscr{F}}=N'_{\mathscr{F}}$ because $j_{M_{\mathscr{F}}}\circ i'_{\mathscr{F}}=i_{\mathscr{F}}\circ\psi_{\mathscr{F}}$ where $i':N'\rightarrow M$ is the canonical injection. Now we prove the last part of the assertion. For each $x\in N$, $\delta_{x}=i\circ g$ where the map $g:R\rightarrow N$ is defined as $r\rightsquigarrow r.x$. Thus $N\subseteq j^{-1}_{M_{\mathscr{F}}}(N_{\mathscr{F}})$. Conversely, let $x\in j^{-1}_{M_{\mathscr{F}}}(N_{\mathscr{F}})$ then there exists an ideal $I\in\mathscr{F}$ such that $I\subseteq\Ann_{R}(x+N)$. Therefore $x+N\in\mathscr{F}\big(M_{\mathscr{F}}/N\big)=0$. $\Box$ \\ Now we prove the final main result of this article:\\ \begin{theorem}\label{theorem 33} Let $\mathscr{F}$ be an of finite type system on the ring $R$ and let $\mathscr{F}'$ be the set of ideals $J$ of $R_{\mathscr{F}}$ such that $R_{\mathscr{F}}/J$ as $R-$module is $\mathscr{F}-$negligible. Then the map $\mathfrak{p}\rightsquigarrow\mathfrak{p}_{\mathscr{F}}$ is a bijection between the set of prime ideals of $R$ which do not belong to $\mathscr{F}$ and the set of prime ideals of $R_{\mathscr{F}}$ which do not belong to $\mathscr{F}'$.\\ \end{theorem} {\bf Proof.} First of all we show that if $\mathfrak{p}$ is a prime ideal of $R$ such that $\mathfrak{p}\notin\mathscr{F}$ then $\mathfrak{p}_{\mathscr{F}}$ is a prime ideal of $R_{\mathscr{F}}$ and that $\mathfrak{p}_{\mathscr{F}}\notin\mathscr{F}'$. Suppose $1_{R_{\mathscr{F}}}=[\pi']\in\mathfrak{p}_{\mathscr{F}}$ where $\pi':R\rightarrow R/\mathscr{F}(R)$ is the canonical map. Thus there exists an ideal $I\in\mathscr{F}$ such that for each $a\in I$ there is some $x_{a}\in\mathfrak{p}$ in which $\Ann_{R}(a-x_{a})\in\mathscr{F}$. But $I$ is not contained in $\mathfrak{p}$ since $\mathfrak{p}\notin\mathscr{F}$. Take $b\in I\setminus\mathfrak{p}$ then clearly $\Ann_{R}(b-x_{b})\subseteq\mathfrak{p}$. This implies that $\mathfrak{p}\in\mathscr{F}$ which is a contradiction therefore $\mathfrak{p}_{\mathscr{F}}$ is a proper ideal of $R_{\mathscr{F}}$.\\ We shall apply Proposition \ref{G-prop 2} to the canonical ring map $\pi:R\rightarrow R/\mathfrak{p}$. If $0\in\mathscr{G}$ then $R/\mathfrak{p}$ will be $\mathscr{F}-$negligible. By applying Lemma \ref{G-lemma 6} to the exact sequence $\xymatrix{0\ar[r]&\mathfrak{p}\ar[r]&R\ar[r]^{\pi\:\:\:\:}&R/\mathfrak{p}}$ we get that $\mathfrak{p}_{\mathscr{F}}=R_{\mathscr{F}}$ which is a contradiction since $\mathfrak{p}_{\mathscr{F}}$ is a proper ideal of $R_{\mathscr{F}}$. Thus $0\notin\mathscr{G}$ and so by Lemma \ref{G-lemma 12}, $(R/\mathfrak{p})_{\mathscr{G}}\simeq(R/\mathfrak{p})_{\mathscr{F}}$ is an integral domain. Suppose $xy\in\mathfrak{p}_{\mathscr{F}}$ for some elements $x,y\in R_{\mathscr{F}}$. Thus $0=\pi_{\mathscr{F}}(xy)=\pi_{\mathscr{F}}(x)\pi_{\mathscr{F}}(y)$. Therefore we have either $\pi_{\mathscr{F}}(x)=0$ or that $\pi_{\mathscr{F}}(y)=0$. Hence $\mathfrak{p}_{\mathscr{F}}$ is a prime ideal.\\ Suppose $\mathfrak{p}_{\mathscr{F}}\in\mathscr{F}'$ then there exists an ideal $I\in\mathscr{F}$ such that $IR_{\mathscr{F}}\subseteq\mathfrak{p}_{\mathscr{F}}$. This implies that $I\subseteq\mathfrak{p}$. Because for each $a\in I$ there exists an ideal $J\in\mathscr{F}$ such that for each $b\in J$ there is some $y_{b}\in\mathfrak{p}$ for which $\Ann_{R}(ab-y_{b})\in\mathscr{F}$. Take some $b\in J\setminus\mathfrak{p}$ then $\Ann_{R}(ab-y_{b})$ is not contained in $\mathfrak{p}$. It follows that $a\in\mathfrak{p}$. But $I\subseteq\mathfrak{p}$ is a contradiction since $\mathfrak{p}\notin\mathscr{F}$, therefore $\mathfrak{p}_{\mathscr{F}}\notin\mathscr{F}'$. \\ The map $\mathfrak{p}\rightsquigarrow\mathfrak{p}_{\mathscr{F}}$ is injective between the foregoing sets. Because suppose $\mathfrak{p}_{\mathscr{F}}=\mathfrak{p}'_{\mathscr{F}}$ where $\mathfrak{p}$ and $\mathfrak{p}'$ are prime ideals of $R$ with $\mathfrak{p},\mathfrak{p}'\notin\mathscr{F}$. For each $a\in\mathfrak{p}$, $j_{\mathfrak{p}}(a)\in\mathfrak{p}'_{\mathscr{F}}$ thus there exists an ideal $J\in\mathscr{F}$ such that for each $b\in J$ there is some $x_{b}\in\mathfrak{p}'$ in which $\Ann_{R}(ab-x_{b})\in\mathscr{F}$. Now take some $b\in J\setminus\mathfrak{p}'$ also take some $c\in\Ann_{R}(ab-x_{b})\setminus\mathfrak{p}'$, then we have $c(ab-x_{b})=0\in\mathfrak{p}'$ and so $a\in\mathfrak{p}'$; symmetrically we have $\mathfrak{p}'\subseteq\mathfrak{p}$, thus $\mathfrak{p}=\mathfrak{p}'$. \\ Finally, we show that the map $\mathfrak{p}\rightsquigarrow\mathfrak{p}_{\mathscr{F}}$ is surjective. Let $\mathfrak{q}$ be a prime ideal of $R_{\mathscr{F}}$ which does not belong to $\mathscr{F}'$. We have $\mathfrak{p}=j_{R}^{-1}(\mathfrak{q})\notin\mathscr{F}$ since $\mathfrak{p}R_{\mathscr{F}}\subseteq\mathfrak{q}$. We also have $\mathscr{F}(R_{\mathscr{F}}/\mathfrak{q})=0$ because pick $x+\mathfrak{q}\in\mathscr{F}(R_{\mathscr{F}}/\mathfrak{q})$, if $x\notin\mathfrak{q}$ then $\Ann_{R}(x+\mathfrak{q})R_{\mathscr{F}}\subseteq\mathfrak{q}$ it follows that $\mathfrak{q}\in\mathscr{F}'$ which is a contradiction. Therefore, by Lemma \ref{G-lemma 13}, $\mathfrak{p}_{\mathscr{F}}=\mathfrak{q}$. $\Box$ \\ The usual localization theory is just a special case of the G-localizations:\\ \begin{proposition}\label{G-th 1} Let $S$ be a multiplicative subset in the ring $R$ and let $\mathscr{F}$ be the set of ideals of $R$ which meeting $S$. Then for each $R-$module $M$, there is a unique $R-$linear map $\phi: S^{-1}M\rightarrow M_{\mathscr{F}}$ such that $j_{M}=\phi\circ\pi$ where $\pi:M\rightarrow S^{-1}M$ is the canonical map. Moreover $\phi$ is bijective. In particular, the $R-$algebras $S^{-1}R$ and $R_{\mathscr{F}}$ are canonically isomorphic.\\ \end{proposition} {\bf Proof.} Clearly $\Ker(\pi)$ and $\Coker(\pi)$ are $\mathscr{F}-$negligible. Therefore, by Proposition \ref{G-lemma 2}, the first part of the assertion is realized. Now for each $I\in\mathscr{F}$, consider the map $\lambda_{I}:\Hom_{R}\big(I,M/\mathscr{F}(M)\big)\rightarrow S^{-1}M$ defined by $f\rightsquigarrow m/s$ where $f(s)=m+\mathscr{F}(M)$ for some $s\in I\cap S$. The map $\lambda_{I}$ is well-defined. Because suppose there is another element $s'\in I\cap S$ such that $f(s')=m'+\mathscr{F}(M)$. We have $s'f(s)=f(ss')=sf(s')$. It implies that $S\cap\Ann_{R}(sm'-s'm)\neq\emptyset$. The map $\lambda_{I}$ is also $R-$linear and clearly $\lambda_{I}=\lambda_{J}\circ u_{I,J}$ for every $I,J\in\mathscr{F}$ with $I\leq J$. Therefore, by the universal property of the colimits, there is a (unique) $R-$linear map $\lambda:M_{\mathscr{F}}\rightarrow S^{-1}M$ such that the map $\lambda_{I}$ factors as $\xymatrix{\Hom_{R}\big(I,M/\mathscr{F}(M)\big)\ar[r] & M_{\mathscr{F}}\ar[r]^{\lambda\:\:\:\:\:}&S^{-1}M}$ for all $I\in\mathscr{F}$. The map $\lambda$ is injective because suppose $\lambda([f])=0$ for some element $[f]\in M_{\mathscr{F}}$ where $f:I\rightarrow M/\mathscr{F}(M)$ is a $R-$linear map and $I\in\mathscr{F}$. We know that $I\cap S\neq\emptyset$. Take some $s\in I\cap S$ then clearly $Rs\in\mathscr{F}$ and $f|_{Rs}=0$. The map $\lambda$ is also surjective. Because for each element $m/s\in S^{-1}M$ then consider the $R-$linear map $g:Rs\rightarrow M/\mathscr{F}(M)$ defined by $rs\rightsquigarrow rm+\mathscr{F}(M)$. Note that the map $g$ is well-defined since if $rs=r's$ then $r-r'\in\mathscr{F}(R)$ and so $(r-r')m\in\mathscr{F}(R)M\subseteq\mathscr{F}(M)$. It is also $R-$linear and clearly $\lambda([g])=m/s$. We have $\lambda^{-1}=\phi$. $\Box$ \\
1990a29d6b2d3883d829074369c04db310196d9b
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Initiated by Gelu Popescu in \cite{MR1670202}, noncommutative Poisson transforms, and subsequently the explicit and analytic construction of isometric dilations, have been proved to be an extremely powerful tools in studying the structure of commuting and noncommuting tuples of bounded linear operators on Hilbert spaces. This is also important in noncommutative domains (and subsequently, noncommutative varieties) classification problems in the operator algebras (see \cite{MR2643314}, \cite{GP10}, \cite{GP11} and references therein). In \cite{MR1668582} Arveson used similar techniques to generalize Sz.-Nagy and Foias dilation theory for commuting tuple of row contractions. These techniques have also led to further recent development \cite{H16, MR1648483, MR2005882, MR2186099, MR2481905, MR1822685, MR2608451, MR2793449} on the structure of bounded linear operators in more general settings. In particular, in \cite{MR2481905} Muhly and Solel introduced Poisson kernel for completely contractive, covariant representations over $W^*$-correspondences. The notion of Poisson kernel for completely contractive, covariant representations over a subproduct system of $W^*$-correspondences was introduced and studied by Shalit and Solel in \cite{MR2608451}. This approach was further investigated by Viselter \cite{MR2793449} for the extension problem of completely contractive, covariant representations of subproduct systems to $C^*$-representations of Toeplitz algebras. Covariant representations on subproduct systems are important since it is one of the refined theories in operator theory and operator algebras that provides a unified approach to study commuting as well as noncommuting tuples of operators on Hilbert spaces. The main purpose of this paper is to investigate a Beurling-Lax-Halmos type invariant subspace theorem in the sense of \cite{MR3346131,J2}, and the notion of curvature in the sense of Arveson \cite{A2000}, Popescu \cite{MR1822685} and Muhly and Solel \cite{MR2005882} for completely contractive, covariant representations of standard subproduct systems. The plan of the paper is the following. In Section 2, we recall several basic results from \cite{MR2793449} including the intertwining property of Poisson kernels. In Section 3 we obtain an invariant subspace theorem for pure completely contractive, covariant representations of standard subproduct systems. As an immediate application we derive a Beurling-Lax-Halmos type theorem. Our objective in Section 4 is to extend, several results on curvature of a contractive tuple by Popescu \cite{MR1822685,MR2643314}, for completely contractive, covariant representations of subproduct systems. We first define the curvature for completely contractive, covariant representations of standard subproduct systems. This approach is based on the definition of curvature for a completely contractive, covariant representation over a $W^*$-correspondence due to Muhly and Solel \cite{MR2005882}. The final section is composed of several results on wandering subspaces which are motivated from our invariant subspace theorem. This section generalizes \cite[Section 5]{BEDS} on wandering subspaces for commuting tuple of bounded operators on Hilbert spaces. \section{Notations and prerequisites} In this section, we recall some definitions and properties about $C^*$-correspondences and subproduct systems (see \cite{MR0355613}, \cite{La95}, \cite{MR1648483}, \cite{MR2608451}). Let $\mathcal M$ be a $C^*$-algebra and let $E$ be a Hilbert $\mathcal M$-module. Let $\mathcal L(E)$ be the $C^*$-algebra of all adjointable operators on $E$. The module $E$ is said to be a {\it $C^*$-correspondence over $\mathcal M$} if it has a left $\mathcal M$-module structure induced by a non-zero $*$-homomorphism $\phi:\mathcal M\to \mathcal L(E)$ in the following sense \[ a\xi:=\phi(a)\xi \quad \quad (a\in\mathcal M, \xi\in E). \] All such $*$-homomorphisms considered in this article are non-degenerate, which means, the closed linear span of $\phi(\mathcal M)E$ equals $E.$ If $F$ is another $C^*$-correspondence over $\mathcal M,$ then we get the notion of tensor product $F\bigotimes_{\phi} E$ (cf. \cite{La95}) which satisfy the following properties: \[ (\zeta_1 a)\otimes \xi_1=\zeta_1\otimes \phi(a)\xi_1, \] \[ \langle\zeta_1\otimes\xi_1,\zeta_2\otimes\xi_2\rangle=\langle\xi_1,\phi(\langle\zeta_1,\zeta_2\rangle)\xi_2\rangle \] for all $\zeta_1,\zeta_2\in F;$ $\xi_1,\xi_2\in E$ and $a\in\mathcal M.$ Assume $\mathcal M$ to be a $W^*$-algebra and $E$ is a Hilbert $\mathcal M$-module. If $E$ is self-dual, then $E$ is called a {\it Hilbert $W^*$-module} over $\mathcal M.$ In this case, $\mathcal L(E)$ becomes a $W^*$-algebra (cf. \cite{MR0355613}). A $C^*$-correspondence over $\mathcal M$ is called a {\it $W^*$-correspondence} if $E$ is self-dual, and if the $*$-homomorphism $\phi:\mathcal M\to \mathcal L(E)$ is normal. When $E$ and $F$ are $W^*$-correspondences, then their tensor product $F\bigotimes_{\phi} E$ is the self-dual extension of the above tensor product construction. \begin{definition} Let $\mathcal M$ be a $C^*$-algebra, $\mathcal H$ be a Hilbert space, and $E$ be a $C^*$-correspondence over $\mathcal M$. Assume $\sigma:\mathcal M\to B(\mathcal H)$ to be a representation and $T: E\to B(\mathcal H)$ to be a linear map. The tuple $(T,\sigma)$ is called a {\rm covariant representation} of $E$ on $\mathcal H$ if \[ T(a\xi a')=\sigma(a)T(\xi)\sigma(a') \quad \quad (\xi\in E, a,a'\in\mathcal M). \] In the $W^*$-set up, we additionally assume that $\sigma$ is normal and that $T$ is continuous with respect to the $\sigma$-topology of $E$ (cf. \cite{MR945550}) and ultra weak topology on $B(\mathcal H)$. The covariant representation is called {\rm completely contractive} if $T$ is completely contractive. The covariant representation $(T,\sigma)$ is called {\rm isometric} if \[ T(\xi)^*T(\zeta)=\sigma(\langle \xi,\zeta\rangle) \quad \quad (\xi,\zeta\in E). \] \end{definition} The following important lemma is due to Muhly and Solel \cite[Lemma 3.5]{MR1648483}: \begin{lemma} The map $(T,\sigma)\mapsto \widetilde T$ provides a bijection between the collection of all completely contractive, covariant representations $(T,\sigma)$ of $E$ on $\mathcal H$ and the collection of all contractive linear maps $\widetilde{T}:~\mbox{$E\bigotimes_{\sigma} \mathcal H\to \mathcal H$}$ defined by \[ \widetilde{T}(\xi\otimes h):=T(\xi)h \quad \quad (\xi\in E, h\in\mathcal H), \] and such that $\widetilde{T}(\phi(a)\otimes I_{\mathcal H})=\sigma(a)\widetilde{T}$, $a\in\mathcal M$. Moreover, $\widetilde T$ is isometry if and only if $(T,\sigma)$ is isometric. \end{lemma} \begin{example}\label{eg-1} Assume $E$ to be a Hilbert space with an orthonormal basis $\{e_i\}^n_{i=1}.$ Any contractive tuple $(T_1, \ldots,T_n)$ on a Hilbert space $\mathcal H$ can be realized as a completely contractive, covariant representation $(T,\sigma)$ of $E$ on $\mathcal H$ where $T(e_i):=T_i$ for each $1\leq i\leq n,$ and when the representation $\sigma$ maps every complex number $\lambda$ to the multiplication operator by $\lambda$. \end{example} Now we recall several definitions and results from \cite{MR2793449} which are essential for our objective. We will use $A^*$-algebra, to denote either $C^*$-algebra or $W^*$-algebra, to avoid repetitions in statements. Similarly we also use $A^*$-module and $A^*$-correspondence. \begin{definition}\label{def2} Let $\mathcal{M}$ to be an $A^*$-algebra and $X=(X(n))_{n \in \mathbb{Z_+}}$ be a sequence of $A^*$-correspondences over $\mathcal{M}$. Then $X$ is said to be a {\rm subproduct system} over $\mathcal{M}$ if $X(0)=\mathcal{M},$ and for each $n,m \in \mathbb{Z_+}$ there exist a coisometric, adjointable bimodule function $$U_{n,m}:X(n)\mbox{$\bigotimes$} X(m) \to X(n+m),$$ such that \begin{itemize} \item[(a)] the maps $U_{n,0}$ and $U_{0,n}$ are the right and the left actions of $\mathcal{M}$ on $X(n)$, respectively, that is, \[ U_{n,0}(\zeta \otimes a):=\zeta a,~U_{0,n}(a \otimes \zeta):=a\zeta \quad \quad (\zeta\in X(n),~a \in \mathcal M, n\in\mathbb{Z_+}), \] \item [(b)] the following associativity property holds for all $n,m,l \in \mathbb{Z_+}$; \[ U_{n+m,l}(U_{n,m}\otimes I_{X(l)})=U_{n,m+l}(I_{X(n)} \otimes U_{m,l}). \] If each coisometric maps are unitaries, then we say the family $X$ is a {\rm product system}. \end{itemize} \end{definition} \begin{definition} \label{def1} Let $\mathcal{M}$ be an $A^*$-algebra and let $ X=(X(n))_{n \in \mathbb{Z_+} }$ be a subproduct system over $\mathcal{M}$. Assume $T=(T_n)_{n \in \mathbb{Z_+}}$ to be a family of linear transformations $T_n: X(n) \to B(\mathcal{H})$, and define $\sigma:=T_0$. Then the family $T$ is called a {\rm completely contractive, covariant representation of $X$ on $\mathcal{H}$} if \begin{itemize} \item [(i)] for every $n \in \mathbb{Z_+}$, the pair $(T_n, \sigma)$ is a completely contractive, covariant representation of the $A^*$-correspondence $X(n)$ on $\mathcal{H}$, and \\ \item [(ii)] for every $n,m \in \mathbb{Z_+},~ \zeta \in X(n)$ and $ \eta \in X(m)$, \be\label{eqn1} T_{n+m}(U_{n,m}(\zeta \otimes \eta))=T_n(\zeta)T_m(\eta). \ee \end{itemize} \end{definition} For $n \in \mathbb{Z_+}$ define the contractive linear map $\widetilde{T}_n:X(n) \bigotimes_{\sigma} \mathcal{H} \to \mathcal{H}$ as (see \cite{MR1648483}) \begin{equation}\label{eqn3} \widetilde{T}_n(\zeta \otimes h):=T_n(\zeta)h \quad \quad (\zeta \in X(n),~h \in \mathcal{H}). \end{equation} Thus we can replace (\ref{eqn1}) by \[ \widetilde{T}_{n+m}(U_{n,m} \otimes I_{\mathcal{H}})=\widetilde{T}_n(I_{X(n)} \otimes \widetilde{T}_m). \] \begin{example}\label{eg-2} The Fock space $\mathcal{F}_X:=\bigoplus_{n \in \mathbb{Z_+}}X(n)$ of a subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ is an $A^*$-correspondence over $\mathcal{M}$. For each $n \in \mathbb{Z_+}$, we define a linear map $S^X_n: X(n) \to \mathcal{L}(\mathcal F_X) $ by $$S^X_n(\zeta)\eta:=U_{n,m}(\zeta \otimes \eta)$$ for every $m \in \mathbb{Z_+},$ $\zeta\in X(n)$ and $\eta \in X(m)$. When $ n \neq 0$ we call each operator $S^X_n$ a {\it creation operator} of $\mathcal{F}_X$, and the family $S^X:=(S^X_n)_{n \in \mathbb{Z_+}}$ is called an {\it $X$-shift}. It is easy to verify that the family $S^X$ is indeed a completely contractive, covariant representation of $\mathcal{F}_X$. From the Definition \ref{def2} it is easy to see that, for each $a\in \mathcal{M},$ the map $S^X_0(a)= \phi _{\infty}(a):\mathcal{F}_X\to \mathcal{F}_X$ maps $(b,\zeta_1,\zeta_2,\ldots)\mapsto (a b, a\zeta_1,a \zeta_2,\ldots)$. \end{example} Let $\mathcal M$ to be an $A^*$-algebra, and let $X=(X(n))_{n \in \mathbb{Z_+}}$ be an $A^*$-correspondences over $\mathcal{M}$. Then $X$ is said to be a \textit{standard subproduct system} if $X(0)=\mathcal{M},$ and for any $n,m \in \mathbb{Z_+}$ the bimodule $X(n+m)$ is an orthogonally complementable sub-module of $X(n)\bigotimes X(m)$. Let $X=(X(n))_{n \in \mathbb{Z_+}}$ be a standard subproduct system and $E:=X(1)$. Then for each $n,$ the bi-module $X(n)$ is an orthogonally complementable sub-module of $E^{\otimes n}$ (here $E^{\otimes 0}=\mathcal{M}$), and hence there exists an orthogonal projection $p_n \in \mathcal{L}(E^{\otimes n})$ of $E^{\otimes n}$ onto $X(n)$. We denote the orthogonal projection $\bigoplus_{n \in \mathbb{Z_+}}p_n$ of $\mathcal{F}_E$, the Fock space of the product system $E=(E^{\otimes n})_{n \in \mathbb{Z_+}}$ with trivial unitaries, onto $\mathcal{F}_X$ by $P$. Note also that here the projections $(p_n)_{n \in \mathbb{Z_+}}$ are bimodule maps and \[ p_{n+m}=p_{n+m}(I_{E^{\otimes n}} \otimes p_{m})=p_{n+m}(p_{n} \otimes I_ {E^{\otimes m}}), \] for all $n,m \in \mathbb{Z_+}$. This implies that if we define each $U_{n,m}$ to be the projection ${p_{n+m}}$ restricted to ${X(n)\bigotimes X(m)},$ then every standard subproduct system becomes a subproduct system over $\mathcal{M}.$ In this case (\ref{eqn1}) reduces to \[ T_{n+m}(p_{n+m}(\zeta \otimes \eta))=T_n(\zeta)T_m(\eta)~\mbox{for all}~\zeta\in E^{\otimes n}\mbox{ and }\eta\in E^{\otimes m}, \] and (\ref{eqn3}) becomes \be \label{eqn6} \widetilde{T}_{n+m}(p_{n+m} \otimes I_{\mathcal{H}})|_{X(n) \bigotimes X(m) \bigotimes_{\sigma} \mathcal{H}}=\widetilde{T}_n(I_{X(n)} \otimes \widetilde{T}_m). \ee Taking adjoints on both the sides we obtain \be \label{eqn5} {\widetilde{T}}_{n+m}^*=(I_{X(n)} \otimes {\widetilde{T}_m}^*){\widetilde{T}_n}^* \quad \quad (n,m\in \mathbb{Z_+}).\ee Note that for the sake of convenience we ignored the embedding of $X(n+m) \bigotimes_{\sigma} \mathcal{H}$ into $X(n)\bigotimes X(m) \bigotimes_{\sigma} \mathcal{H}$ in the previous formula. We further deduce that \be ~\label{eqn4} {\widetilde{T}_{n+1}}^*=(I_E \otimes {\widetilde{T}_n}^*){\widetilde{T}_1}^*=(I_{X(n)} \otimes {\widetilde{T}_1}^*){\widetilde{T}_n}^*,\ee and \[ {\widetilde{T}_n}^*=(I_{X(n-1)} \otimes {\widetilde{T}_1}^*)(I_{X(n-2)} \otimes {\widetilde{T}_1}^*)\ldots(I_E \otimes {\widetilde{T}_1}^*){\widetilde{T}_1}^*, \] for all $n \in \mathbb{Z_+}$. \begin{example}\label{eg-3} If $X(n)$ is the $n$-fold symmetric tensor product of the Hilbert space $X(1),$ then $X=(X(n))_{n\in\mathbb Z_+}$ becomes a standard subproduct system of Hilbert spaces (cf. \cite[Example 1.3]{MR2608451}). Moreover, let $\{e_1,\ldots,e_d\}$ be an orthonormal basis of $X(1)$. Then \[ T \leftrightarrow (T_1(e_1),T_1(e_2),\ldots,T_1(e_d)) \] induces a bijection between the set of all completely contractive covariant representations $T$ of $X$ on a Hilbert space $\mathcal H$ onto the collection of all commuting row contractions $(T_1, \ldots,T_d)$ on $\mathcal H$ (cf. \cite[Example 5.6]{MR2608451}). \end{example} \vspace{0.2in} Before proceeding to the notion of Poisson kernels, we make a few comments: \begin{enumerate} \item We use the symbol sot-$\lim$ for the limit with respect to the strong operator topology. From Equation \ref{eqn4} we infer that $\{\widetilde{T}_n {\widetilde{T}_n^*}\}_{n \in \mathbb{Z_+}}$ is a decreasing sequence of positive contractions, and thus $Q:=\mbox{sot-}\displaystyle\lim_{n \rightarrow \infty}\widetilde{T}_n {\widetilde{T}_n^*}$ exists. If $Q=0$, then we say that the covariant representation $T$ is {\it pure}. Note that $T$ is pure if and only if $\mbox{sot-}\displaystyle\lim_{n \rightarrow \infty}{\widetilde{T}_n^*}=0$. \item Let $\psi$ be a representation of $\mathcal M$ on a Hilbert space $\mathcal E$. Then the induced covariant representation $S \otimes I_{\mathcal E}:=(S_n(\cdot)\otimes I_{\mathcal E})_{n\in\mathbb{Z}_+}$ is pure, where each $S_n(\cdot)\otimes I_{\mathcal E}$ is an operator from $X(n)$ into $B(\mathcal F_X \bigotimes_\psi \mathcal E)$. \item It is proved in \cite[Lemma 6.1]{MR2608451} that every subproduct system is isomorphic to a standard subproduct system. Therefore it is enough to consider standard subproduct systems. \end{enumerate} \vspace{0.2in} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a completely contractive, covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$. We denote the positive operator $(I_{\mathcal{H}}- \widetilde{T}_1 {\widetilde{T}_1^*} )^{1/2} \in B(\mathcal{H})$ by $\triangle_*(T)$ and the defect space $\overline{\mbox{\rm Im} ~\triangle_*(T)}$ by $\mathcal{D}$. It is proved in \cite[Proposition 2.9]{MR2793449} that $\triangle_*(T) \in \sigma(\mathcal{M})^\prime.$ Therefore $\mathcal{D}$ reduces $\sigma(a)$ for each $a \in \mathcal{M}$. Thus using the reduced representation $\sigma ^\prime $ we can form the tensor product of the Hilbert space $\mathcal{D}$ with $X(n)$ for each $n \in \mathbb{Z_+},$ and hence with $\mathcal{F}_X .$ For simplicity we write $\sigma$ instead of $\sigma ^\prime.$ The {\it Poisson kernel} of $T$ is the operator $K(T):\mathcal{H} \to \mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}$ defined by \[ K(T)h:=\sum_{n \in \mathbb{Z_+}}(I_{X(n)}\otimes \triangle_*(T)){\widetilde{T}_n}^*h \quad \quad (h \in \mathcal{H}). \] In the next proposition we recall the properties of the Poisson kernel from \cite{MR2793449}: \bpp\label{prop2} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a completely contractive, covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$. Then $K(T)$ is a contraction and \[ K(T)^*(S_n(\zeta)\otimes I_{\mathcal{D}})=T_n(\zeta)K(T)^* \quad \quad (n \in \mathbb{Z_+},~\zeta \in X(n)). \] Moreover, $K(T)$ is an isometry if and only if $T$ is pure. \epp \bp For each $h \in \mathcal{H}$, from (\ref{eqn4}) and (\ref{eqn6}) it follows that $$ \begin{array}{ccl} \displaystyle\sum_{n \in \mathbb{Z_+}}\|(I_{X(n)} \otimes \triangle_*(T)) {\widetilde{T}_n}^*h\|^2 & = & \displaystyle\sum_{n \in \mathbb{Z_+}} \langle {\widetilde{T}_n}(I_{X(n)}\otimes \triangle_*(T)^2){\widetilde{T}_n}^*h,h\rangle \\ & = & \displaystyle\sum_{n \in \mathbb{Z_+}} \langle {\widetilde{T}_n}(I_{X(n)}\otimes (I_{\mathcal{H}}-{\widetilde{T}_1}{\widetilde{T}_1}^*) ){\widetilde{T}_n}^*h,h\rangle \\ & = & \displaystyle\sum_{n \in \mathbb{Z_+}} \langle \widetilde{T}_n{\widetilde{T}_n}^*-\widetilde{T}_{n+1}{\widetilde{T}_{n+1}}^*h,h\rangle \\ &=&\langle h,h \rangle- {\lim}_{n \rightarrow \infty}\langle {\widetilde{T}_n}{\widetilde{T}_n}^*h,h\rangle, \end{array} $$ here we also used ${\widetilde{T}_0}{\widetilde{T}_0}^*=I_{\mathcal{H}}$. So $K(T)$ is a well-defined contraction, and it is an isometry if $T$ is pure. Now for each $n\in \mathbb{Z_+}$ and $z_n \in X(n) \bigotimes_{\sigma} \mathcal{D}$ we have \[ K(T)^*\left(\displaystyle\sum_{n \in \mathbb{Z_+}} z_n\right)=\displaystyle\sum_{n \in \mathbb{Z_+}} \widetilde{T}_n(I_{X(n)} \otimes \triangle_*(T))z_n. \] Therefore for every $m \in \mathbb{Z_+},$ $\eta \in X(m)$ and $h \in \mathcal{D}$, (\ref{eqn4}) gives $$ \begin{array}{ccl} K(T)^*(S_n(\zeta)\otimes I_{\mathcal{D}})(\eta \otimes h) & = & K(T)^*(p_{n+m}(\zeta \otimes \eta)\otimes h) \\ & = &{\widetilde{T}_{n+m}}(p_{n+m}(\zeta \otimes \eta) \otimes \triangle_*(T)h) \\ & = & {\widetilde{T}_n}(\zeta \otimes {\widetilde{T}_m}(\eta \otimes \triangle_*(T)h))\\ &=&T_n(\zeta)K(T)^*(\eta \otimes h). \qedhere\end{array} $$ \ep \section{Invariant subspaces of covariant representations} In this section we first introduce the notion of invariant subspaces for completely contractive, covariant representations and then in Theorem \ref{thm1} we obtain a far reaching generalization of \cite[Theorem 2.2]{MR3346131}. Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a completely contractive, covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$. A closed subspace $\mathcal{S}$ of $\mathcal{H}$ is called \textit{invariant} for the covariant representation $T$ if $\mathcal{S}$ is invariant for $\sigma(\mathcal{M})$ and if $\mathcal{S}$ is left invariant by each operator in the set $\{ T_n(\zeta) : \zeta \in X(n), n \in \mathbb{N} \}$. \bt\label{thm1} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a pure completely contractive, covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$, and let $\mathcal S$ be a non-trivial closed subspace of $\mathcal H$. Then $\mathcal{S}$ is invariant for $T$ if and only if there exist a Hilbert space $\mathcal{D},$ a representation $\pi$ of $\mathcal M$ on $\mathcal D,$ and a partial isometry $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{D}\to \mathcal{H} $ such that $\mathcal S = \mbox{ran~} \Pi$ and $$ \Pi (S_n(\zeta)\otimes I_{\mathcal{D}})=T_n(\zeta)\Pi \quad \quad (\zeta\in X(n), ~n\in \mathbb{Z_+}). $$ \et \bp Since $\mathcal{S}$ is invariant for $T=(T_n)_{n \in \mathbb{Z_+}}$, we get a covariant representation $(V_n:={T_n}|_{\mathcal{S}})_{n \in \mathbb{Z_+}}$ of the standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ on $\mathcal S.$ We denote $V_0$ by $\pi$. Now for each $ n \in \mathbb{N},~s\in \mathcal S,$ and $\zeta\in X(n)$, $$ \langle \zeta\otimes s,\zeta\otimes s \rangle=\langle s,\pi(\langle \zeta,\zeta\rangle) s\rangle=\langle s,\sigma(\langle \zeta,\zeta\rangle)s\rangle=\langle \zeta\otimes s, \zeta\otimes s\rangle, $$ yields an embedding $j_n$ from $X(n)\bigotimes_\pi \mathcal{S}$ into $X(n)\bigotimes_{\sigma} \mathcal{H}.$ Thus for each $ n \in \mathbb{N},$ $j_nj^*_n$ is an orthogonal projection. \noindent For each $ n \in \mathbb{N}$, from the definition of the map ${\widetilde{V}_n}: X(n)\bigotimes_{\pi} \mathcal{S} \to \mathcal{S}$ it follows that \[ {\widetilde{V}_n}(\zeta \otimes s)=V_n(\zeta)s=T_n(\zeta)s=\widetilde{T}_n\circ j_n(\zeta\otimes s), \] for all $\zeta\in X(n)$ and $s\in\mathcal S$. It also follows that \[ \langle\widetilde{V}_n\widetilde{V}^*_n s,s\rangle=\langle\widetilde{T}_nj_nj^*_n\widetilde{T}^*_n s,s\rangle\leq \langle\widetilde{T}_n\widetilde{T}^*_n s,s\rangle, \] for all $n\in \mathbb N$ and $s\in\mathcal S$. Hence the covariant representation $V$ is pure as well as completely contractive. \noindent Since the defect space $\mathcal{D}=\overline{\mbox{\rm Im} ~\triangle_*(V)}$ of the representation $V$ is reducing for $\pi$, it follows from Proposition \ref{prop2} that the Poisson kernel $K(V): \mathcal{S} \to \mathcal{F}_X \bigotimes_\pi \mathcal{D}$, defined by $$ K(V)(s)=\displaystyle\sum_{n \in \mathbb{Z_+}}(I_{X(n)}\otimes \triangle_*(V)){{\widetilde{V}_n}}^*s \quad \quad (s\in\mathcal S), $$ is an isometry and \[ K(V)^*(S_n(\zeta) \otimes I_{\mathcal{D}})=V_n(\zeta)K(V)^*, \] for all $n \in \mathbb{Z_+},$ and $\zeta\in X(n)$. Let $i_{\mathcal S}:\mathcal S \to \mathcal{H}$ be the inclusion map. Clearly $i_{\mathcal S}$ is an isometry and \[ i_{\mathcal S}T_n(\cdot)|_{\mathcal{S}}=T_n(\cdot) i_{\mathcal S}. \] Therefore we get a map $\Pi:\mathcal{F}_X \bigotimes_{\pi} \mathcal{D} \to \mathcal{H}$ defined by $\Pi:=i_{\mathcal S}K(V)^* $. Then $$\Pi \Pi^*=i_{\mathcal S}K(V)^*(i_{\mathcal S}K(V)^*)^*=i_{\mathcal S}i^*_{\mathcal S}=P_{\mathcal{S}},$$ the projection on $\mathcal S$. Hence $\Pi$ is a partial isometry and the range of $ \Pi$ is $\mathcal{S}$. From $i_{\mathcal S}V_n=i_{\mathcal S}T_n|_{\mathcal{S}}=T_n i_{\mathcal S}$ and the intertwining property of the Poisson kernel we deduce that \[ \Pi(S_n(\zeta) \otimes I_{\mathcal{D}})=i_{\mathcal S}K(V)^*(S_n(\zeta) \otimes I_{\mathcal{D}})=i_{\mathcal S}V_n(\zeta)K(V)^*=T_n(\zeta) \Pi. \] Conversely, suppose that there exists a partial isometry $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{D}\to \mathcal{H} .$ Then $ran \Pi$ is a closed subspace of $\mathcal{H}$ and the intertwining relation for $\Pi$ implies that $ran \Pi$ is a $T=(T_n)_{n \in \mathbb{Z_+}}$ invariant subspace of $\mathcal{H}$. \ep \begin{corollary} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a pure completely contractive, covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$, and $\mathcal S$ be a non-trivial closed subspace of $\mathcal H.$ Then $\mathcal{S}$ is invariant for $T$ if and only if there exist a Hilbert space $\mathcal{D},$ a representation $\pi$ of $\mathcal M$ on $\mathcal D,$ and a bounded linear operator $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{D}\to \mathcal{H} $ such that $P_{\mathcal S}=\Pi\Pi^*,$ and \[ \Pi (S_n(\zeta)\otimes I_{\mathcal{D}})=T_n(\zeta)\Pi \quad \quad (\zeta\in X(n), ~n\in \mathbb{Z_+}). \] \end{corollary} \begin{definition} Let $X=(X(n))_{n \in \mathbb{Z_+}}$ be a standard subproduct system over an $A^*$-algebra $\mathcal M$. Assume $\psi$ and $\pi$ to be representations of $\mathcal M$ on Hilbert spaces $\mathcal E$ and $\mathcal E',$ respectively. A bounded operator $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{E'}\to \mathcal F_X\bigotimes_\psi\mathcal E $ is called {\rm multi-analytic} if it satisfies the following condition $$\Pi (S_n(\zeta)\otimes I_{\mathcal{E'}})=(S_n(\zeta)\otimes I_{\mathcal E})\Pi~\mbox~{whenever}~\zeta\in X(n), ~n\in \mathbb{Z_+}.$$ Further we call it {\rm inner} if it is a partial isometry. \end{definition} As an application, we have the following Beurling-Lax-Halmos type theorem (cf. \cite[Theorem 3.2]{MR2643314}) which extends \cite[Theorem 2.4]{Po95} and \cite[Corollary 4.5]{J2}: \begin{theorem}\label{Beur} Assume $X=(X(n))_{n \in \mathbb{Z_+}}$ to be a standard subproduct system over an $A^*$-algebra $\mathcal M$ and assume $\psi$ to be a representation of $\mathcal M$ on a Hilbert space $\mathcal E$. Let $\mathcal S$ be a non-trivial closed subspace of the Hilbert space $\mathcal F_X\bigotimes_\psi\mathcal E.$ Then $\mathcal{S}$ is invariant for $S\otimes I_{\mathcal E}$ if and only if there exist a Hilbert space $\mathcal{E'},$ a representation $\pi$ of $\mathcal M$ on $\mathcal E',$ and an inner multi-analytic operator $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{E'}\to \mathcal F_X\bigotimes_\psi\mathcal E $ such that $\mathcal S$ is the range of $\Pi$. \end{theorem} \begin{proof} Let $\mathcal{S}$ be an invariant subspace for $S\otimes I_{\mathcal E}$. By Theorem \ref{thm1} we know that there exist a Hilbert space $\mathcal{E'},$ a representation $\pi$ of $\mathcal M$ on $\mathcal E',$ and a partial isometry $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{E'}\to \mathcal F_X\bigotimes_\psi\mathcal E$ such that $\mathcal S = \mbox{ran~} \Pi$ and \[ \Pi (S_n(\zeta)\otimes I_{\mathcal{E'}})=(S_n(\zeta)\otimes I_{\mathcal E})\Pi \quad \quad (\zeta\in X(n), ~n\in \mathbb{Z_+}). \] For the reverse direction, if we start with a partial isometry $\Pi: \mathcal{F}_X\bigotimes_{\pi}\mathcal{E'}\to \mathcal F_X\bigotimes_\psi\mathcal E$, then $ran \Pi$ is a closed subspace of $\mathcal F_X\bigotimes_\psi\mathcal E$ and the intertwining relation for $\Pi$ implies that $ran \Pi=\mathcal{S}$ is invariant for $S\otimes I_{\mathcal E}.$ \end{proof} For Beurling type classification in the tensor algebras setting see also Muhly and Solel \cite[Theorem 4.7]{MS99}. \section{Curvature} The notion of a curvature for commuting tuples of row contractions was introduced by Arveson \cite{A2000}. This numerical invariant is an analogue of the Gauss-Bonnet-Chern formula from Riemannian geometry, and closely related to rank of Hilbert modules over polynomial algebras. It has since been further analyzed by Popescu \cite{MR1822685} (see also \cite{GP15} for recent results on a general class), Kribs \cite{DK} in the setting of noncommuting tuples of operatos, and by Muhly and Solel \cite{MR2005882} in the setting of completely positive maps on $C^*$ algebras of bounded linear operators. The purpose of this section is to study curvature for a more general framework, namely, for completely contractive, covariant representations of subproduct systems. We begin by recalling the definition of left dimension \cite{MR1473221} for a $W^*$-correspondences $E$ over a semifinite factor $\mathcal{M}$ (see Muhly and Solel, Definition 2.5, \cite{MR2005882}). Let $\mathcal M$ be a semifinite factor and $\tau$ be a faithful normal semifinite trace, and let $L^2(\mathcal M)$ be the GNS construction for $\tau$. Note that for each $a\in\mathcal M$ there exists a left multiplication operator, denoted by $\lambda (a)$, and a right multiplication operator, denoted by $\rho (a)$, on $L^2(\mathcal M)$. Each unital, normal, $*$-representation $\sigma:\mathcal M\to B(\mathcal H)$ defines a {\it left $\mathcal M$-module} $\mathcal H$. This yields an $\mathcal M$-linear isometry $V:\mathcal H\to L^2(\mathcal M)\bigotimes l_2$. Here $\mathcal M$-linear means \[ V\sigma(a)=(\lambda(a)\otimes I_{l_2})V \quad \quad (a\in\mathcal M). \] Moreover \[ V\sigma(\mathcal M)'V^*=p(\lambda(\mathcal M)\otimes I_{l_2})'p\subseteq (\lambda(\mathcal M)\otimes I_{l_2})', \] where \[ p:=VV^*\in (\lambda(\mathcal M)\otimes I_{l_2})', \] is a projection. One can observe that $(\lambda(\mathcal M)\otimes I_{l_2})'$ equals the semifinite factor $\rho(\mathcal M)\bigotimes B(l_2)$ whose elements can be written as matrices of the form $(\rho(a_{ij}))$. For each positive element $x\in \sigma(\mathcal M)'$, we express $VxV^*$ in the form $(\rho(a_{ij}))$, and define $$tr_{\sigma(\mathcal M)'} (x):=\sum \tau (a_{ii}).$$ Note that $tr_{\sigma(\mathcal M)'}$ is a faithful normal semifinite trace on $\sigma(\mathcal M)'$. The {\it left dimension} of $\mathcal H$ is defined by $$dim_{l}(\mathcal H):=tr_{\sigma(\mathcal M)'}(p).$$ For each $W^*$-correspondence $E,$ the Hilbert space $E\bigotimes_{\sigma} L^2(\mathcal M)$ has a natural left $\mathcal M$-module structure. The left dimension of $E\bigotimes_{\sigma} L^2(\mathcal M)$ will be denoted by $dim_l(E).$ Now let $X=(X(n))_{n \in \mathbb{Z_+}}$ be a standard subproduct system of $W^*$-correspondences over a semifinite factor $\mathcal{M}$. Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a completely contractive, covariant representation of $X$ on a Hilbert space $\mathcal{H}$. Define a contractive, normal and completely positive map $\Theta_{T}:\sigma(\mathcal{M})'\to \sigma(\mathcal{M})'$ by $$ \Theta_T(a):=\widetilde{T}_1(I_{E}\otimes a)\widetilde{T}^*_1, $$ for all $a\in\sigma(\mathcal{M})'$. It follows from (\ref{eqn6}), (\ref{eqn5}) and (\ref{eqn4}) that \begin{align*} \Theta^2_T(a)&=\Theta_T(\Theta_T(a))=\widetilde{T}_1(I_{E}\otimes (\widetilde{T}_1(I_{E}\otimes a)\widetilde{T}^*_1))\widetilde{T}^*_1 \\ &=\widetilde{T}_1(I_{E}\otimes \widetilde{T}_1)(I_{E^{\otimes 2}}\otimes a)(I_{E}\otimes \widetilde{T}^*_1)\widetilde{T}^*_1 \\ & = \widetilde{T}_2(p_2\otimes I_{\mathcal H})(I_{E^{\otimes 2}}\otimes a)(p^*_2\otimes I_{\mathcal H})\widetilde{T}^*_2 \\&=\widetilde{T}_2(I_{X(2)}\otimes a)\widetilde{T}^*_2, \end{align*} for all $a\in\sigma(\mathcal{M})'$. Inductively, we get \begin{align*} \Theta^n_T(a)&=\Theta_T(\Theta^{n-1}_T(a))=\widetilde{T}_1(I_{E}\otimes (\widetilde{T}_{n-1}(I_{X(n-1)}\otimes a)\widetilde{T}^*_{n-1}))\widetilde{T}^*_1 \\&=\widetilde{T}_1(I_{E}\otimes \widetilde{T}_{n-1})(I_{E}\otimes I_{X(n-1)}\otimes a)(I_{E}\otimes \widetilde{T}^*_{n-1})\widetilde{T}^*_1 \\&= \widetilde{T}_n(p_n\otimes I_{\mathcal H})(I_{E}\otimes I_{X(n-1)}\otimes a)(p^*_n\otimes I_{\mathcal H})\widetilde{T}^*_n \\&=\widetilde{T}_n(I_{X(n)}\otimes a)\widetilde{T}^*_n, \end{align*} for all $a\in\sigma(\mathcal{M})'$ and $n\geq 2$. The following is a reformulation of Muhly and Solel's result in our setting \cite[Proposition 2.12]{MR2005882}: \bpp\label{prop6.1} Let $X=(X(n))_{n \in \mathbb{Z_+}}$ be a standard subproduct system of left-finite $W^*$-correspondences over a finite factor $\mathcal{M}$. If $T=(T_n)_{n \in \mathbb{Z_+}}$ is a completely contractive, covariant representation of $X$ on $\mathcal{H}$ then $$tr_{\sigma(\mathcal{M})'}(\Theta^n_{T}(x))\leq \|\widetilde{T}_n\|^2 \: dim_l(X(n))tr_{\sigma(\mathcal{M})'}(x),$$ for all $x \in \sigma(\mathcal{M})_+'$. \epp Let $X=(X(n))_{n \in {\mathbb{Z_+}}}$ be a standard subproduct system of left-finite $W^*$-correspondences over a semifinite factor $\mathcal{M}$. The \textit{curvature} of a completely contractive, covariant representation $T=(T_n)_{n \in \mathbb{Z_+}}$ of $X$ on a Hilbert space $\mathcal{H}$ is defined by \begin{align}\label{eqn_curv} Curv(T)=\lim_{k \to \infty}\frac{tr_{\sigma(\mathcal{M})'}(I-\Theta_T^k(I))}{\sum_{j=0}^{k-1} dim_l(X(j))}, \end{align} if the limit exists. The following result is well known (cf. Popescu \cite[p.280]{MR1822685}). \bl \label{lemp} Let $\{a_j\}_{j=0}^{\infty}$ and $\{b_j\}_{j=0}^{\infty}$ be two real sequences, and let $a_j \geq 0$ and $b_j>0$ for all $j \geq 0$. Consider the partial sums $A_k:=\sum_{j=0}^{k-1}a_j$ and $B_k:=\sum_{j=0}^{k-1}b_j$, and suppose that $B_k \to \infty$ as $k \to \infty$. Then \[ \lim_{k \to \infty}\frac{A_k}{B_k}=L, \] whenever $L:=\lim_{j \to \infty}\frac{a_j}{b_j}$ exists. \el Coming back to our definition of curvatures, we note that \[ \begin{split} tr_{\sigma(\mathcal{M})'}(I-\Theta_T^k(I)) & = \displaystyle \sum^{k-1}_{j=0} tr_{\sigma(\mathcal{M})'}\Theta_T^j(I-\Theta_T(I)) \\ & = \displaystyle \sum^{k-1}_{j=0} tr_{\sigma(\mathcal{M})'}\Theta_T^j(\triangle_*(T)^2). \end{split} \] From this, and our previous lemma, it follows that $Curv(T)$ is well defined whenever the following two conditions are satisfied: \begin{itemize} \item [(1)] $ \lim_{j \to \infty}\frac{tr_{\sigma(\mathcal{M})'}\Theta_T^j(\triangle_*(T)^2)}{ dim_l(X(j))}$ exists, and \item [(2)] $ \lim_{k \to \infty} \sum_{j=0}^{k-1}dim_l(X(j))=\infty$. \end{itemize} The next result concerns the existence of curvatures in the setting of completely contractive, covariant representations on product systems (cf. \cite[Example 1.2]{MR2608451}). This is an analogue of the result by Muhly and Solel \cite[Theorem 3.3]{MR2005882}. The curvature for completely contractive, covariant representation of the standard subproduct system (see Example \ref{eg-3}) will be discussed at the end of this section. \bt Let $X=(X(n))_{n \in \mathbb{Z_+}}$ be a product system of $W^*$-correspondences over a finite factor $\mathcal{M}$, that is, $X(n)=E^{\otimes n}$ where $E:=X(1)$ is a left-finite $W^*$-correspondence. Set $d: =dim_l(E)$. If $T=(T_n)_{n \in \mathbb{Z_+}}$ is a completely contractive, covariant representation of $X$ on $\mathcal{H}$, then the following holds: \begin{itemize} \item [(1)] The limit in the definition of $Curv(T)$ exists, either as a positive number or $+ \infty$. \item [(2)] $Curv(T)=\infty$ if and only if $tr_{\sigma(\mathcal{M})'}(I-\Theta_T(I))=\infty$. \item [(3)] If $tr_{\sigma(\mathcal{M})'}(I-\Theta_T(I)) < \infty$, then $Curv(T) < \infty$. Moreover, in this case we have the following: \begin{enumerate} \item [(3a)] for $d\geq1$ we have \[ Curv(T)=\lim_{k \to \infty} \frac{tr_{\sigma(\mathcal{M})'}(\Theta_{T}^k(I)-\Theta_{T}^{k+1}(I))}{d^k}, \] in particular if $d>1$, then we further get \[ Curv(T)=(d-1)\lim_{k \to \infty}\frac{tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I))}{d^k}; \] \item [(3b)] for $d<1$, $\lim_{k \to \infty}tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I)) < \infty$, and \[ Curv(T)=(1-d)\left(\lim_{k \to \infty}tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I))\right). \] \end{enumerate} \end{itemize} \et \bp From \cite[Theorem 3.3]{MR2005882} it follows that $dim_l(X(j))=d^j.$ Let $ a_k=tr_{\sigma(\mathcal{M})'}(\Theta_{T}^k(I)-\Theta_{T}^{k+1}(I))$ for $k\geq 0$. Then Proposition \ref{prop6.1} yields $$\begin{array}{cll} a_{k+1} & = &tr_{\sigma(\mathcal{M})'}(\Theta_{T}( \Theta_{T}^k(I)-\Theta_{T}^{k+1}(I) ))\\ & \leq & \|\widetilde{T}_1\|^2 dim_l(E) tr_{\sigma(\mathcal{M})'}( \Theta_{T}^k(I)-\Theta_{T}^{k+1}(I) )\\ & \leq& da_k, \end{array} $$ for all $k\geq 0$. If $a_0=\infty$, then the fact that $\{\widetilde{T}_n\widetilde{T}_n^*\}_{n \in \mathbb{Z_+}}$ is a decreasing sequence of positive contractions implies that \[ tr_{\sigma(\mathcal{M})'}(I-\Theta_T^k(I))=\infty \quad \quad (k \geq 0). \] If $a_0<\infty,$ then $\{\frac{a_j}{d^j}\}^{\infty}_{j=0}$ is a non-increasing sequence of non-negative numbers. Then $0\leq L \leq a_0$ where $L := \lim \frac{a_j}{d^j}$. \noindent Let $d\geq1$. Since \[ tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I))=\sum_{j=0}^{k-1}a_j, \] by Lemma \ref{lemp} (for $b_j=d^j$) the limit defining $Curv(T)$ exists and $Curv(T) = L$. \noindent Now let $d>1$. Then $\sum_{j=0}^{k-1}d^j=\frac{d^k-1}{d-1}$ and $\lim_{k \to \infty}\frac{d^k-1}{d^k}=1$ yields $$ \begin{array}{cll} Curv(T) & = & \lim_{k \to \infty }\frac{tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I)) }{\frac{d^k-1}{d-1}} \\ & = & (d-1)\lim_{k \to \infty}\frac{tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I))}{d^k-1} \lim_{k \to \infty}\frac{d^k-1}{d^k} \\ & = & (d-1)\lim_{k \to \infty}\frac{tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I))}{d^k}. \end{array} $$ This proves statement $(3a)$. \noindent Finally, let $d<1$ so that $\sum_{j=0}^{\infty} d^j = 1/(1-d)$. Since $a_j \leq d^j a_0$ for all $j\geq 0$, $\lim_{k \to \infty}tr_{\sigma(\mathcal{M})'}(I-\Theta_{T}^k(I))$ exists and is finite. This completes the proof of $(3b)$. The proof of statements $(1)$ and $(2)$ follows by noting that whenever $a_0$ is finite, the limit defining $Curv(T)$ exists and is finite. \ep Recall that \[ \Theta_T(x)=\widetilde{T_1}(I_{E} \otimes x)\widetilde{T_1}^*, \] for all $x\in\sigma(\mathcal{M})'$, and \[ Q=\lim_{n \to \infty}\widetilde{T}_n\widetilde{T}_n^*=\lim_{n \to \infty}\Theta_{T}^n(I_{\mathcal{H}}). \] Using the intertwining property of the Poisson kernel \[ K(T)^*(S_{n}(\zeta)\otimes I_{\mathcal{D}})=T_n(\zeta)K(T)^*, \] we have $$ \begin{array}{cll} \widetilde{T}_n(I_{X(n)} \otimes K(T)^*)(\zeta \otimes k) & = & \widetilde{T}_n(\zeta \otimes K(T)^*k) \\ & = & T_n(\zeta)K(T)^* k \\ & = & K(T)^*(S_n(\zeta) \otimes I_{\mathcal{D}})k, \end{array} $$ for all $\zeta \in X(n), k\in \mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}, n \in \mathbb{Z_+}$. Then $$ \widetilde{T}_n(I_{X(n)} \otimes K(T)^*)=K(T)^*\widetilde{(S_n(\cdot) \otimes I_{\mathcal{D}})}, $$ and hence $\Theta_T^{n}(Q)=Q$ and \[ K(T)^*K(T)=I_{\mathcal{H}}-Q, \] yields \[ \begin{split} K(T)^* & (I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}-\Theta_{S \otimes I_{\mathcal{D}}}^n (I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}))K(T) \\ & = K(T)^*K(T)- K(T)^*(\widetilde{S_n(\cdot) \otimes I_{\mathcal{D}}})(\widetilde{S_n(\cdot) \otimes I_{\mathcal{D}}})^*K(T) \\ & = K(T)^*K(T)-\widetilde{T}_n(I_{X(n)} \otimes K(T)^*)(\widetilde{S_n(\cdot) \otimes I_{\mathcal{D}}})^* K(T) \\ & = I_{\mathcal{H}}-Q-\widetilde{T}_{n}(I_{X(n)} \otimes K(T)^*)(I_{X(n)} \otimes K(T)) \widetilde{T}_{n}^* \\ & = I_{\mathcal{H}}-Q-\widetilde{T}_{n}(I_{X(n)} \otimes K(T)^*K(T)) \widetilde{T}_{n}^* \\ & =I_{\mathcal{H}}-Q- \widetilde{T}_{n}(I_{X(n)} \otimes (I_\mathcal{H}-Q)) \widetilde{T}_{n}^* \\ & = I_{\mathcal{H}}-Q-\Theta_T^{n}(I_\mathcal{H}-Q) \\ & = I_{\mathcal{H}}-\Theta_T^{n}(I_\mathcal{H}). \end{split} \] Therefore one can compute the curvature, in terms of Poisson kernel, in the following sense: \begin{proposition}\label{proposition1} Let $X=(X(n))_{n \in {\mathbb{Z_+}}}$ be a standard subproduct system of left-finite $W^*$-correspondences over a finite factor $\mathcal{M}$. If $T=(T_n)_{n \in \mathbb{Z_+}}$ is a completely contractive, covariant representation of $X$ on a Hilbert space $\mathcal{H}$, then the curvature of $T$ is given by \begin{align}\label{eqn2} &~~~~~~~~Curv(T)\\&\nonumber =\lim_{k \to \infty}\frac{tr_{\sigma(\mathcal{M})'}(K(T)^*(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}-\Theta_{S \otimes I_{\mathcal{D}}}^k(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}))K(T))}{\sum_{j=0}^{k-1}dim_l(X(j))}, \end{align} if the limit exists. \end{proposition} The following theorem generalizes \cite[Theorem 2.1]{MR1822685}. \begin{theorem}\label{theorem1} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ to be a completely contractive, covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ of $A^*$-correspondences over an $A^*$-algebra $\mathcal{M}$. Then there exist a Hilbert space $\mathcal E,$ a representation $\psi$ of $\mathcal M$ on $\mathcal E,$ and an inner multi-analytic operator $\Pi:\mathcal F_X\bigotimes_\psi\mathcal E \to\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D} $ such that $$ I_{\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D}}-K(T)K(T)^*=\Pi\Pi^*. $$ \end{theorem} \begin{proof} Proposition \ref{prop2} implies that $(ran K(T))^\perp$ is invariant for the covariant representation $S\otimes I_{\mathcal D}.$ Now we use Theorem \ref{Beur} and obtain a Hilbert space $\mathcal E,$ a representation $\psi$ of $\mathcal M$ on $\mathcal E,$ and a partial isometry $\Pi:\mathcal F_X\bigotimes_\psi\mathcal E \to\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D} $ such that $(ran K(T))^\perp$ is the range of $\Pi,$ and $$ \Pi(S_n(\zeta)\otimes I_{\mathcal E})=(S_n(\zeta)\otimes I_{\mathcal{D}})\Pi, $$ for all $\zeta\in X(n)$ and $n\in \mathbb{Z_+}$. Finally, using the fact that $\Pi$ is a partial isometry and \[ (ran K(T))^\perp =ran(I_{\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D}}-K(T)K(T)^*), \] we get the desired formula. \end{proof} The following is an analogue of \cite[Theorem 3.32]{MR2643314} in our context in terms of multi-analytic operators. \bt Let $X=(X(n))_{n \in {\mathbb{Z_+}}}$ be a standard subproduct system of left-finite $W^*$-correspondences over a finite factor $\mathcal{M}$. If $T=(T_n)_{n \in \mathbb{Z_+}}$ is a completely contractive, covariant representation of $X$ on a Hilbert space $\mathcal{H}$, and \begin{align}\label{eq1.2} tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}-\Theta^k_{S \otimes I_{\mathcal{D}}}(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}})) < \infty, \end{align} for all $k \geq 1$, then there exist a Hilbert space $\mathcal E,$ a representation $\psi$ of $\mathcal M$ on $\mathcal E,$ and an inner multi-analytic operator $\Pi:\mathcal F_X\bigotimes_\psi\mathcal E \to\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D} $ such that \begin{align*} &~~~~Curv(T)\\&=\lim_{k \to \infty}\frac{ tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}((I_{\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D}}-\Pi\Pi^*)(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}-\Theta_{S \otimes I_{\mathcal{D}}}^k(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}})))}{\sum_{j=0}^{k-1}dim_l(X(j))}. \end{align*} \et \begin{proof} For simplicity of notation we use $I$ for $I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}$ and also use $\Theta $ for $\Theta_{S \otimes I_{\mathcal{D}}}$. Define a representation $\rho$ of $\mathcal M$ on $\mathcal H\bigoplus \left(\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}\right)$ by \[ \rho(a)=\begin{pmatrix} {\sigma(a)} & {0} \\ {0} & \phi_{\infty}(a)\otimes I_{\mathcal D} \end{pmatrix}, \] for all $a \in \mathcal M$. Then \begin{align} \nonumber & tr_{\sigma(\mathcal{M})'}(K(T)^*(I-\Theta^k(I))K(T)) \\ \nonumber&=tr_{\rho(\mathcal{M})'}\begin{pmatrix} {K(T)^*(I-\Theta^k(I))K(T)} & {0} \\ {0} & 0 \end{pmatrix} \\ \nonumber&=tr_{\rho(\mathcal{M})'}\left(\begin{pmatrix} {0} & {K(T)^*(I-\Theta^k(I))^{\frac{1}{2}})} \\ {0} & 0 \end{pmatrix} \begin{pmatrix} {0} & {0} \\ {(I-\Theta^k(I))^{\frac{1}{2}}K(T)} & 0 \end{pmatrix}\right) \\ \nonumber&=tr_{\rho(\mathcal{M})'}\left( \begin{pmatrix} {0} & {0} \\ {(I-\Theta^k(I))^{\frac{1}{2}}K(T)} & 0 \end{pmatrix}\begin{pmatrix} {0} & {K(T)^*(I-\Theta^k(I))^{\frac{1}{2}})} \\ {0} & 0 \end{pmatrix} \right) \\ \nonumber&=tr_{\rho(\mathcal{M})'}\begin{pmatrix} {0} & {0} \\ {0} & (I-\Theta^k(I))^{\frac{1}{2}}K(T)K(T)^*(I-\Theta^k(I))^{\frac{1}{2}} \end{pmatrix} \\ \label{eqna1}&= tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}((I-\Theta^k(I))^{\frac{1}{2}}K(T)K(T)^*(I-\Theta^k(I)) ^{\frac{1}{2}}). \end{align} Now by Theorem \ref{theorem1}, there exist a Hilbert space $\mathcal E,$ a representation $\psi$ of $\mathcal M$ on $\mathcal E$, and an inner multi-analytic operator $\Pi:\mathcal F_X\bigotimes_\psi\mathcal E \to\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D}$ such that \begin{align*} &~~~~Curv(T)\\& =\lim_{k \to \infty}\frac{ tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}((I_{\mathcal{F}_X\bigotimes_{\sigma}\mathcal{D}}-\Pi\Pi^*)(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}}-\Theta_{S \otimes I_{\mathcal{D}}}^k(I_{\mathcal{F}_X \bigotimes_{\sigma} \mathcal{D}})))}{\sum_{j=0}^{k-1}dim_l(X(j))}. \end{align*} Then equation (\ref{eqna1}) and Proposition \ref{proposition1} yields \begin{align*} Curv(T) =&\lim_{k \to \infty}\frac{tr_{\sigma(\mathcal{M})'}(K(T)^*(I-\Theta^k(I))K(T))}{\sum_{j=0}^{k-1}dim_l(X(j))} \\ =&\lim_{k \to \infty}\frac{ tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}((I-\Theta^k(I))^{\frac{1}{2}}K(T)K(T)^*(I-\Theta^k(I))^{\frac{1}{2}})}{\sum_{j=0}^{k-1}dim_l(X(j))} \\ = &\lim_{k \to \infty}\frac{ tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'} ((I-\Theta^k(I))^{\frac{1}{2}}(I-\Pi\Pi^*)(I-\Theta^k(I))^{\frac{1}{2}})}{\sum_{j=0}^{k-1}dim_l(X(j))} \\ =&\lim_{k \to \infty}\frac{ tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}((I-\Pi\Pi^*)(I-\Theta^k(I)))}{\sum_{j=0}^{k-1}dim_l(X(j))}. \end{align*} The third equality follows from the observation that: since $tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}(I-\Theta^k(I))$ is finite, $(I-\Theta^k(I))^{\frac{1}{2}}$ belongs to the ideal $$\{x:tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}(x^*x)<\infty\},$$ and hence $tr_{(\phi_{\infty}(\mathcal{M})\otimes I_{\mathcal D})'}(I-\Theta^k(I))^{\frac{1}{2}}<\infty$, for all $k$. \qedhere \end{proof} \begin{remark} Consider the standard subproduct system of Example \ref{eg-3}, and let $\mbox{dim~} X(1) = d<\infty$. It is easy to verify that \begin{align*} dim_l(X(j))&= \binom{j+d-1}{j}, \end{align*} for all $j \geq 1$. By induction it follows that \begin{align*} \displaystyle \sum^{k-1}_{j=0} dim_l(X(j))&=\frac{k(k+1)\ldots (k+d-1)}{d!}\sim\frac{k^d}{d!} \quad \quad (k \geq 1). \end{align*} Therefore, in this case, our curvature defined in (\ref{eqn_curv}) coincides with the Arveson's curvature for the row contraction \[ (T_1(e_1),T_1(e_2),\ldots,T_1(e_d)) \] (under the finite rank assumption, that is, $rank~(I_{\mathcal H} - \sum_{i=1}^d T_1(e_i) T_1(e_i)^*) < \infty$) on the Hilbert space $\mathcal H$ (cf. \cite[Theorem C]{A2000}). One can also compare the curvature obtained in (\ref{eqn2}) with Popescu's curvature (cf. \cite[Equation 2.11]{MR1822685}). Furthermore, observe that the condition (\ref{eq1.2}) is automatic for finite rank row contractions. \end{remark} \section{Wandering subspaces} The notion of wandering subspaces of bounded linear operators on Hilbert spaces was introduced by Halmos \cite{PH}. With this as a motivation we extend the notion of wandering subspace (cf. \cite[p. 561]{H16}) for covariant representations of standard subproduct systems, as follows: Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$. A closed subspace $\mathcal{S}$ of $\mathcal{H}$ is called \textit{wandering} for the covariant representation $T$ if it is $\sigma(\mathcal M)$-invariant, and if for each $n \in \mathbb{N}$ the subspace $\mathcal S$ is orthogonal to \[ \mathfrak{L}_n(\mathcal{S},T):=\bigvee \{T_n(p_n(\zeta))s\: : \: \zeta \in E^{\otimes n},s \in \mathcal{S}\}. \] When there is no confusion we use the notation $\mathfrak{L}_n(\mathcal{S})$ for $\mathfrak{L}_n(\mathcal{S},T),$ and also use $\mathfrak{L}(\mathcal{S})$ for $\mathfrak{L}_1(\mathcal{S})$. A wandering subspace $\mathcal W$ for $T$ is called \textit{generating} if $\mathcal H=\overline{\mbox{span}} \{\mathfrak L_n (\mathcal W):n\in\mathbb Z_+\}$. In the following proposition we prove that the wandering subspaces are naturally associated with invariant subspaces of covariant representations of standard subproduct systems. \bpp \label{prop5.1} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$. If $\mathcal S$ is a closed $T$-invariant subspace of $\mathcal{H},$ then $\mathcal{S}\ominus\mathfrak{L}(\mathcal{S})$ is a wandering subspace for $T|_{\mathcal S}:=({T_n}|_{\mathcal S})_{n\in \mathbb{Z_+}}.$ \epp \bp Let $n \geq 1$ and $\eta=\xi_1\otimes\xi_{n-1} \in E^{\otimes n}$ for some $\xi_1 \in E$ and $\xi_{n-1}\in E^{\otimes n-1}$. Let $x, s \in \mathcal{S}\ominus \mathfrak{L}(\mathcal{S})$ so that $ y=T_n(p_n(\eta))s \in \mathfrak{L}_n(\mathcal{S}\ominus\mathfrak{L}(\mathcal{S}))$. Then \[ \begin{split} \langle x, y \rangle & = \langle x ,T_n(p_n(\eta))s\rangle \\ & =\langle x, T_n(p_n(\xi_1 \otimes \xi_{n-1}))s \rangle \\ & =\langle x,T_{1+(n-1)}(p_{1+(n-1)}(\xi_1 \otimes \xi_{n-1}))s\rangle \\ & = \langle x,T_1(\xi_1)T_{n-1}(\xi_{n-1})s\rangle \\ & =0, \end{split} \] since $\mathcal{S}$ in invariant under $T_{n-1}(\xi_{n-1})$. Therefore $ \mathcal{S}\ominus \mathfrak{L}(\mathcal{S})$ is orthogonal to $\mathfrak{L}_n(\mathcal{S}\ominus\mathfrak{L}(\mathcal{S}))$, $n \geq 1$ and hence $\mathcal{S}\ominus\mathfrak{L}(\mathcal{S})$ is a wandering subspace for $T|_{\mathcal S}=({T_n}|_{\mathcal S})_{n\in \mathbb{Z_+}}$. \ep Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$. Suppose $\mathcal W$ is a wandering subspace for $T.$ Set \[ \mathcal G_{T,\mathcal W}:=\bigvee_{n\in\mathbb Z_+}\mathfrak L_n(\mathcal W). \] Note that \begin{align*} &\mathfrak L\left(\bigvee_{n\in\mathbb Z_+}\mathfrak L_n(\mathcal W)\right) \\& = \overline{\mbox{span}}\{T_1(p_1(\zeta))T_n(p_n(\eta))w:\zeta\in E, \eta\in E^{\otimes n},w\in \mathcal W, n\in\mathbb Z_+\} \\ & = \overline{\mbox{span}}\{T_{n+1}(p_{n+1}(p_1(\zeta)\otimes p_n(\eta))w:\zeta\in E, \eta\in E^{\otimes n}, w\in \mathcal W, n\in\mathbb Z_+\}\\ &\subset \bigvee_{n\in\mathbb N}\mathfrak L_n(\mathcal W). \end{align*} In the other direction, we have \begin{align*} &\bigvee_{n\in\mathbb N}\mathfrak L_n(\mathcal W) \\&= \overline{\mbox{span}}\{T_{n}(p_{n}(p_1(\zeta)\otimes p_{n-1}(\eta))w:\zeta\in E, \eta\in E^{\otimes {n-1}},w\in \mathcal W, n\in\mathbb N\} \\ & = \overline{\mbox{span}}\{T_{1}(p_1(\zeta))T_{n-1}( p_{n-1}(\eta))w:\zeta\in E, \eta\in E^{\otimes {n-1}},w\in \mathcal W, n\in\mathbb N\} \\ &\subset \mathfrak L\left(\bigvee_{n\in\mathbb Z_+}\mathfrak L_n(\mathcal W)\right). \end{align*} Thus these sets are equal, and it follows that \begin{align*} \mathcal G_{T,\mathcal W}\ominus \mathfrak L(\mathcal G_{T,\mathcal W})=\bigvee_{n\in\mathbb Z_+}\mathfrak L_n(\mathcal W)\ominus \mathfrak L\left(\bigvee_{n\in\mathbb Z_+}\mathfrak L_n(\mathcal W)\right)=\mathcal W. \end{align*} Hence we have the following uniqueness result: \begin{proposition}\label{prop6} Let $T=(T_n)_{n \in \mathbb{Z_+}}$ be a covariant representation of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}}$ over an $A^*$-algebra $\mathcal M$. If $\mathcal W$ is a wandering subspace for $T,$ then $$\mathcal W=\mathcal G_{T,\mathcal W}\ominus \mathfrak L(\mathcal G_{T,\mathcal W}).$$ Moreover, if $\mathcal W$ is also generating, then $\mathcal W=\mathcal H\ominus \mathfrak L(\mathcal H).$ \end{proposition} In Theorem \ref{thm1} we observed that each non-trivial closed subspace $\mathcal S\subset\mathcal H,$ which is invariant under a pure completely contractive, covariant representation $T=(T_n)_{n \in \mathbb{Z_+}}$ of a standard subproduct system $X=(X(n))_{n \in \mathbb{Z_+}},$ can be written as $\mathcal S=\Pi(\mathcal{F}_X\bigotimes_{\pi}\mathcal{D}).$ In the following theorem we study wandering subspaces in a general situation when $T$ is not necessarily pure. \bt \label{Theorem 5.3} Let $X=(X(n))_{n \in \mathbb{Z}_+}$ be a standard subproduct system over an $A^*$-algebra $\mathcal M$. Let $\pi:\mathcal M\to B(\mathcal E)$ be a representation on a Hilbert space $\mathcal E$ and $T=(T_n)_{n \in \mathbb{Z}_+}$ be the covariant representation of $X$. Let $\Pi: \mathcal{F}_X \bigotimes_\pi \mathcal{E} \to \mathcal{H}$ be a partial isometry such that $\Pi(S_n(\zeta)\otimes I_{\mathcal{E}})=T_n(\zeta)\Pi$ for every $\zeta \in X(n), n \in \mathbb{Z}_+$. Then $\mathcal{S}:=\Pi(\mathcal{F}_X \bigotimes_\pi \mathcal{E})$ is a closed $T$-invariant subspace, $\mathcal{W}:=\mathcal{S} \ominus \mathfrak{L}(\mathcal{S})$ is a wandering subspace for $T|_{\mathcal{S}},$ and $\mathcal{W}=\Pi((ker \Pi)^{\bot}\bigcap \mathcal{M}\bigotimes_\pi \mathcal{E})$. \et \bp Define $F=(ker \Pi)^{\bot}\bigcap \mathcal{M}\bigotimes_\pi \mathcal{E}$. Since $\mathcal{S}$ is the range of $\Pi,$ it is a closed $T$-invariant subspace. Therefore by Proposition \ref{prop5.1}, the subspace $\mathcal{W}$ is a wandering subspace for $T|_{\mathcal{S}}$. $$ \begin{array}{cl} &\mathfrak{L}(\mathcal S, T) \\& = \mathfrak{L}(\Pi(\mathcal{F}_X \bigotimes_\pi \mathcal{E}),T) \\& = \bigvee\{ \: T_1(\zeta)k \:\: : \:\: k \in\Pi(\mathcal{F}_X \bigotimes_\pi \mathcal{E}),\zeta\in X(1) \} \\ & = \bigvee\{ \: T_1(\zeta)\Pi(l) \:\: : \:\: l \in \mathcal{F}_X \bigotimes_\pi \mathcal{E},\zeta\in X(1) \}\\ & = \bigvee\{ \: \Pi(S_1(\zeta) \otimes I_{\mathcal{E}})(l_m\otimes e) \:\: : \:\: l_m\otimes e \in X(m) \bigotimes_\pi \mathcal{E},\zeta\in X(1), m \in \mathbb{Z}_+ \}. \end{array} $$ For $x \in (ker \Pi)^{\bot}\bigcap \mathcal{M}\bigotimes_\pi \mathcal{E}$ and $ l_m\otimes e \in X(m) \bigotimes_\pi \mathcal{E}$ we have $$ \begin{array}{ll} \langle \Pi x,\Pi(S_1(\zeta) \otimes I_{\mathcal{E}})(l_m\otimes e) \rangle & = \langle \Pi ^*\Pi x,(S_1(\zeta) \otimes I_{\mathcal{E}})(l_m\otimes e)\rangle \\ & = \langle x,(S_1(\zeta) \otimes I_{\mathcal{E}})(l_m\otimes e)\rangle\\&= \langle x,P_{1+m}(\zeta\otimes l_m) \otimes e\rangle \\ & =0, \end{array} $$ and hence $\Pi((ker \Pi)^{\bot}\bigcap \mathcal{M}\bigotimes_\pi \mathcal{E})\subset\mathcal W$. \noindent For the converse direction, let $x \in \mathcal{S} \ominus \mathfrak{L}(\mathcal{S},T)=\mathcal W$, and $\Pi(y)=x$ for some $y\in (ker\Pi)^{\bot}$. Therefore for any $\zeta \in X(1),\eta \otimes e\in \mathcal{F}_X\bigotimes_\pi \mathcal E$ we have \begin{eqnarray}\label{err1} \langle y,(S_1(\zeta)\otimes I_{\mathcal{E}})(\eta \otimes e)\rangle&=&\langle \Pi y,\Pi(S_1(\zeta)\otimes I_{\mathcal{E}})(\eta \otimes e)\rangle=0. \end{eqnarray} Recall that by definition we have \begin{align*} &\mathfrak L(\mbox{$\mathcal{F}_X \bigotimes_\pi \mathcal{E}$},S\otimes I_{\mathcal E})\\&=\bigvee\{(S_1(\zeta) \otimes I_{\mathcal{E}})(\eta \otimes e) : \eta \in X(m), \zeta \in X(1), e \in \mathcal{E}, m \in \mathbb{Z}_+\}. \end{align*} Since $\mathcal M\bigotimes_\pi \mathcal E$ is a generating wandering subspace for the covariant representation $S\otimes I_{\mathcal E},$ it follows from Proposition \ref{prop6} that $(\mathcal{F}_X \bigotimes_\pi \mathcal{E}) \ominus\mathfrak L(\mathcal{F}_X \bigotimes_\pi \mathcal{E},S\otimes I_{\mathcal E})=\mathcal M\bigotimes_\pi\mathcal{E}$, and hence (\ref{err1}) implies that $y\in \mathcal M\bigotimes_\pi\mathcal{E}$. Hence we get $\Pi((ker \Pi)^{\bot}\bigcap \mathcal{M}\bigotimes_\pi \mathcal{E})=\mathcal W$. \ep Since each commuting tuple of operators defines a covariant representation, the previous theorem is a generalization of \cite[Theorem 5.2]{BEDS}. Indeed, we get the following corollary: \begin{corollary} With the same notation of Theorem \ref{Theorem 5.3} we have $$\bigvee_{n \in \mathbb{Z_+}}\mathfrak L_n(\mathcal W,T)=\Pi\left(\bigvee_{n \in \mathbb{Z_+}}\mathfrak{L}_n(F,S\otimes I_{\mathcal E})\right )$$ where $F=(ker\Pi)^{\bot}\bigcap \mathcal{M} \mbox{$\bigotimes_{\pi}$}\mathcal{E}$. Moreover, $F$ is wandering subspace for the representation $S\otimes I_{\mathcal{E}}$, that is, $F\bot \mathfrak{L}_n(F,S\otimes I_{\mathcal E})$ for each $n\in\mathbb N.$ \end{corollary} \bp For each $f,f' \in F $ we have \begin{eqnarray*}\langle f, (S_n(\zeta)\otimes I_{\mathcal{E}})f'\rangle&=& \langle \Pi^{*}\Pi f,(S_n(\zeta)\otimes I_{\mathcal{E}})f' \rangle=\langle \Pi f,\Pi(S_n(\zeta)\otimes I_{\mathcal{E}})f' \rangle\\&=&0. \end{eqnarray*} Therefore $F$ is wandering subspace for the representation $S\otimes I_{\mathcal E}$. Moreover, since $\mathcal{W}=\Pi F$, we have that \[ \begin{split} \bigvee\{\mathfrak L_n(\mathcal W,T) : n \in \mathbb{Z_+}\} & = \bigvee\{(T_n(\zeta)\Pi(F) : \zeta \in X(n), n \in \mathbb{Z_+}\} \\ &= \bigvee\{\Pi(S_n(\zeta)\otimes I_{\mathcal{E}})(F) : \zeta \in X(n), n \in \mathbb{Z_+}\} \\ &= \Pi(\bigvee \{ (S_n(\zeta)\otimes I_{ \mathcal{E}})(F) : \zeta \in X(n), n \in \mathbb{Z_+}\})\\ &= \Pi(\bigvee\{\mathfrak{L}_n(F,S\otimes I_{\mathcal E}) : n \in \mathbb{Z_+}\} ).\qedhere \end{split} \] \ep \vspace{0.5cm} \noindent{\bf Acknowledgements:} We would like to sincerely thank the anonymous referee for valuable suggestions that improved the content of the Section 4. The research of Sarkar was supported in part by (1) National Board of Higher Mathematics (NBHM), India, grant NBHM/R.P.64/2014, and (2) Mathematical Research Impact Centric Support (MATRICS) grant, File No : MTR/2017/000522, by the Science and Engineering Research Board (SERB), Department of Science \& Technology (DST), Government of India. Trivedi thanks Indian Statistical Institute Bangalore for the visiting scientist fellowship. Veerabathiran was supported by DST-Inspire fellowship.
a07fec5bb6a5fbda4496469b05e0982e399eb377
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Appendix: $\< TT \>$ at 2-loops} The holographic formula for the power spectrum reads, \begin{equation} \label{power_sp_h} \Delta^2_{{\cal R}}(q) = -\frac{q^3}{4 \pi^2} \frac{1}{{\rm Im} \<\!\<T(q)T(-q)\>\!\>} , \end{equation} where $T=T^i_i$ is the trace of the energy momentum tensor and the double bracket notation indicates that the momentum conserving delta function (times $(2 \pi)^3$) has been removed. The imaginary part in (\ref{power_sp_h}) is taken after analytic continuation \begin{equation} \label{analytic} q \to -i q, N \to -i N, \end{equation} where $q$ is the magnitude of the momentum. This formula was derived in \cite{McFadden:2009fg} using the domain-wall/cosmology correspondence and also follows from the wave-function of the universe approach. There is a similar holographic formula for the tensor power spectrum involving the transverse traceless part of the 2-point function of $T_{ij}$. This class of theories we consider has the important property that if one promotes $g_{YM}^2$ to a new field that transforms appropriately under conformal transformation, the theory becomes conformally invariant \cite{Jevicki:1998ub, Kanitscheider:2008kd}. We say that the theory has a ``generalized conformal structure''. This is not a bona fide symmetry of the theory but, nevertheless, controls many of its properties. The generalised conformal structure implies that the 2-point function of the energy momentum tensor, to leading order in the large $N$ limit (planar diagrams), is given by \begin{eqnarray} \<\!\< T(q) T(-q) \>\!\> = q^3 N^2 f(\geff^2), \end{eqnarray} where $\geff^2=\gYM^2 N/q$ is the effective dimensionless 't Hooft coupling constant and $f(\geff^2)$ is function of $\geff^2$ \cite{Kanitscheider:2008kd}. The overall factor of $q^3$ reflects the fact that the energy momentum has dimension 3 in three dimensions and the overall factor of $N^2$ is because this is the leading order term in the large $N$ limit. Under the analytic continuation (\ref{analytic}), \begin{equation} \geff^2(q) \to \geff^2(q), \quad N^2 q^3 \to -i N^2 q^3 \end{equation} and therefore for this class of theories, \begin{equation} \label{power_sp3} \Delta^2_{{\cal R}}(q) = \frac{1}{4 \pi^2 N^2 f(g_{eff}^2)}, \end{equation} which is the formula we used in the main text. Perturbation theory is valid when $g_{\rm eff}^2 \ll 1$. In the perturbative regime, the function $f(\geff^2)$ is given by \begin{equation} \label{app:2loop} f(\geff^2) = f_0 \left( 1 - f_1 \, \geff^2 \ln \geff^2 + f_2 \, \geff^2 + O(\geff^4) \right). \end{equation} The function $f_0$ is determined by a 1-loop computation, while $f_1$ and $f_2$ come from 2-loops. The presence of the logarithm is due to UV and IR divergences in the computation of the 2-point function of the energy momentum tensor. The 1-loop computation was done in \cite{McFadden:2009fg, McFadden:2010na} and we summarize the 2-loop computation of \cite{CDS} here. Since this a gauge theory we first need to gauge fix. The gauged fixed action is \begin{eqnarray} S &=& \frac{1}{\gYM^2} \int d^3 x \, \tr \left[ \frac{1}{2} F_{ij} F^{ij} + (\partial^i A_i)^2 + 2 \partial^i \bar{c} \mD_i c \right. \nn \\ && + (\mD \Phi^M)^2 + 2 \bar{\psi}^L \mD \psi^L + 2 \sqrt{2} \, \mu_{M L_1 L_2} \Phi^M \bar{\psi}^{L_1} \psi^{L_2} \nn \\ && \left. + \frac{1}{6} \lambda_{M_1 M_2 M_3 M_4} \Phi^{M_1} \Phi^{M_2} \Phi^{M_3} \Phi^{M_4} \right] . \end{eqnarray} where $b, c$ is the ghost sector. The energy-momentum tensor is obtained by coupling the theory to a background metric $g_{ij}$ and then using $T_{ij} = (2/\sqrt{g}) (\delta S/\delta g^{ij}) |_{g_{ij}=\delta_{ij}}$. This procedure defines a unique energy momentum tensor, except that there is a choice of how to couple the scalars to gravity. One may include the non-minimal coupling $1/(2 \gYM^2) \sum_M \xi_M R (\Phi^M)^2$ in the action, where the sum is over all scalars and different scalars may have a different $\xi_M$. The variation of this term contributes an ``improvement term'' to the energy momentum tensor. If $\xi_M=0$ the scalar is a called a minimal scalar, while if $\xi_M=1/8$ it is a conformal scalar. The energy momentum obtained in this fashion is given by \begin{equation} T_{ij} = T^A_{ij} + T^{g.f.}_{ij} + T^{gh}_{ij} + T^{\psi}_{ij} + T^{\Phi}_{ij} + T^{Y}_{ij}, \end{equation} where the different terms denote, respectively, the contribution of the gauge fields, the gauge-fixing term , the ghost sector, the fermions, the scalars and the Yukawa interactions. These are explicitly given by \begin{eqnarray} \label{T} T^{A}_{ij} &=& \frac{1}{\gYM^2} \tr \left[ F_{ik} F_{jk} - \delta_{ij} \frac{1}{4} F_{kl} F_{kl} \right] , \\ T^{g.f.}_{ij} &=& \frac{1}{ \gYM^2} \tr \left[ A_i \partial_j \left( \partial_k A_k\right) + A_j \partial_i \left( \partial_k A_k\right) \right. \nonumber \\ &&\left. - \delta_{ij} \left( A_k\ \partial_k \partial_l A_l + \frac{1}{2} (\partial_k A_k) (\partial_l A_l) \right)\right] , \nn \\ T^{gh}_{ij} &=& \frac{1}{\gYM^2} \tr \left[ \partial_i \bar{c} \mD_j c + \partial_j \bar{c} \mD_i c - \delta_{ij} \partial_k \bar{c} \mD_k c\right] , \nn \\ T^{\psi}_{ij} &=& \frac{1}{\gYM^2} \tr \left[ \frac{1}{2} \bar{\psi}^L \gamma_{(i} \oa{\mD}_{j)} \psi^L - \delta_{ij} \frac{1}{2} \bar{\psi}^L \gamma_k \oa{\mD}_k \psi^L \right] ,\nn \\ T^{\Phi}_{ij} &=& \frac{1}{\gYM^2} \tr \left[ \mD_i \Phi^M \mD_j \Phi^M - \delta_{ij} \left( \frac{1}{2} (\mD \Phi^M)^2 \right. \right. \nonumber \\ && \left. \qquad \qquad + \frac{1}{4!} \lambda_{M_1 M_2 M_3 M_4} \Phi^{M_1} \Phi^{M_2} \Phi^{M_3} \Phi^{M_4} \right) \nonumber \\ && \left. \qquad \qquad + \, \xi_M \left( \delta_{ij} \partial^2 - \partial_i \partial_j\right) (\Phi^M)^2\right] , \nn \\ T^{Y}_{ij} &=& \frac{1}{\gYM^2} \tr \left[ - \delta_{ij} \, \, \mu_{M L_1 L_2} \Phi^M \bar{\psi}^{L_1} \psi^{L_2} \right] . \nn \end{eqnarray} The topology of the 2-loop diagrams that need to be computed is given in Fig. \ref{Fig.TT2L}. \begin{figure}[h] \centering \includegraphics[scale=1]{TT2Lc} \hspace{0.5cm} \includegraphics[scale=1]{TT2Lb} \hspace{0.5cm} \includegraphics[scale=1]{TT2La} \hspace{0.5cm} \includegraphics[scale=1]{TT2Ld} \hspace{0.5cm} \includegraphics[scale=1]{TT2Le} \caption{2-loop diagrams contributing to $\< TT \>$. The blob in the fourth diagram represents an insertion of a 1-loop self-energy. \label{Fig.TT2L}} \end{figure} The 2-point function has a UV divergence that can be cancelled by the counterterm \begin{equation} a_{CT} \int d^3 x \sqrt{g} R, \end{equation} where $R$ is the curvature scalar of the background metric $g_{ij}$, with appropriately chosen $a_{CT}$. As usual, this process introduces a renormalization scale $\mu$, which leads to scheme dependence. The correlator also has an IR divergence (unless all scalars are minimal), which we regulated with an IR cut-off, $\mu_*$. This leads to the following result for $\ln \beta$, \begin{equation} \ln \beta=\ln \frac{1}{|g|} -\frac{a_0}{f_1} - \left(\ln \frac{q_*}{\mu} + \frac{64/\pi^2}{f_1 \mN_{(B)} }\Sigma_\Phi \ln \frac{\mu}{\mu_*}\right), \end{equation} where we have made use of the definition of $g$, $g q_* = f_1 \gYM^2 N$. The renormalization scale $\mu$ is arbitrary and so is the pivot scale $q_*$. We fix this scheme dependence by setting, $\mu=q_*$. An alternative scheme is to set $\mu$ to the (inverse of the) smallest scale in the data (i.e. equal to 0.17 Mpc${}^{-1}$). We have checked that the results we present in the main text are not sensitive to the choice of scheme, except possibly at very low $l$'s. As argued in the main text, this is precisely the regime where one should not trust the perturbative computation. Regarding the IR divergence now. It was argued in \cite{Jackiw:1980kv, Appelquist:1981vg} that superrenormalizable theories with a dimensionful coupling constant (as in our case) are non-perturbative IR finite, with $\gYM^2$ providing the IR cut-off . We therefore set $\mu_*= c \gYM^2$, where $c$ is a number that can only be determined non-perturbatively. This leads to our final formula for $\ln \beta$, \begin{equation} \label{app:final_log} \ln \beta=\ln \frac{1}{|g|} -\frac{a_0}{f_1} - \frac{64/\pi^2}{ f_1 \mN_{(B)} }\Sigma_\Phi \ln \frac{N f_1}{c g} \end{equation} In the main text we set $c=1$ but we also checked that the results do not change qualitatively if we change $c$. An alternative way to deal with the IR issues is to get $\mu_*$ equal to the (inverse of the) largest scale in the data (i.e. $1.4 \times 10^{-4}$ Mpc${}^{-1}$). As in the case of scheme dependence, we have checked that the results are not very sensitive to how we treat $\mu_*$, except possibly at very low $l$s.
121d84f876a5f476fc2969fbf1d60ec480453317
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The study of mixtures of Bose-Einstein condensates (BECs) has opened rich opportunities for novel experimental and theoretical investigations. Mixtures of ultracold atoms offer great flexibility thanks to the variety of atomic species and the additional degree of freedom related to the hyperfine structure \cite{Myatt97,Stenger98,Pu98,Ho98,Timmermans98,Ohmi98} (for a recent overview see \cite{Stamper13}). For a weakly interacting mixture of two BECs, the ground state of the system can either be a miscible mixture of the two components or a phase separated configuration \cite{Colson78}. Nevertheless, the stability of mixtures very close to the critical region is sensitive to other effects, such as asymmetries in the trapping potential \cite{Kevrekidis07}. Moreover, for systems in which the intracomponent coupling constants do not exactly coincide, one of the two components will experience a positive buoyancy and will ``float'' on the other. Previous experiments involving two internal states of rubidium were affected by both of these problems \cite{Hall98,Weld10,Hamner11,Nicklas15,Eto16} hence setting strong limits to explore the many-body properties of miscible binary BECs. In particular, such conditions prevent the study of the static and dynamic response of an unpolarized system close to the transition between the miscible and immiscible phases, where interaction effects are particularly important despite the weakly interacting nature of the gas \cite{Jezek02}. Here we report on the first measurement of the spin-dipole (SD) polarizability of a two-component BEC, as well as the frequency of the SD oscillation, by using an ultracold mixture of the $|3^2S_{1/2}, F=1, m_F=+1\rangle \equiv \ket{\uparrow}$ and $|3^2S_{1/2}, F=1, m_F=-1\rangle \equiv \ket{\downarrow}$ states of atomic sodium. The polarizability characterizes in a fundamental way the thermodynamic behavior of binary ultracold gases and exhibits a divergent behavior at the transition between the miscible and immiscible phases, with the occurrence of important spin fluctuations \cite{Recati11,Abad13,Bisset15}. On the other hand, the SD oscillation is the simplest collective excitation supported by the system in the presence of harmonic trapping and is characterized by the motion of the two components with opposite phase around equilibrium. The SD oscillation is the analog of the famous giant dipole resonance of nuclear physics, where neutrons and protons oscillate with opposite phase \cite{Bohr69}. Actually, collective modes are a popular subject of research in quantum many body systems (see, \emph{e.g.}, \cite{Pitaevskii16}) where experiments are able to determine the corresponding frequencies with high precision, providing a good testbed for detailed comparison with theory and an accurate determination of the relevant interaction parameters. Collective dynamics has been already investigated in quantum binary mixtures of atomic gases like repulsive gases of Fermi atoms \cite{Vichi99,DeMarco02,Recati11,Valtolina16}, Bose-Bose \cite{Hall98,Sinatra99,Maddaloni00,Modugno01,Modugno02,Jezek02,Mertes07,Nicklas11,Zhang12,Egorov13,Sartori13,Sartori15} and Bose-Fermi mixtures \cite{Ferlaino03} as well as Bose-Fermi superfluid mixtures \cite{Ferrier14,Roy16}. In the case of Bose-Bose mixtures both the polarization and the SD oscillation frequency are predicted to be crucially sensitive to the difference between the value of the intra and intercomponent interactions \cite{Jezek02,Sartori15} which is particularly small in our case. The dramatic change of the density profile of the trapped gas, caused by a small displacement of the minima of the trapping potentials of the two species near the miscible-immiscible phase transition, was first investigated theoretically in \cite{Jezek02}. Our mixture is not subject to buoyancy as $g_{\uparrow\uparrow}=g_{\downarrow\downarrow}\equiv g$ and is on the miscible side $g_{\uparrow\downarrow} < g$ near the boundary of the phase transition ($g$ and $g_{\uparrow\downarrow}$ are respectively the intra and intercomponent coupling constants). The fact that $(g - g_{\uparrow\downarrow}) / g \simeq 7 \, \%$, as given by the scattering lengths $a_{\uparrow\uparrow}=a_{\downarrow\downarrow}=54.54(20) a_0$ and $a_{\uparrow\downarrow}=50.78(40)a_0$, where $a_0$ is the Bohr radius \cite{Knoop11}, ensures the stability of the mixture and, together with the absence of buoyancy, allows us to overcome the ultimate limits to measure both the polarizability and SD oscillation frequency. \section{Mixture preparation} Our experiment is based on the apparatus introduced in \cite{Lamporesi13} and starts with a nearly pure BEC of $^{23}$Na atoms in the $\ket{\downarrow}$ state in a crossed optical dipole trap with frequencies $\left[ \omega_x,\omega_y,\omega_z \right] / 2 \pi = \left[47.7(2),207.2(3),156.8(2) \right] \, \text{Hz}$. The magnetic fields along the three spatial directions are calibrated with a precision of $1 \, \text{mG}$ using RF spectroscopy techniques. The first step towards the creation of the spin mixture is to perform a Landau--Zener transition to the $|F=1, m_F=0\rangle \equiv \ket 0 $ state with nearly $100 \, \%$ transfer efficiency. This is realized at a magnetic field of $100 \, \text{G}$ to isolate a two-level system exploiting the quadratic Zeeman shifts. The second step consists in inducing a Rabi oscillation among the three Zeeman sublevels to obtain a 50/50 spin mixture of $\ket{\downarrow}$ and $\ket{\uparrow}$ \cite{Zibold16}. The bias field along $\hat x$ is taken small enough to allow us to neglect the quadratic Zeeman shifts compared to the Rabi frequency and is kept on during the whole experimental sequence following the Rabi pulse. The number of atoms in each spin component is $N_\uparrow = N_\downarrow \simeq 10^6$ and the total chemical potential of the cloud is $\mu_{\text{tot}}/k_{\text{B}} \simeq 200 \, \text{nK}$. Fig. \ref{Fig1}(a) shows typical absorption images of the spinor BEC after a $10 \, \text{ms}$ Stern--Gerlach (SG) expansion in a magnetic field gradient along $\hat z$. In order to prevent the decay of the mixture to $|0\rangle$ by spin changing collisions, we lift this level by $\sim 10 \, \text{kHz}$ using blue detuned microwave dressing on the transition to $|F=2, m_F=0\rangle$ (see Fig. \ref{Fig1}(b)). \begin{figure}[t] \centerline{{\includegraphics[width=0.5\textwidth]{Fig1.pdf}}} \caption{(a) Absorption images taken after a SG expansion for I) the dipole loading, II) the Landau--Zener transition, III) the Rabi pulse leading to the creation of the binary mixture. (b) Stabilization of the two components by shifting the $|0\rangle$ state using microwave dressing on the transition to $|F=2, m_F=0\rangle$. } \label{Fig1} \end{figure} \section{Spin-dipole polarizability} The SD polarizability of a spin mixture describes the ability of the system to adapt itself to a displacement in opposite direction of the trapping potentials of the two components. After realizing a fully overlapped configuration, we adiabatically apply a magnetic field gradient $B'_x$ along $\hat x$ using a pair of coils in anti-Helmholtz configuration. The gradient is controlled with a resolution at the level of $4 \, \text{mG/cm}$. This displaces the minima of the trapping potentials such that $V_{\uparrow,\downarrow}=m\omega_x^2(x\pm x_0)^2/2$ where $ x_0 = g_F \mu_B B'_x/( m \omega_x^2)$ ($g_F$ is the Land\'e factor, $\mu_B$ the Bohr magneton and $m$ the atomic mass). The SD polarizability is defined as \begin{equation} \mathcal{P}(x_0) \equiv \frac{d(x_0)}{2 x_0}, \end{equation} where $d=x_\downarrow-x_\uparrow$ is the \emph{in-situ} relative displacement between the centers of mass of each component (see Fig. \ref{Fig2}(a)). After a $2 \, \text{ms}$ SG expansion, we measure $d$ by fitting each spin component density distribution to independent Thomas--Fermi (TF) profiles to extract their centers $x_{\uparrow,\downarrow}$. The individual density profiles are not exactly TF-like, but we verified, using a Gross--Pitaevskii equation (GPE) simulation, that this approximate fitting procedure results in an overestimation of $\mathcal{P}$ by at most 6\%. Later in the text we discuss the additional correction to the measurement of $\mathcal{P}$ related to interactions during the SG expansion. Fig. \ref{Fig2}(a) shows the experimental results where the value of $x_0$ is estimated after calibrating $B'_x$. We observe that all data points strongly deviate from the prediction $d=2 x_0$ for a mixture without intercomponent interactions (green solid line), revealing the large SD polarizability of the system. \begin{figure}[t] \centerline{{\includegraphics[width=0.5\textwidth]{Fig2.pdf}}} \caption{(a) Relative displacement $d =x_\downarrow-x_\uparrow$ between the spin components as a function of $x_0$ (orange dots). The green solid line corresponds to the situation of no intercomponent interaction $d = 2 x_0$. The figure also shows a sketch of the experimental conditions (the dashed curve is the total cloud density). (b) SD oscillations for different values of $x_0$ (positive and negative) observed using the second experimental protocol. The solid lines are fit to the data according to Eq. (\ref{SGEvol}). In each figure of the paper, data error bars are the sum in quadrature of systematic and statistical errors (one standard deviation of the mean).} \label{Fig2} \end{figure} We use a second experimental protocol to determine the polarizability which will later prove to be useful for measuring the SD oscillation frequency. It consists in realizing the Rabi pulse $\ket 0 \rightarrow \ket{\uparrow,\downarrow}$ in the presence of a magnetic field gradient. As the minima of the trapping potentials for the $\ket{\uparrow}$ and $\ket{\downarrow}$ states are shifted by $\mp x_0$ with respect to the initial state $|0 \rangle$, this makes the two components oscillate out of phase after the Rabi pulse. The \emph{in-situ} time evolution of the relative displacement $D(t)=x_\downarrow(t)-x_\uparrow(t)$ is expected to be given by $D(x_0,t) = d(x_0) \left[1- \cos \left[ \omega(x_0) t \right] \right]$. Measurements of such oscillations after a SG expansion of $t_{\text{SG}} = 10 \, \text{ms}$ for different values of $x_0$ varying the magnetic field gradient are reported in Fig. \ref{Fig2}(b). After the SG expansion, the displacement between the spin components is given by $D_{\text{SG}}(x_0,t,t_{\text{SG}}) = D(x_0,t) + \partial_t D (x_0,t) \, t_{\text{SG}}$ such that we analyze the data by fitting it with the following function: \begin{equation} \label{SGEvol} D_{\text{SG}} = A(x_0,t_{\text{SG}}) \cos \left[ \omega(x_0) t + \phi(x_0,t_{\text{SG}}) \right] + d(x_0), \end{equation} where $A(x_0,t_{SG} ) = -d(x_0) \sqrt{1+\omega^2(x_0) t^2_{\text{SG}} } $ and $ \phi(x_0,t_{\text{SG}}) = \arctan \left[ \omega(x_0) t_{\text{SG}}\right]$. Eq. (\ref{SGEvol}) allows us to extract the value of $d(x_0)$ neglecting here again intercomponent interactions during the expansion. \begin{figure}[t] \centerline{{\includegraphics[width=0.5\textwidth]{Fig3.pdf}}} \caption{SD polarizability extracted from the data of Fig. \ref{Fig2}(a) (orange dots) and (b) (green triangles). The black (red) solid line is the prediction computed using the GPE (LDA). The shaded regions give the uncertainties taking into account error bars on the value of the coupling constants \cite{Knoop11}. The green solid line corresponds to the situation of no intercomponent interaction $\mathcal P = 1$. We also provide the density profiles $n_{\uparrow,\downarrow}(x,0,0)$ from the GPE for $x_0/R_x=0.001, 0.01, 0.05$. The experimental points overestimate the actual value of $\mathcal{P}$ due to the approximation of the TF fit and the interaction effect during the expansion (see text).} \label{Fig3} \end{figure} Fig. \ref{Fig3} shows the SD polarizability as a function of $x_0/R_x$ ($R_x$ is the TF radius along $\hat x$) using the data of Fig. \ref{Fig2}. We notice a strong nonlinear dependence of the polarizability on the separation between the two trapping potential minima, which is maximal in the linear limit ($x_0 \rightarrow 0$) and tends to $1$ for large separation ($x_0\gg R_x$). In the same figure, we also plot the theoretical predictions obtained within the local density approximation (LDA) and the numerical integration of the GPE performed with the experimental parameters. We identify three regions along $\hat x$, with the outer two regions occupied by either the $\ket{\uparrow}$ or $\ket{\downarrow}$ component and the inner region occupied by both of them. In the linear limit ($x_0 \to 0$) the LDA gives the result \begin{equation} \label{LDAP} \mathcal{P}(x_0 \rightarrow 0) = \frac{g+g_{\uparrow\downarrow}}{g-g_{\uparrow\downarrow}}, \end{equation} for the polarizability \cite{Sartori15}\footnote{The SD polarizability Eq. (\ref{LDAP}) should not be confused with the magnetic polarizability $\chi_M = 1/\left[n(g-g_{\uparrow\downarrow})\right]$ which is defined in uniform matter in terms of the energy cost $\delta E = M^2/(2 \chi_M)$ associated with a small polarization $M=(N_\uparrow-N_\downarrow)/(N_\uparrow+N_\downarrow)$ of the gas.}, pointing out its divergent behavior near the phase transition occurring at $g_{\uparrow\downarrow}=g$. The agreement between the LDA and the GPE is excellent except in the region of small minima separation where the LDA becomes less and less adequate because of the large value of the spin healing length $\hbar/\sqrt{2mn(g-g_{\uparrow\downarrow})}$ ($n$ is the total density of the cloud). In general, we observe a good agreement between the theoretical predictions and the experimental data. In particular, the huge effect on the polarizability caused by the vicinity to the miscible-immiscible phase transition is clearly revealed and the scaling with $x_0/R_x$ is well reproduced. The data analysis presented so far has however been performed neglecting interactions between the spin components during the SG expansion. Indeed, GPE simulations of the expansion in the presence of interactions show that the experimentally measured polarizability is overestimated by $5 \%$ ($30 \%$) for the $2 \, \text{ms}$ ($10 \, \text{ms}$) SG expansion. This explains the remaining difference between the experimental points of Fig. \ref{Fig3} and the theoretical predictions. \section{Spin-dipole oscillation} A useful estimate of the SD frequency is obtained by employing a sum rule approach \cite{Pitaevskii16} based on the ratio $\hbar^2 \omega_{\text{SD}}^2 = M_1/M_{-1}$, where $M_1=N \hbar^2/2m$ ($N=N_\uparrow+N_\downarrow$) is the model independent energy weighted sum rule relative to the SD operator $\sum_{i} (x_{i\downarrow} - x_{i\uparrow})$, and $M_{-1}=N\mathcal{P}(x_0 \rightarrow 0)/(2m\omega_x^2)$ is the inverse energy weighted sum rule fixed according to linear response theory by the linear SD polarizability \cite{Sartori15}. This leads to the following relation between the SD frequency and polarizability \begin{equation} \label{relationPolSD} \omega_{\text{SD}}= \frac{\omega_x}{\sqrt{\mathcal{P}(x_0 \rightarrow 0)}} \, . \end{equation} Using the LDA expression (\ref{LDAP}) for the polarizability one derives the following prediction for the SD frequency \begin{equation} \label{SD_ratio} \omega_{\text{SD}}= \sqrt{ \frac{g-g_{\uparrow\downarrow}}{g+g_{\uparrow\downarrow}}} \, \omega_x \, . \end{equation} The same result can be directly obtained by generalizing the hydrodynamic theory developed in \cite{Stringari96} for density oscillations to the case of SD oscillations \cite{Pitaevskii16}. Eq. (\ref{SD_ratio}) explicitly points out the crucial role played by the intercomponent coupling constant $g_{\uparrow\downarrow}$ in softening the frequency of the SD mode with respect to the value $\omega_x$ characterizing the frequency of the in-phase center-of-mass oscillation. We check, using time-dependent GPE simulations of the SD oscillations for our experimental parameters, that the sum rule prediction (4) provides $\omega_{SD}$ with an accuracy better than $1\%$ when substituting the value of the static SD polarizability $\mathcal P(x_0\rightarrow0)$ from the GPE. This demonstrates that an accurate SD frequency measurement can be used to determine the value of the SD polarizability. As shown on Fig. \ref{Fig2}(b), the Rabi pulse in the presence of a magnetic field gradient gives rise to the excitation of SD oscillations whose frequency can be extracted as a function of the induced displacement $x_0$. A first estimate of the SD frequency is obtained considering that $\omega_{\text{SD}} = \omega(x_0 \rightarrow 0)$. Indeed, for large values of $x_0$, the oscillation frequency $\omega(x_0)$ approaches $\omega_x$ while it decreases to $\omega_{\text{SD}}$ as $x_0 \rightarrow 0$. Since in the small $x_0$ limit the amplitude of the oscillation tends to zero, we perform a linear fit to the curve of $\omega(x_0)/\omega_x$ as a function of the oscillation amplitude $A(x_0)$ and extract $\omega_{\text{SD}}/\omega_x=0.18(1)$ from the y-intercept of the linear fit (see Fig. \ref{Fig4}(a)). This method shows a good agreement with the LDA prediction Eq. (\ref{SD_ratio}) $\omega_{\text{SD}}/\omega_x = 0.189(15)$ and with the GPE simulations yielding $\omega_{\text{SD}}/\omega_x = 0.213(17)$ (uncertainties take into account error bars on the value of the coupling constants \cite{Knoop11}). The different values of $\omega_{\text{SD}}$ from the LDA and GPE calculations have the same origin as the one discussed in the case of the polarizability and are due to the large value of the spin healing length in the vicinity of the quantum phase transition. \begin{figure}[t] \centerline{{\includegraphics[width=0.5\textwidth]{Fig4.pdf}}} \caption{(a) Ratio $\omega(x_0)/\omega_x$ as a function of the amplitude $A(x_0)$ for the data of Fig. \ref{Fig2}(b) (same marker styles). The black (red) marker is the prediction of the GPE $\omega_{\text{SD}}/\omega_x = 0.213(17)$ (LDA $\omega_{\text{SD}}/\omega_x = 0.189(15)$). Extrapolating the linear fit of $\omega(x_0)/\omega_x$ (solid blue line) for vanishing amplitude gives $\omega_{\text{SD}} /\omega_x = 0.18(1)$. (b) SD oscillations using the alternative method (blue dots) giving $\omega_{\text{SD}}/\omega_x = 0.218(2)$. We also show two density profiles $n_{\uparrow,\downarrow}(x,0,0)$ illustrating the out-of-phase SD oscillations (data obtained from the GPE taking the equilibrium state for $x_0/R_x=0.01$ as initial condition before setting $x_0=0$ to start the oscillations).} \label{Fig4} \end{figure} An alternative and more efficient way to excite the SD mode and to measure its frequency consists in first creating two perfectly overlapped spin states where $B'_x = 0$ ($x_0=0$) and then applying a magnetic field gradient $B'_x = 0.1 \, \text{G/cm}$ ($x_0 = 1.3 \, \mu \text{m}$) for $3 \, \text{ms} \ll 2 \pi / \omega(x_0)$ before restoring $B'_x = 0$. This leads to an \emph{in-situ} dipole oscillation shown in Fig. \ref{Fig4}(b) after $10 \, \text{ms}$ of SG expansion. We measure $\omega_{\text{SD}}/\omega_x = 0.218(2)$ which is slightly larger than the previous estimate based on the data of Fig. \ref{Fig2}(b) and shows better agreement with the prediction from the GPE simulations. For a precise determination of $\omega_{\text{SD}}$, it is important to ensure that the SD mode has a small \emph{in-situ} amplitude: here we estimate $D_{\text{SD}} = D_{\text{SG}} / \sqrt{1+\omega^2_{\text{SD}} t^2_{\text{SG}}} = 5.4 \, \mu\text{m}$ which is relatively small compared to the TF radius $R_x=40 \, \mu\text{m}$. In all experiments, we observe oscillations without noticeable damping on very long timescales (they are ultimately limited by the cloud lifetime). Indeed, the maximal relative velocities of the two superfluid components $v_{\text{max}} = 1.2 \, \text{mm/s}$ for the data of Fig. 2(b), and $v_{\text{max}} = 0.4 \, \text{mm/s}$ for the data of Fig. 4(b) are smaller than the critical velocity for the dynamical counterflow instability $v_{\text{cr}} = \sqrt{\mu_{\text{tot}}(1- g_{\uparrow\downarrow}/g)/2m} = 1.8 \, \text{mm/s}$ \cite{Abad15}. \section{Conclusion} In conclusion, we reported on the experimental measurements of the polarizability and of the frequency of the SD oscillation in a two-component BEC of sodium. Because of the vicinity to the miscible-immiscible quantum phase transition both quantities are very sensitive to the value of the intercomponent interaction and their behavior deviates by large factors from the values predicted in the absence of intercomponent interaction. This represents a major difference with respect to other available superfluid quantum mixtures, like the Bose-Fermi mixtures of lithium gases \cite{Ferrier14,Delehaye15}, where the role played by the intercomponent interaction is much less crucial. Similarly to the case of \cite{Ferrier14,Delehaye15} our mixture is characterized by two interacting superfluids oscillating with opposite phase and the observed SD oscillation is undamped for small amplitude as a consequence of superfluidity. For large amplitude motion the Landau's critical velocity will, however, behave very differently, being very sensitive to the value of the intercomponent interaction \cite{Abad15}. Another interesting feature concerns the behavior of the SD oscillation at finite temperature. While the damping of the SD oscillation was actually observed in the old experiments of \cite{DeMarco02} carried out on a normal Fermi gas, understanding the behavior of the collective modes in the presence of both a condensed (superfluid) and thermal (non-superfluid) components remains extremely challenging \cite{Armaitis15, Lee16}. Other topics of interest concern the experimental realization of magnetic solitons \cite{Qu16} and the inclusion of coherent coupling between the two spin components. The Bose mixtures realized and investigated here then represent an ideal platform to explore important equilibrium and dynamic properties of binary superfluids. \begin{acknowledgments} We thank C. Salomon and G. Roati for discussions and critical reading of the manuscript, and L. Festa for technical assistance at the early stage of the experiment. We acknowledge funding by the Provincia Autonoma di Trento, the QUIC grant of the Horizon 2020 FET program, and by the Istituto Nazionale di Fisica Nucleare. \end{acknowledgments}
fe89f6242b0f09e3e015ed4eb04b014a8704871b
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In this paper, we introduce a novel objective for the generation of all $k$-subsets of $n$ elements and we discuss the structure of the resulting combinatorial optimization task. In general, improving the order of elements in a sequence towards some objective is recognized to be a complex optimization task \cite{dewar} with interesting applications in computer science such as unit test coverage \cite{bryce,rothermel}. Star trackers (see Figure \ref{fig:hydra}) are a common sensor used by a spacecraft to determine its attitude by looking at fixed stars. The problem we here formalize and tackle was suggested by the work of Mortari et al. \cite{MortariEtAl2004} on the design of efficient algorithms for star identification. In that paper, the authors consider a ``lost-in-space" spacecraft attitude identification problem: find the orientation of a spacecraft in deep space using a single star tracker image. Such a problem corresponds to that of identifying $k$ stars in a ``scene" (i.e. a picture taken by the star tracker) containing $n$ spikes of which $t$ are unknown stars and the rest are artifacts due to various disturbances present in harsh space environments. Part of the algorithm proposed in that paper, called the Pyramid Algorithm and today widely used in many star-trackers in orbit, needs to generate all $\binom{n}{3}$ combinations in a smart order that allows the discovery of three true stars in the scene from a minimal number of star catalog queries. \section{Background} A star tracker utilizes a star catalog containing the positions of all known stars having brightness larger than some predefined threshold. A camera is used as sensory device to capture the light coming from the stars which will then create spikes in the image. The crucial software part of a star tracker is the star identification algorithm whose task is to match spikes extracted from the image with stars in the star catalog. With the matched stars it is then possible to calculate the orientation of the spacecraft. One major problem for star identification algorithms is that not all spikes in an image are associated to real stars. Spikes can be caused by reflections on debris, radiation or other spurious sources. A correct and fast star identification is critical in space: a failure of the spacecraft to detect its own attitude promptly may lead to the complete failure of the mission. \begin{figure}[tbp] \centering \includegraphics[width=0.8\columnwidth]{hydra_star_tracker} \caption{The flight model of the hydra star-tracker currently flying on the ESA Sentinel 3A satellite mission.} \label{fig:hydra} \end{figure} \subsection{The Pyramid Algorithm} A widely used star identification algorithm \cite{SpratlingEtAl2009} is the Pyramid Algorithm developed by Mortari et al. \cite{MortariEtAl2004}. The algorithm is based on the distances between two stars in an image. A database stores the distance between each star pair from the catalog up to the maximum distance that is possible within the field of view of the camera. The look-up of an arbitrary distance within the measurement tolerance typically results in a number of possible star pairs from the catalog. It is therefore necessary to build higher order graphs with the spikes in the image, such as a triangle consisting of three distances. The resulting star pairs from the distance look-ups in the database have to be matched to form an actual triangle given the IDs of the stars in the star catalog. \begin{figure}[tbp] \centering \includegraphics[width=0.8\columnwidth]{scene2} \caption{A random scene as a star tracker would see it. Ten ($n=10$) detected spikes in the image are numbered arbitrarily from 0 to 9 and marked as circles. Artifacts are colored in red, actual stars ($t=5$) are colored in green. The pyramid algorithm detected the four stars connected with blue lines as a valid pyramid.} \label{fig:scene1} \end{figure} An important part of the algorithm is to select spike triplets resulting in a sequence of queries to the database able to find a positive match as quickly as possible. This problem equals to that of wisely selecting the order of all possible $k$-subsets of $n$ given spikes, where, in this case, $k=3$. For example, as discussed also in the original paper by Mortari, the lexicographic order (which tries the spikes in the sequence 0-1-2, 0-1-3, 0-1-4, etc.) is clearly problematic, as the first spike is repeated until all possible combinations containing it have been also tried. In the case of such a spike being spurious this approach results in many unnecessary database queries. As part of the Pyramid Algorithm, Mortari et al. proposed an algorithm to generate the subsets in a smarter fashion, while acknowledging that the mathematically optimal solution is of interest, but unknown. This stateless algorithm, which we call \emph{pattern shifting} algorithm, is described in the original paper and here reported for convenience: \begin{verbatim} for dy from 1 to n-2 for dz from 1 to n-dy-1 for x from 0 to n-dy-dz-1 y = x + dy z = y + dz next combination is [x, y, z] \end{verbatim} This code would produce the sequence 0-1-2, 1-2-3, 2-3-4, etc. \section{Problem Description} Consider a star tracker camera retrieving an image containing $n$ spikes arbitrarily numbered from 0 to $n-1$ of which an unknown subset $s = \{p_1, p_2, \dots, p_t\}$ of cardinality $t$ are actual stars and the remaining spikes artifacts. A fixed $n$, $s$ defines what will here be called a scene, which can be thought of as the underlying, unknown, ground truth. A query $q = \{c_1, c_2, \dots, c_k\}$ is a $k$-subset of the spikes and is said to discover the scene if $q \subseteq s$. We study the problem of generating, for a fixed $n$, an ordered sequence $Q=\{q_j, j\in 1..M\}$ of $M$ queries able to discover the unknown scene, on average, at the smallest trial $j$. To guarantee that $j$ always exists, we only consider a sequence $Q$ if it contains all $M = N = \binom{n}{k}$ possible queries (in which case we refer to $Q$ as \emph{complete}) and we assume $t\ge k$ as scenes with less than $t$ spikes cannot be discovered. The set $S = \{s_1, s_2, \dots \}$ for a fixed $n$ contains all possible scenes that are discoverable ($t \ge k$). Due to lack of prior knowledge about $S$, we assume all scenes to have equal probability $|S|^{-1}$, where: \begin{equation} |S| = \sum_{t=k}^{n} \binom{n}{t} \end{equation} We formalize the problem as: \begin{equation}\label{eq:ttd} \begin{array}{rl} \mbox{find:} & Q \\ \mbox{to minimize:} & T(Q) = \frac{1}{|S|}\sum_{s \in S} \tau\left(Q, s\right) \end{array} \end{equation} where $T(Q)$ is the expected time to discovery, that is the average across all possible scenes $s \in S$, of $\tau(Q, s)$: the index $j$ of the first $q_j \in Q$ discovering $s$. \subsection{Discovery of Scenes} Let $D(q_i)$ be the number of scenes discovered by the query $q_i$ but not by any of the previous queries $q_{j< i}$. Clearly, for each of such scenes, $\tau(Q, s)=i$. We may then rearrange the terms in the expression for the expected time to discovery in Eq.(\ref{eq:ttd}) grouping together all scenes discovered by $q_i$ and thus summing over all $q_i \in Q$: \begin{equation}\label{eq:TQq} T(Q) = \frac{1}{|S|} \sum_{i=1}^{|Q|} i D(q_i). \end{equation} In this form, the objective function suggests that the number of scenes discovered at each $i$ should be larger than those discovered at any $j>i$, a simple thought that will form the basis of some of the most successfull algorithms here introduced. \subsection{Example} Consider a simple scenario with $n=5$ spikes and $k=3$. We assume the spikes are numbered from 0 to 4. Let us compute the time to discovery $T$ of the sequence $Q$ generated by the pattern shifting algorithm by Mortari et al.: \begin{equation*} Q = \begin{Bmatrix} \{ 0, 1, 2 \}, \\ \{ 1, 2, 3 \}, \\ \{ 2, 3, 4 \}, \\ \{ 0, 1, 3 \}, \\ \{ 1, 2, 4 \}, \\ \{ 0, 1, 4 \}, \\ \{ 0, 2, 3 \}, \\ \{ 1, 3, 4 \}, \\ \{ 0, 2, 4 \}, \\ \{ 0, 3, 4 \} \\ \end{Bmatrix} \end{equation*} As a start, consider the scene $s_e=\{1, 2, 3, 4\}$. As $q_1 = \{0,1,2\}\not \subseteq s_e$ does not discover the scene, but $q_2 =\{1,2,3\}\subseteq s_e$ does, the number of queries necessary to discover this particular scene is $\tau\left(Q, s_e\right) = 2$. To compute the total time to discovery $T(Q)$, we must sum over all possible scenes (of which there are 16 with $t \ge 3$ : one with $t=5$, five with $t=4$ and ten with $t=3$). Listing all possibilities it is not difficult to find that we have four scenes discovered by $q_1$, two discovered for each $q_2$ to $q_4$ and one for each $q_5$ to $q_{10}$. This, according to Eq.(\ref{eq:TQq}) gives a final score of: $$ T(Q) = \frac{1}{16} \left( 4\cdot1 + \sum_{i=2}^{4} 2 \cdot i + \sum_{i=5}^{10} 1\cdot i \right) = \frac{67}{16} $$ The optimum in this case could be reached, for example, by swapping $q_2$ and $q_{10}$ leading to a score of $T(Q) = \frac{65}{16}$. The lexicographic sequence, for reference, has, in this case, a score of $T(Q) = \frac{71}{16}$. \section{Analysis}\label{sec:analysis} We discuss some properties useful to design algorithms aimed at solving the problem stated in Eq.(\ref{eq:ttd}). \subsection{Equivalent sequences}\label{sec:uniqueness} No sequence $Q$ has a unique score. Consider the equivalence class $[Q]$ containing all sequences having the same time to discovery. The cardinality of this set is at least $n!$. Since this property is also valid for the optimal sequence, we conclude that there are at least $n!$ solutions. This result follows immediately noting that any permutation of the $n$ spike IDs does not change the score. Therefore, applying the same permutation to all elements in each $q_i \in Q$ will result in an equally scored $Q'$. \subsection{Scene Discovery Monotonicity}\label{sec:monotonicity} If $Q$ is optimal, then the values $D(q_i)$ monotonically decrease with $i$. This property, called the monotonicity property, states that the number of scenes discovered by each successive query in an optimal sequence is a monotonically decreasing sequence. It is illustrated in Table \ref{table:monotonicity}. The sequence on the left cannot be optimal as it violates the monotonicity property. By swapping the highlighted queries we obtain a better sequence. \begin{table} \centering \caption{Monotonicity of the Optimal Order} \label{table:monotonicity} \def\arraystretch{1.1 \begin{tabular}{cc|cc} $q_i$ & $D(q_i)$ & $q^*_i$ & $D(q^*_i)$ \\ \hline \{0, 1, 2\} & 4 & \{0, 1, 2\} & 4 \\ \{0, 3, 4\} & 3 & \{0, 3, 4\} & 3 \\ \textbf{\{0, 1, 3\}} & \textbf{1} & \textbf{\{2, 3, 4\}} & \textbf{2} \\ \textbf{\{2, 3, 4\}} & \textbf{2} & \textbf{\{0, 1, 3\}} & \textbf{1} \\ \{1, 2, 4\} & 1 & \{1, 2, 4\} & 1 \\ \{0, 1, 4\} & 1 & \{0, 1, 4\} & 1 \\ \{0, 2, 3\} & 1 & \{0, 2, 3\} & 1 \\ \{1, 3, 4\} & 1 & \{1, 3, 4\} & 1 \\ \{0, 2, 4\} & 1 & \{0, 2, 4\} & 1 \\ \{1, 2, 3\} & 1 & \{1, 2, 3\} & 1 \\ \cline{2-2} \cline{4-4} & $T(Q)$ = 64 & & $T(Q^*)$ = 63 \end{tabular} \end{table} \begin{proof} Consider two consecutive queries $q_i, q_{i+1} \in Q$ such that $D(q_{i}) < D(q_{i+1})$, we show that swapping the subsets $q_i, q_{i+1}$ results in a new sequence $Q^*$, where $q^*_{i} = q_{i+1}$ and $q^*_{i+1} = q_{i}$, with a lower average time to discovery implying that, for the optimal sequence, $D(q_i)$ monotonically decreases with $i$. Due to the swap, $D(q_{i}) \ge D(q^*_{i+1})$ because of scenes that can be discovered by both $q_i$ and $q_{i+1}$. According to Eq. \eqref{eq:TQq}, The contribution of the queries $q_i$, $q_{i+1}$ to $T(Q)$ is: $$T_i = i D(q_i) + (i+1) D(q_{i+1})$$ $T^*_i$ can be similarly defined as the contribution of the queries $q^*_i$, $q^*_{i+1}$ to $T(Q^*)$. Using $D(q_{i}) + D(q_{i+1}) = D(q^*_i) + D(q^*_{i+1})$, as both queries together always discover the same amount of scenes, we get $$T_i - T^*_i = D(q_{i+1}) - D(q^*_{i+1}) > 0,$$ as $D(q_{i+1}) > D(q_{i}) \ge D(q^*_{i+1})$. Swapping the two queries does not alter the contributions of the previous and following queries. Therefore, $T_i > T^*_i$ results in $T(Q^*) < T(Q)$ \end{proof} \subsection{The Average $T$} Consider the set $\mathcal Q$ of all sequences that are \emph{complete} and do not contain repetitions. Following Eq.(\ref{eq:ttd}) and Eq.(\ref{eq:TQq}) we may restate the fundamental problem subject of this paper as: \begin{equation} \begin{array}{rl} \mbox{find:} & Q \in \mathcal Q\\ \mbox{to minimize:} & T(Q) = \frac{1}{|S|} \sum_{i=1}^{|Q|} i D(q_i). \end{array} \end{equation} The dimension of our search space, that is the cardinality of $\mathcal Q$, can be computed noting that the number of queries $q_i \in Q$ is $N = \binom{n}{k}$. The total number of different sequences in $\mathcal Q$, the search space dimension, is then $N!$. It is of interest to compute $\sigma$, the average across the whole search space of $T(Q)$. Any algorithm producing a sequence scoring less than $\sigma$ will be considered as making good use of the problem structure, while the opposite can be said for algorithms that generate sequences scoring more than $\sigma$. The value $\sigma$ can be computed as follows: \begin{equation*} \sigma = \frac{1}{N!} \sum_Q T(Q) \end{equation*} Computing $\sigma$ directly from this definition is quite expensive, thus we derive a simpler formula that has a much lower complexity. Let us start by assuming the first $i-1$ queries in some $Q$ have not discovered a scene $s_t$ having $t$ true stars. The probability that the following query $q_i \in Q$ will discover it is denoted with $p(q_i, t)$. This probability is the fraction between the number of queries able to discover $s_t$ and the total number of remaining, possible queries: \begin{equation*} p\left(q_i, t\right) = \frac{\binom{t}{k}}{N - i + 1}, \end{equation*} This probability increases with $i$ and is well defined only when $i \le N+1-\binom{t}{k}$. We may then compute the unconditional probability $p_i(t)$ that $s_t$ will be detected at the $i$-th query $q_i$ as: \begin{equation*} p_i(t) = p\left(q_i, t\right) \prod_{j=1}^{i-1} \left(1 - p\left(q_j, t\right)\right) \end{equation*} Using this equation, it can be shown that $\sigma$ can be derived as: \begin{equation}\label{eq:sigma} \sigma = \frac{1}{|S|} \sum_{t=k}^{n} \left( \left|S_t\right| \sum_{i=1}^{N} i p_i(t) \right), \end{equation} where $\left|S_t\right| = \binom{n}{t}$ is the number of possible scenes with exactly $t$ true stars. The complexity of computing this is $O(N n)$. Some examples are shown in Table~\ref{tab:ET}, showing that the expectation for a fixed $k$ has a maximum at some $n$. \begin{table} \centering \caption{Values of $\sigma$ for Different $n$ and $k$.} \label{tab:ET} \begin{tabular}{c|c|r} $n$ & $k$ & \multicolumn{1}{c}{$\sigma$} \\ \hline 5 & 3 & 4.2 \\ 10 & 3 & 17.4 \\ 20 & 3 & 16.0 \\ 50 & 3 & 9.9 \\ 100 & 3 & 8.8 \\ \end{tabular} \hspace{1em} \begin{tabular}{c|c|r} $n$ & $k$ & \multicolumn{1}{c}{$\sigma$} \\ \hline 20 & 1 & 2.0 \\ 20 & 2 & 5.1 \\ 20 & 5 & 322.5 \\ 20 & 10 & 32528.1 \\ 20 & 15 & 5748.7 \\ \end{tabular} \end{table} \section{Algorithms} We propose several algorithms to generate a solution to the problem \eqref{eq:ttd} that we group in three families. First we detail algorithms able to find an optimal solution. Since these reveal to be computationally intractable with growing $n$, we introduce a second family including algorithms making use of different heuristics to search in $\mathcal Q$ resulting in a great reduction of the computational cost but without the guarantee of finding the optimal sequence. Finally, the algorithms of the last family are stateless $k$-subset (queries) generators. They have a considerably lower computational cost, making them attractive for real-time generation in star trackers. This family includes a generalization of the pattern shifting algorithm described by Mortari et al. \cite{MortariEtAl2004}. \subsection{Finding an optimal solution} We start with the naive brute-force approach computing the score for all $Q \in \mathcal Q$ and returning the minimum. The same result, implemented as a branch and prune algorithm making use of the properties introduced in section \ref{sec:analysis} to prune large portion of the search space not containing the optimal solution. \subsubsection{Brute-force} In the brute-force approach one simply determine the time to discovery for all possible sequences of queries. This way of determining the best sequence comes at a high cost, with a time complexity of $O\left(N! k |S|\right)$. Every permutation of the $N$ queries has to be checked against every possible scene. It is easy to understand that this complexity grows out of any reasonable bounds even for low numbers of $n$ and $k$. \subsubsection{Branch and prune} The brute-force algorithm can be implemented as branch and prune tree search. Each level in the tree adds another query to the sequence. This means the root node has $N$ neighbors and the branching factor at each level is decreasing by one from there, resulting in a depth of $N$ for all the leaves in the tree. To optimize computational performance, scenes and queries can be stored and processed in a binary representation. The bit-wise \emph{AND} operation allows to check if a query would discover a scene. The search for the optimum can then easily be done using any complete tree search algorithm and evaluating the score at each leaf of the tree to find the minimum. The first optimization to the brute-force algorithm is to compute the score during the exploration of the nodes in the tree, while doing a depth first search. If a node in the tree has a higher score than the minimum so far, the whole branch can be pruned. Further pruning is possible by using the monotonicity property (section~\ref{sec:monotonicity}). If the number of scenes discovered by a query is higher than in the previous step, the whole branch can also be pruned. The property also allows to determine a lower bound on the score of a branch, which allows pruning earlier when compared to the current minimum score. Another optimization considers the possibility to permute the elements making up the queries as described in section~\ref{sec:uniqueness}. When the search tree is pruned by only allowing one of these permutations, the number of sequences to be checked is a factor $n!$ lower. Additionally, at some point, all the remaining queries remove exactly one scene each. All possible permutations of these queries will lead to the same score. When looking for only one optimum the search can just take any random permutation of the remaining queries at this point or directly compute all permutations without recalculating the score for each of them. Although the required time to find the optimal sequence drastically decreases, these optimizations do not influence the overall computational complexity, making it infeasible to compute the optimal sequence for $n>6$, $k>3$ on computers that were available to us at the time of writing. \subsection{Goal-driven Algorithms} In this section we propose two algorithms that allow us to compute sub-optimal solutions for higher values of ($n$, $k$). Both algorithms use heuristics to assign a score $\delta(q_i)$ to a candidate query $q_i$ that indirectly estimates how well it contributes to minimize the sequence's overall time to discovery. Then, the query with the highest $\delta(q_i)$ is executed. \subsubsection{Greedy Scene Elimination (GSE)} \label{sec:GSE} In this greedy algorithm, a sequence is built by simply selecting the queries $q_i$ that discovers the highest number of scenes $\delta_{\text{GSE}}(q_i) = D(q_i)$ at each step. This approach implicitly produces sequences having the monotonicity property described in section~\ref{sec:monotonicity}. Ties can be resolved in different ways. Our current implementation chooses the query that comes first in lexicographic order. We found that the choice influences the final score, which is evidence for the non-optimality of the algorithm. Initially, all the scenes are generated and stored. Then, with each query, the discovered scenes are eliminated from the list of scenes, hence the name of the algorithm. This equals following just one path to a single leaf in the tree search algorithm. The computational complexity of this algorithm is therefore lowered to $O\left(Nk |S|\right)$, while the memory consumption is $O\left(n |S|\right)$ to store the scenes that need to be checked. While the branch and prune algorithm only allowed us to compute the optimal solution for $n < 7$, the greedy scene elimination algorithm allowed us to compute solutions for $n < 32$ within a few days. \subsubsection{Minimally Intersecting Subsets (MIS)} Given the computational requirements of GSE, a simpler scoring function is defined by considering the overlap between the candidate $q_i$ and each one of queries previously executed. For each $q_i$, we select the one with the highest: $$\delta_{\text{MIS}}(q_i) = - \sum_{j=1}^{i-1}2^{|q_i \cap q_j|} - 1$$ The term $2^{|q_i \cap q_j|}$ corresponds to the number of subsets in the intersection $q_i \cap q_j$. We exclude the empty subset from the count to avoid undesirable results: two queries with $|q_i \cap q_j|=0$ would be equivalent to one with $|q_i \cap q_j|=1$. The same procedure as in GSE is used to break ties. The score $\delta_{\text{MIS}}(q_i)$ can be seen as an approximation to $\delta_{\text{GSE}}(q_i)$. Let $d(q_i)$ be the set of scenes $\subseteq S$ that could be discovered by $q_i$, we can express $\delta_{\text{GSE}}(q_i)$ as: \begin{IEEEeqnarray*}{ll} \hspace{2em}&\hspace{-2em} \delta_{\text{GSE}}(q_i) = D(q_i) = | d(q_i) | \\ & - \sum_{j=1}^{i-1} | d(q_i) \cap d(q_j) | \\ & + \sum_{j=1}^{i-1}\sum_{k=1}^{j-1} | d(q_i) \cap d(q_j) \cap d(q_k) | \\ & - \sum_{j=1}^{i-1}\sum_{k=1}^{j-1} \sum_{l=1}^{k-1} | d(q_i) \cap d(q_j) \cap d(q_k)\cap d(q_l) | \\ & + \ldots \IEEEyesnumber \label{eq:sc_removed} \\ \end{IEEEeqnarray*} where the first term is equal for all $q$ and thus removing it results in an equivalent score function. Here we consider an approximation to $\delta_{\text{GSE}}$ by only considering the second term: $$\delta_{\text{MIS}^*}(q_i) = - \sum_{j=1}^{i-1} | d(q_i) \cap d(q_j) | $$ It can be shown that the size of the intersection between the scenes removed by $q_i, q_j$ can be expressed more easily as: $$\delta_{\text{MIS}^*}(q_i) = -\sum_{j=1}^{i-1} 2 ^{n - 2k + |q_i \cap q_j|}$$ This corresponds to $\delta_{\text{MIS}}(q_i)$, except for a factor $2^{n - 2k}$ and the exclusion of the empty subset in the count as explained above. Increasingly better approximations to GSE could be obtained by considering more and more terms of $\eqref{eq:sc_removed}$, an option we have not studied in this paper. \subsection{Stateless Sequence Generation Algorithms} The algorithms described in this section do not explicitly consider the objective in equation~\eqref{eq:ttd} to generate $k$-subsets. Thus they do not require to store the previously executed queries or the remaining stars, not having a representation of the current state of the problem. This leads to implementations with lower computational and memory complexity. These algorithms can be seen as analogous to those used to solve the minimal change ordering problem such as the revolving door algorithm~\cite{cages}, the Eades-McKay algorithm~\cite{matters}, and Chase's sequence~\cite{knuth}. Trying to minimize the distance between consecutive $k$-subsets, they perform badly in our case, as close, similar subsets discover fewer scenes. \subsubsection{Generalized Pattern Shifting Algorithm} All the previous methods require the evaluation of a score function for every query, which requires a high computational cost even for simple functions as $n$ and $k$ increase. Those algorithms cannot run in real-time on a star tracker. In \cite{MortariEtAl2004}, Mortari et al. propose an algorithm for producing $k=3$ subset sequences. The algorithm is notable for the low computational complexity with which it generates sequences, while maintaining some degree of diversity among its sequentially produced subsets. We present here a generalization of the pattern shifting algorithm to arbitrary $k$, and arbitrary reference sequences. Let $E = \{e_0, \ldots, e_{n-1}\}$ represent the $n$ elements being grouped into $k$-element subsets. A reference sequence generator is made to iteratively produce its sequence of $(k-1)$-element subsets from $E \setminus \{e_0\}$. Each subset $q'$ produced this way is extended with $e_0$, so as to compose a valid $k$-element subset: $q_i = \{e_0\} \cup q'$. Multiple variants of this subset are then produced, by incrementing/shifting subset elements: $q_{i+1} = \{e_{m+1} | e_m \in q_i\}$. This will lead to a sub-sequence $\{q_i, \ldots, q_j\}$ of $Q$, ending at the subset $q_j$ for which $e_{n-1} \in q_j$. Should this condition hold already for $q_i$, the sub-sequence will then include a single subset. In the following sections we take lexicographic as the reference sequence generator, which, for $k=3$, then perfectly replicates the sequences produced by the algorithm given in \cite{MortariEtAl2004}. However, the formulation presented above allows for other reference generators to be considered. It can even use itself recursively stopping at $k=0$ which simply results in a single empty subset. Naturally, the quality of the sequences generated by this algorithm are closely tied to the quality of the sequences produced by the reference generator. Exchanging lexicographic for revolving door \cite{cages}, for instance, will improve the overall quality. \subsubsection{Unranking Algorithms} We here consider a class of sequence generators that rely on unranking functions to produce their sequences. These are functions that, for a given rank $r$, will generate the $(r+1)$-th subset in some reference sequence. In \cite{cages}, unranking functions are provided for the lexicographic, co-lexicographic and revolving door sequence generators. By relying on an unranking function, the problem of generating a sequence is recast as the problem of how to best generate a sequence of $N$ ranks in $\{0, \ldots, N-1\}$. A pseudorandom number generator (PRNG) can be used to produce these ranks. If the uniqueness of ranks is enforced, for instance via a full cycle PRNGs such as linear feedback shift registers, we then obtain a pseudorandom sequence without repetitions, effectively a pseudorandom permutation of the $N$ subsets. In terms of scalability, unranking functions enable such sequences to be generated without the need to obtain ahead the full sequence $Q$ and then perform a shuffling operation in memory over it. This approach can however be improved through analysis of the reference sequences for which the unranking functions are designed. Lexicographic, co-lexicographic and revolving door are all sequences where similar subsets can be found clustered with similar ranks. A logical approach is then to seek a process that generates ranks that are locally as dissimilar as possible, and preserves this property throughout the full sequence. Techniques to achieve that can be found in the domain of low-discrepancy, or quasi-random sequence generators. We introduce here an approach, under the name ``base unrank", that is based on the van der Corput sequence \cite{corput}. Let $L = \lceil log_b{N} \rceil$ represent the number of digits required in order to count from 0 to $N-1$ in base $b$. The sequence of ranks is produced by counting from $0$ to $b^L - 1$, in base $b$, while prioritizing increments to the \emph{most significant digit}. This is equivalent to a sequence that increments ranks as per the normal rules of addition, but at each step then reverses the $L$ digits. Ranks $r \geq N$ produced by this process are simply skipped in the sequence. This sequence generation equals using the van der Corput sequence from $0$ to $b^L - 1$ multiplied by $b^L$ as index for the unranking function. Consider the example of base unranking for $n=5$, $k=3$ ($N=10$). Taking $b=2$ will result in a counting over $L=4$ digits. In binary, the sequence would be $\{0000, 1000, 0100, \ldots, 1111\}$, which would decode in base 10 to $\{0, 8, 4, 12, 2, 10, 6, 14, 1, 9, 5, 13, 3, 11, 7, 15\}$. Skipping the ranks $r \geq N$ would then lead to the sequence $\{0, 8, 4, 2, 6, 1, 9, 5, 3, 7\}$. These ranks, produced sequentially, and mapped through the unranking function for some reference sequence such as the revolving door (the default used in the remainder of this paper), will generate a valid $k$-elements subset sequence with no repetitions, and little overhead (computationally, the costlier operation will be the unranking itself). \iftrue \subsubsection{Random Queries} For completeness, we also consider the expected score of purely random queries, selecting $k$-subsets at uniform probability from the set of all possible $k$-subsets. This means, that $Q$ would contain repetitions and has infinite length. In practice it would be necessary to abort the discovery at some point to avoid running an endless loop, when a scene cannot be discovered. With a derivation similar to that of Eq.(\ref{eq:sigma}), we obtain: \begin{equation} \expect{T(Q)} = \frac{1}{|S|} \sum_{t=k}^{n} \left( |S_t| \cdot \frac{N}{\binom{t}{k}} \right) \end{equation} \fi \section{Results} \begin{figure}[t] \centering \subfloat[][Optimal]{ \includegraphics[width=.1\textwidth]{sequences/optimal_6_3}\vspace{1em} } \subfloat[][GSE]{ \includegraphics[width=.1\textwidth]{sequences/greedy_scene_elimination_6_3}\vspace{1em} } \subfloat[][Scenes Discovered]{ \includegraphics[width=.2\textwidth]{optimum_scenes} } \caption{For $n=6$ and $k=3$ an optimal solution (a) cannot be found by the greedy scene elimination (b) algorithm. This is because the second query of the optimal solution discovers less scenes than GSE at this step, but is then able to discover more scenes with the subsequent two queries.} \label{fig:sequences_6_3} \end{figure} \begin{figure*}[t] \centering \includegraphics[width=\textwidth]{triple_plot} \caption{Comparison of the performance of the algorithms presented. The left and center panel show the score $T(Q)$ of the sequences $Q$ generated by the algorithms depending on the number of spikes $n$ (with $k=3$, left panel) and the subset length $k$ (with $n=10$, center panel). The right panel shows the number of scenes left to discover for the sequences $Q$ after every query $q_i$ for $i$ from 0 to $N-1$ for the scenario with $n=20$ and $k=3$.} \label{fig:results} \end{figure*} We test all the proposed algorithms in scenarios with different values of ($n$, $k$), reporting the average time to discovery $T(Q)$ of each case. We pay particular attention to the $k=3$ cases, due to their relevance to star trackers. Additionally, we analyse the sequences produced in some of these cases to study how the different algorithms behave. The branch and prune algorithm can find an optimum sequence for scenarios up to $n = 6$. In the case of $n=6, k=3$, it is interesting to note that there is an overlap between the first two queries, $[0, 1, 2]$ followed by $[0, 3, 4]$ as illustrated in Figure~\ref{fig:sequences_6_3}. It is interesting to note that such sequence is not generated by GSE and MIS algorithms, given that they try to select the query discovering the maximum amount of scenes at each step, thus minimizing the overlap with respect to the previous sequence. As a reference, we include the sequences generated by the algorithms for $n=10, k=3$ in Figure \ref{fig:sequences_10_3}. Note that for GSE, in this case, after 48 queries, all the sequences remove one scene, thus resulting in a lexicographic ordering of the remaining queries. Figure~\ref{fig:results} shows the results of the algorithms in scenarios with different ($n$, $k$). Excluding the random queries case, the remaining algorithms can be ordered according to their performance from best to worst as follows: $$\text{GSE} \prec \text{MIS} \prec \text{base unrank} \prec \text{pattern-shifting} \prec \text{lex}$$ It is interesting to note that the pattern-shifting and lexicographic (lex) algorithms exhibit a performance worse than $\sigma$. Although the pattern-shifting algorithm has a remarkably simple implementation and has low complexity both in terms of computations and memory, a random ordering is likely to perform better. Moreover, base-unrank offers even better results with comparable costs. Another interesting point is that, for $n >> k$, as repetitions are less likely, the performance of the random queries converge to $\sigma$, as can be seen in Figure~\ref{fig:results} in the left panel. In this case, it becomes better than the pattern shifting algorithm for $n \ge 17$. The magnified areas in Figure~\ref{fig:results} show that the margin between the algorithms is relatively small, especially between GSE and MIS. A similar difference is expected between GSE and the optimal solution. To investigate the behavior of the algorithms, Figure~\ref{fig:results} (right panel) shows how many scenes are undiscovered in the same scenario ($n=20$, $k=3$), while iterating over the queries in the sequence. The lexicographic algorithm shows a repeated pattern where less and less scenes are removed up to some point, where suddenly more scenes are removed again. The pattern shifting algorithm shows a similar behaviour, though less pronounced. Interestingly, the lines of $\sigma$ and the base unrank algorithm are intersecting. The better score of the base unrank algorithm seems to be caused by it removing more scenes in the beginning. Of all algorithms without repetitions of queries, we saw that the lexicographic sequence performs worst. We find that the lexicographic sequence seems to be the worst case sequence for the described optimization problem. For $n < 6$, the lexicographic sequence is always among the sequences that maximizes the score as we found during the brute-force searches. Reversing the choice of GSE to choose the query that removes the least amount of scenes also yields a sequence with exactly the same score of the lexicographic sequence and showing the exact same behavior in the plot shown in Figure~\ref{fig:results} The best solutions generated for each one of the scenarios considered in this paper have been made available online\footnote{\url{http://www.esa.int/gsp/ACT/ai/projects/star_trackers.html}}. \begin{figure*}[p] \captionsetup[subfigure]{labelformat=empty} \captionsetup[subfigure]{justification=centering} \centering \subfloat[][Lexicographic]{ \includegraphics[width=.099\textwidth]{sequences/lexicographic_10_3} } \hspace{10pt} \subfloat[][Revolving Door]{ \includegraphics[width=.099\textwidth]{sequences/revdoor_10_3} } \hspace{10pt} \subfloat[][Pattern Shifting]{ \includegraphics[width=.099\textwidth]{sequences/mortari_10_3} } \hspace{10pt} \subfloat[][Greedy Scene\\Elimination]{ \includegraphics[width=.099\textwidth]{sequences/greedy_scene_elimination_10_3} } \hspace{10pt} \subfloat[][Minimally\\Intersecting\\Subsets]{ \includegraphics[width=.099\textwidth]{sequences/min_inters_10_3} } \hspace{10pt} \subfloat[][Base Unrank]{ \includegraphics[width=.099\textwidth]{sequences/base_unrank_10_3} } \caption{Visualisation of the sequences generated by the algorithms implemented in this paper, for $n=10$ and $k=3$.} \label{fig:sequences_10_3} \end{figure*} \section{Conclusion} We consider a star identification problem and we map it to a $k$-subsets (queries) optimal ordering problem. We provide an in depth analysis of its structure proving interesting mathematical properties that are used in the design and assessment of solution algorithms. A number of algorithms with different complexity and performance, covering a wide spectrum of the parameters $n$ and $k$, is proposed and proved to advance the state-of-the-art in star identification research. For small $n<7$ we are able to provide the optimal solutions, while for higher $n$ ($< 32$) our algorithms are only able to compute and score suboptimal solutions. We released our best solutions online\footnotemark[1]. The problem complexity is such that for even higher $n$ it is extremely challenging to even compute the considered scoring function. Nevertheless, we presented a class of algorithms with polynomial complexity in the output size $N$, producing better sequences than the average random sequence at least for the parameter range that could be tested. \bibliographystyle{IEEEtran}
08e99ea6a1c050c45c79e023219878485542c727
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} There are some techniques that have been used to recover the losses of entanglement and protect it from decoherence. Among of these techniques is the quantum purification \cite{Bennt,Deutsch,Metwally2002,Metwally2006}. Moreno et. al.,\cite{Moreno} showed that, it is possible to improve the coherence by using nested environments. Xiao et. al \cite{Xiao2016} displayed that the teleportation of Fisher information can be enhanced by using partial measurements. The possibility of purifying and concentrating entanglement by using local filtering is discussed by Yashodamma et. al, \cite{Yas2014}. The immunity and protecting the entanglement of accelerated qubit-qutrit system by using the local filtering technique is investigated by Metwally\cite{Metwally2016-0}. Recently, it has been shown that the local weak-reverse measurements,(WRM) can be used as a technique to protect and improve the correlation in qubit-qutrit systems \cite{Liang}. Xiao et. al.\cite{Xiao}, showed that, the entanglement losses caused by amplitude damping coherence can be retrieved by using the WRM. It is well known that, the accelerated systems lose some of its correlations, and consequently, their efficiencies to perform quantum information tasks decrease\cite{Metwally2013,Downes,Sagheer}. This decay depends on the value of the acceleration and the dimensional of the accelerated system\cite{Metwally2016}. Therefore we are motivated to investigate the possibility of using the weak-reverse measurement's technique to improve some properties of accelerated system. The main task of this contribution is quantifying the amount of entanglement between the two partners in the presence of weak-reverse measurements, investigating the dynamics of the accelerated/non accelerated information ( which is encoded on the accelerated/non accelerated particle) and the coherent information between the partners. The outline of this manuscript is described as follows: The suggested protocol is described in Sec.2. The suggested system and its analytical solutions are given in Sec.$3$. The effect of the weak-reverse'strengths on the degree of entanglement is discussed in Sec.$4$., where different initial states settings are considered. The dynamics of the accelerated, non-accelerated and coherent information is investigated in Sec.$5$. Finally, a conclusion is given in Sec.$6$. \section{Model} Our idea is to consider two partners, Alice and Bob, who share a two-qubit or two-qutrit systems. These systems are initially prepared in maximum or partial entangled states. Here, we consider only Alice's particle is accelerated, which may be a qubit or a qutrit. The partners decided in advance to use the weak-reverse measurement to protect their quantum communication channel. In the following steps, we describe the proposed protocol: \begin{enumerate} \item{Weak measurements:\\} Here, both users Alice and Bob perform the weak measurements on their own particles, either qubit or qutrit. We assume that, the two partners share initially a state defined by $\rho_{ab}^{ini}$. After performing the weak measurements, the partners share a new state defined by $\rho_{ab}^{out}$, \begin{equation} \rho_{ab}^{out}=\mathcal{W}_j\mathcal{W}_j\left(\rho_{ab}^{ini}\right)\mathcal{W}_j^{\dagger}\mathcal{W}_j^{\dagger}, ~i,j=q,t, \end{equation} where $\mathcal{W}_i$ are the weak measurements which are defined by the following operators \cite{kim,Xiao2013} \begin{eqnarray}\label{Wm} \mathcal{W}_q&=&\ket{0}\bra{0}+\sqrt{1-\alpha_q^{(\ell)}}~\ket{1}\bra{1}~\quad\mbox{(for qubit)},\nonumber\\ \mathcal{W}_t&=&\ket{0}\bra{0}+\sqrt{1-\alpha_t^{(1)}}~\ket{1}\bra{1}+\sqrt{1-\alpha_t^{(2)}}~\ket{2}\bra{2}~\quad\mbox{(for-qutrit)}, \end{eqnarray} and $\alpha_q^{(\ell)}$, $\alpha_t^{(\ell)}, \ell=1,2$ are the strengths of the weak measurements of the qubit and the qutrit, respectively. \item{Acceleration step:\\} It is assumed that, only Alice' particle is moving with a uniform acceleration meanwhile Bob's partial is assumed to be inertial. If the shared particles are of fermions types, then in the Minkowski frame the qubit system is transformed in the Rindler frame as, \begin{eqnarray} \ket{0_M}&=&\cos r\ket{0}_I\ket{0}_{II}+\sin r\ket{1}_I\ket{1}_{II},\quad \ket{1_M}=\ket{1}_{I}\ket{0}_{II} \end{eqnarray} while, for the qutrit system, the vacuum, the spin up and spin down states are transformed into Rindler space as, \begin{eqnarray} \ket{0_M}&=&\cos^2r\ket{0}_{I}\ket{0}_{II}+e^{i\phi}\sin r\cos r(\ket{\mathcal{U}}_{I}\ket{\mathcal{D}}_{II}+\ket{\mathcal{D}}_{I}\ket{\mathcal{U}}_{II}) +e^{2i\phi}sin^2r\ket{\mathcal{D}}_{I}\ket{\mathcal{P}}_{II}, \nonumber\\ \ket{\mathcal{U}_M}&=&cos r\ket{\mathcal{U}}_I\ket{0}_{II}+e^{i\phi}\sin r\ket{\mathcal{P}}_I\ket{\mathcal{U}}_{II}, \nonumber\\ \ket{\mathcal{D}_M}&=&cos r\ket{\mathcal{D}}_I\ket{0}_{II}-e^{i\phi}\sin r\ket{\mathcal{P}}_I\ket{\mathcal{D}}_{II}, \end{eqnarray} where $\ket{\mathcal{U}}, \ket{\mathcal{D}}$ and $\ket{\mathcal{P}}$ are the spin up, spin down and pair states, respectively. The acceleration $r$ is defined such that $\tan r=Exp[-\pi\omega\frac{c}{a}]$, ~$0\leq r\leq \pi/4$, $-\infty\leq a\leq\infty$, $\omega$ is the frequency, $c$ is the speed of light and $\phi$ is the phase space which can be absorbed in the definition of the operators \cite{Jason2013}. After tracing the particle in the second region (II), the final state represents the accelerated quantum channel between Alice and Bob, $\rho_{ab}^{acc}$. \item {Reverse measurement step:\\} In this step, the users apply the reverse measurement operations on the accelerated state $\rho_{ab}^{acc}$ to obtain the final state as, \begin{equation}\label{Final} \rho_{ab}^{Final}=\mathcal{R}_i\mathcal{R}_i\left(\rho_{ab}^{acc}\right)\mathcal{R}^{\dagger}_{i}\mathcal{R}^\dagger_{i},\quad i=q,t, \end{equation} where \begin{eqnarray}\label{Rm} \mathcal{R}_q&=&\sqrt{1-\beta_q^{(\ell)}}~\ket{0}\bra{0}+\ket{1}\bra{1},\quad \ell=1,2 \nonumber\\ \mathcal{R}_t&=&\sqrt{(1-\beta_t^{(1)})(1-\beta_t^{(2)})}~\ket{0}\bra{0}+\sqrt{1-\beta_t^{(1)}}~\ket{1}\bra{1} +\sqrt{1-\beta_t^{(2)}}~\ket{2}\bra{2}, \end{eqnarray} \end{enumerate} and $\beta_q^{(\ell)}$, $\beta_t^{(\ell)}, \ell=1,2$ are the strengths of the reverse measurement operations. Some properties of the final state Eq.(\ref{Final}), such as the behavior of entanglement, accelerated local non-local information are examined. Particularly, the effect of the initial state settings and the local operations strengths on these properties. \subsection{The suggested Systems} \begin{enumerate} \item {\bf Two-qubit system\\} In this section, it is assumed that the partners, Alice and Bob, share a partially entangled state, which is called $X-$ state. This state can be written, using the computational basis as, \begin{eqnarray}\label{x-state} \rho_x&=&\mathcal{B}_1\left(\ket{00}\bra{00}+\ket{11}\bra{11}\right)+\mathcal{B}_2\left(\ket{00}\bra{11}+\ket{11}\bra{00}\right) \nonumber\\ &+&\mathcal{B}_3\left(\ket{01}\bra{01}+\ket{10}\bra{10}\right)+\mathcal{B}_4\left(\ket{01}\bra{10}+\ket{10}\bra{01}\right) \end{eqnarray} where $\mathcal{B}_1=\frac{1}{4}(1+c_{33}), \mathcal{B}_2=\frac{1}{4}(c_{11}+c_{22}), \mathcal{B}_3=\frac{1}{4}(1-c_{33}), $ and $\mathcal{B}_4=\frac{1}{4}(c_{11}-c_{22})$, and $c_i,i=1,2,3$ are the diagonal elements of $3\times 3$ cross dyadic. The state (\ref{x-state}) can be described by its Bloch vectors, $\row{s}_j=0,j=1,2$ and a dyadic $\dyadic{C}$ with $c_{ij}=0$ for $i\neq j$. From this state, one can get maximum entangled classes, MES. For example, if we set $c_{11}=c_{22}=c_{33}=-1$, one gets the singled state, $\rho_{\psi^-}=\ket{\psi^-}\bra{\psi^-}$. The Werener state can be obtained if we set $c_{11}=c_{22}=c_{33}=-x$ , etc. As mentioned earlier, only Alice's particle will be accelerated \cite{MetwallyJMPB}. Due to the acceleration the degree of entanglement decreases. Therefore, the partners try to improve the degree of entanglement by applying the weak-reverse measurements. At the end of the protocol, the partners share the state. \begin{eqnarray}\label{Final-qubit} \rho_{ab}^{Final}&=&\frac{1}{\mathcal{N}_q}\Bigl\{\mathcal{B}_1\ket{00}\bra{00}+\tilde\mathcal{B}_2\ket{00}\bra{11} +\tilde\mathcal{B}_3\ket{01}\bra{01}+\tilde\mathcal{B}_4\ket{01}\bra{10} \nonumber\\ &+&\mathcal{B}_5\ket{10}\bra{10}+\tilde\mathcal{B}_6\ket{10}\bra{01} +\tilde\mathcal{B}_7\ket{11}\bra{11}+\tilde\mathcal{B}_8\ket{11}\bra{00}\Bigr\} \end{eqnarray} where, \begin{eqnarray} \tilde\mathcal{B}_1&=&c_1^2\mathcal{B}_1(1-\beta_q^{(1)})(1-\beta_q^{(2)}),\quad \nonumber\\ \tilde\mathcal{B}_2&=&c_1\mathcal{B}_2\sqrt{1-\beta_q^{(1)}}\sqrt{1-\beta_q^{(2)}}\sqrt{1-\alpha_q^{(1)}}\sqrt{1-\alpha_q^{(2)}}, \nonumber\\ \tilde\mathcal{B}_3&=&c_1^2\mathcal{B}_3(1-\beta_q^{(1)})(1-\alpha_q^{(2)}),\quad \nonumber\\ \tilde\mathcal{B}_4&=&c_1\mathcal{B}_4\sqrt{1-\beta_q^{(1)}}\sqrt{1-\beta_q^{(2)}}\sqrt{1-\alpha_q^{(1)}}\sqrt{1-\alpha_q^{(2)}}, \nonumber\\ \tilde\mathcal{B}_5&=&(1-\beta_q^{(2)})\left[s_1^2\mathcal{B}_1+(1-\alpha_q^{(1)})\mathcal{B}_3\right],\quad \tilde\mathcal{B}_6=\tilde\mathcal{B}_4 \nonumber\\ \tilde\mathcal{B}_7&=&(1-\alpha_q^{(1)})(1-\alpha_q^{(2)}) \mathcal{B}_1,~\quad \tilde\mathcal{B}_8=\tilde\mathcal{B}_2, \end{eqnarray} and $\mathcal{N}_q=\tilde\mathcal{B}_1+\tilde\mathcal{B}_3+\tilde\mathcal{B}_5+\tilde\mathcal{B}_6$ is the normalization factor and $c_1=\cos r, s_1=\sin r$. \item{Two-qutrit system:\\} In this section, we investigate the effect of the local-weak-reverse operations on a system consists of two qutrits, where it is assumed that only Alice's qutrit will be accelerated. In the computational basis, the initial sate of this system can be written as \cite{Parisio}, \begin{equation} \ket{\psi_{t}}=\frac{1}{\sqrt{2+\gamma^2}}\left(\ket{00}+\ket{11}+\gamma\ket{22}\right), \end{equation} where it turns into a maximum entangled state for $\gamma=1$. The users apply the protocol which is described in Sec.2. At the end the partners share the following state, \begin{eqnarray}\label{Final-qutrit} \rho_{ab}^{Final}&=&\frac{1}{\mathcal{N}_t}\Bigl\{\mathcal{D}_1\ket{00}\bra{00}+\mathcal{D}_2\ket{00}\bra{11}+\mathcal{D}_3\ket{10}\bra{10}+ \mathcal{D}_4\ket{11}\bra{00}+\mathcal{D}_5\ket{11}\bra{11} \nonumber\\ &&+\mathcal{D}_6\ket{20}\bra{20}+\mathcal{D}_7\ket{22}\bra{00}+\mathcal{D}_8\ket{22}\bra{11}+\mathcal{D}_9\ket{22}\bra{22}+\mathcal{D}_{10}\ket{00}\bra{22} \nonumber\\&& +\mathcal{D}_{11}\ket{11}\bra{22}\Bigr\}, \end{eqnarray} where \begin{eqnarray} \mathcal{D}_1&=&c_1^2\mathcal{R}_{00}^2\mathcal{A}_1, \quad \mathcal{D}_2=c_1^3\mathcal{R}_{00}\mathcal{R}_{11}\mathcal{A}_2, \quad \mathcal{D}_3=c_1^2s_1^2\mathcal{R}_{10}^2\mathcal{A}_1, \nonumber\\ \mathcal{D}_4&=&c_1^3\mathcal{R}_{00}\mathcal{R}_{11}\mathcal{A}_4,\quad \mathcal{D}_5=c_1^3\mathcal{R}_{11}^2\mathcal{A}_5,\quad \mathcal{D}_6=c_1^2s_1^2\mathcal{R}_{20}^2\mathcal{A}_1, \nonumber\\ \mathcal{D}_7&=&c_1^3\mathcal{R}_{22}\mathcal{R}_{00}\mathcal{A}_7,\quad \mathcal{D}_8=c_1^2\mathcal{R}_{22}\mathcal{R}_{11}\mathcal{A}_8,\quad \mathcal{D}_9=c_1^2\mathcal{R}_{22}^2\mathcal{A}_9, \nonumber\\ \mathcal{D}_{10}&=&\mathcal{R}_{00}\mathcal{R}_{22}\mathcal{A}_3c^3,\quad \mathcal{D}_{11}=\mathcal{R}_{11}\mathcal{R}_{22}\mathcal{A}_6c^2,\quad \end{eqnarray} The normalization factor is given by $\mathcal{N}_t=\mathcal{D}_1+\mathcal{D}_3+\mathcal{D}_5+\mathcal{D}_6+\mathcal{D}_9$ and the coefficients $\mathcal{A}_i, i=1..9$ are given by, \begin{eqnarray*} \mathcal{A}_1&=&\frac{1}{2+\gamma^2},~\quad\quad \mathcal{A}_2=\frac{\sqrt{1-\alpha_t^{(1)}}\sqrt{1-\alpha_t^{(2)}}}{2+\gamma^2},\quad \mathcal{A}_3=\frac{\gamma\sqrt{1-\alpha_t^{(1)}}\sqrt{1-\alpha_t^{(2)}}}{2+\gamma^2},~\quad\quad \nonumber\\ \mathcal{A}_4&=&\mathcal{A}_2,~\quad\quad \mathcal{A}_5=(2+\gamma^2)\mathcal{A}_2^2,~\quad\quad \mathcal{A}_6=\gamma\mathcal{A}_2\sqrt{1-\alpha_t^{(1)}}\sqrt{1-\alpha_t^{(2)}},\quad \nonumber\\ \mathcal{A}_7&=&\frac{\gamma}{2+\gamma^2}\sqrt{1-\alpha_t^{(1)}}\sqrt{1-\alpha_t^{(2)}},~\quad \mathcal{A}_8=\frac{\gamma(1-\alpha_t^{(1)})}{2+\gamma^2}\sqrt{1-\alpha_t^{(2)}}\sqrt{1-\alpha_t^{(2)}},\quad \nonumber\\ \mathcal{A}_9&=&\frac{\gamma^2}{2+\gamma^2}(1-\alpha_t^{(1)})(1-\alpha_t^{(2)}), \end{eqnarray*} and $\mathcal{R}_{ij},i,j=0,1,2$ are given by \begin{eqnarray*} \mathcal{R}_{00}&=&\sqrt{(1-\beta_1^{(1)})(1-\beta_2^{(1)})}\sqrt{(1-\beta_1^{(2)})(1-\beta_2^{(2)})},\quad \nonumber\\ \mathcal{R}_{01}&=&\sqrt{1-\beta_1^{(2)}}\sqrt{(1-\beta_1^{(1)})(1-\beta_2^{(1)})},\quad \nonumber\\ \mathcal{R}_{02}&=&\sqrt{1-\beta_2^{(2)}}\sqrt{(1-\beta_1^{(1)})(1-\beta_2^{(1)})},\quad \nonumber\\ \mathcal{R}_{10}&=&\sqrt{1-\beta_1^{(1)}}\sqrt{(1-\beta_1^{(2)})(1-\beta_2^{(2)})}, \nonumber\\ \mathcal{R}_{11}&=&\sqrt{1-\beta_1^{(1)}}\sqrt{1-\beta_1^{(2)}},~\quad\quad \mathcal{R}_{12}=\sqrt{1-\beta_1^{(1)}}\sqrt{1-\beta_2^{(2)}}, \nonumber\\ \mathcal{R}_{20}&=&\sqrt{1-\beta_2^{(1)}}\sqrt{(1-\beta_1^{(2)})(1-\beta_2^{(2)})},~\quad\quad \mathcal{R}_{22}=\sqrt{1-\beta_2^{(1)}}\sqrt{1-\beta_2^{(2)}}. \end{eqnarray*} \end{enumerate} \section{Entanglement} To quantify the survival degree of entanglement $\mathcal{E}$ contained in the system, we use the negativity as a measure. The negativity is defined as, \begin{equation} \mathcal{E}=max\left(0,\sum_i{\lambda_i}\right), \end{equation} where $\lambda_i,$ are the eigenvalues of the partial transpose of $\rho^{T_a}_{ab}$\cite{Karpat}. \begin{figure} \begin{center} \includegraphics[width=19pc,height=19pc]{Fig1a.eps}~~\quad \put(-160,35){\Large$ r$} \put(-240,110){\Large$\mathcal{E}$} \put(-20,70){\Large$\alpha_q$} \put(-60,210){$(a)$}~~\quad \includegraphics[width=19pc,height=19pc]{Fig1b.eps} \put(-160,35){\Large$r$} \put(-235,100){\Large$\mathcal{E}$} \put(-15,70){\Large$\alpha_q$} \put(-60,210){$(b)$} \caption{The entanglement of the final state of the two-qubit system (8) where we set $\alpha_q^{(1)}=\alpha_q^{(2)}= \beta_q^{(1)}=\beta_q^{(2)}=\alpha_q$. (a) The initial state is prepared in a maximum entangled states, i.e., $(c_{11}=c_{22}=c_{33}=1$). (b) The initial state is prepared in a partial entangled state (Werner state, $c_{11}=c_{22}=c_{33}=0.7$), } \end{center} \end{figure} Fig.(1a), displays the behavior of entanglement of the final state (\ref{Final-qubit}), where it is assumed that, the partners share initially a singlet state i.e., $c_{11}=c_{22}=c_{33}=1$ and we have set $\alpha_q^{(1)}=\alpha_q^{(2)}=\beta_q^{(1)}=\beta_q^{(2)}=\alpha_q$. It is clear that, for $r=0$ the degree of entanglement is maximum i.e., ($\mathcal{E}=1)$. The general behavior shows that, the entanglement decreases as the acceleration parameter, $r$ increases. On the other hand, as one increases the strengths of the local operations, the entanglement increases. However, for further values of $\alpha\in[0.75,1]$, the entanglement decreases. The behavior of entanglement of a system initially prepared in a partially entangled state of Werner type is displayed in Fig.(1b). The general behavior is similar to that for the MES (Fig.(1a)), but with smaller decay rate. On the other hand, for $\alpha\in[0.75,1]$, the entanglement re-appears again while, it vanishes completely for the MES. \begin{figure}[t!] \begin{center} \includegraphics[width=19pc,height=19pc]{Fig2a.eps}~~\quad \put(-180,35){\Large$ r$} \put(-240,130){\Large$\mathcal{E}$} \put(-25,35){\Large$\alpha_t$} \put(-60,210){$(a)$}~~\quad \includegraphics[width=19pc,height=19pc]{Fig2b.eps} \put(-190,35){\Large$r$} \put(-235,130){\Large$\mathcal{E}$} \put(-30,40){\Large$\alpha_t$} \put(-60,210){$(b)$} \caption{The of entanglement, of the two-qutrit system Eq.(11), where (a) maxim entangled state ($\gamma=1$) and (b) partial entangled state i.e., ($\gamma=0.5$), where set $\alpha_t^{(1)}=\alpha_t^{(2)}= \beta_t^{(1)}=\beta_t^{(2)}=\alpha_t$. } \end{center} \end{figure} The dynamics of entanglement under the effect of the local weak-reverse measurement for a system is initially prepared in a two-qutrit system is described in Fig.(2). The general behavior is similar to that predicted in Figs.$(1)$, where the entanglement decays as $r$ increases. However, the upper and lower bounds of entanglement depend on initial state settings. It is clear that, for the two-qutrit system the entanglement increases slowly compared with that shown in Fig.(1). Also, as displayed in Fig.(2b), the increasing rate of entanglement for PES is larger than that shown for MES. From the previous figures, one concludes that the weak and reverse measurements can recover the loss of entanglement for small values of accelerations as the local operations' strengths increase for different intervals. The length of these intervals depends on the initial states settings: two-qubit, or two qutrits, maximum / partial entangled states. The increasing rate of entanglement is much larger for systems that are initially prepared in partial entangled states. The partially entangled 2- qubits state is more robust than the partially entangled two-qutrit state. \section{Local and Non-local Information} \subsection{Two-qubit system} In this subsection, we quantify Alice's and Bob's information, which represents the accelerated $(\mathcal{I}_a)$ and non-accelerated,($\mathcal{I}_b$) information, respectively. Moreover, the amount of the non-local information between the partners defined by the coherent information $(\mathcal{I}_{coh})$ will be quantified also. In an explicit form, theses three types of information can be written as, \begin{eqnarray} \mathcal{I}_{a}&=&-\frac{\tilde\mathcal{B}_1+\tilde\mathcal{B}_3}{\mathcal{N}_{q}}\log\left(\frac{\tilde\mathcal{B}_1+\tilde\mathcal{B}_3}{\mathcal{N}_{q}}\right)- \frac{\tilde\mathcal{B}_5+\tilde\mathcal{B}_7}{\mathcal{N}_{q}}\log\left(\frac{\tilde\mathcal{B}_5+\tilde\mathcal{B}_7}{\mathcal{N}_{q}}\right) \nonumber\\ \mathcal{I}_{b}&=&-\frac{\tilde\mathcal{B}_1+\tilde\mathcal{B}_5}{\mathcal{N}_{q}}\log\left(\frac{\tilde\mathcal{B}_1+\tilde\mathcal{B}_5}{\mathcal{N}_q}\right)- \frac{\tilde\mathcal{B}_3+\tilde\mathcal{B}_7}{\mathcal{N}_{q}}\log\left(\frac{\tilde\mathcal{B}_3+\tilde\mathcal{B}_7}{\mathcal{N}_{q}}\right) \nonumber\\ I_{coh}&=&\sum_{i=1}^{4}\mu_i\log(\mu_i), \end{eqnarray} where \begin{eqnarray*} \mu_{1,2}&=&\frac{1}{2\mathcal{N}_{q}}\Bigl\{(\tilde\mathcal{B}_1+\tilde\mathcal{B}_7) \pm\sqrt{(\tilde\mathcal{B}_1-\tilde\mathcal{B}_7)^2+4\tilde\mathcal{B}_2\tilde\mathcal{B}_8}\Bigr\}, \nonumber\\ \mu_{3,4}&=&\frac{1}{2\mathcal{N}_{q}}\Bigl\{(\tilde\mathcal{B}_3+\tilde\mathcal{B}_5) \pm\sqrt{(\tilde\mathcal{B}_3-\tilde\mathcal{B}_5)^2+4\tilde\mathcal{B}_4\tilde\mathcal{B}_6}\Bigr\}, \end{eqnarray*} \begin{figure} \begin{center} \includegraphics[width=19pc,height=19pc]{Fig3a.eps}~~\quad \put(-160,30){\Large$r$} \put(-255,100){$\mathcal{I}_{coh}$} \put(-20,60){\Large$\alpha_q$} \put(-60,200){$(a)$}~\quad \includegraphics[width=19pc,height=19pc]{Fig3b.eps} \put(-160,30){\Large$r$} \put(-235,97){$\mathcal{I}_a$} \put(-15,60){\Large$\alpha_q$} \put(-60,200){$(b)$} \caption{The local and non-local information for a system is initially prepared in a MES of two-qutrit system. (a) The coherent information between Alice and Bob, $\mathcal{I}_{coh}$ and (b) the accelerated local information $\mathcal{I}_a$. } \end{center} \end{figure} Fig.(3) displays the behavior of the coherent information and the accelerated local information, where we set $\alpha_q^{(1)}=\alpha_q^{(2)}= \beta_q^{(1)}=\beta_q^{(2)}=\alpha_q$. At zero acceleration, both types of information are maximum. As Alice's particle is accelerated, the coherent information decreases. However, for a fixed value of $r$, the coherent information increases as the strengths of the local operations increase. Meanwhile, the accelerated local information increases at the expense of the coherent information. For larger values of $\alpha_q$, the accelerated information vanishes while the coherent information increases. This coherent information, no longer represents quantum information but it describes a classical information, because the two particle are almost separable for larger values of $\alpha_q$ as displayed in Fig.(1). In Fig.(4), we scrutinize the effect of the WRM on the local and non local information, as well as on the coherent information. The behavior of the accelerated information $\mathcal{I}_a$ and the non-accelerated information $\mathcal{I}_b$ are described in Fig.(4a), where it is assumed that, the system is either initially prepared in a maximum entangled state (MES) or in a partial entangled state (PES). At zero acceleration $(r=0)$ the information which is encoded in Alice's and Bob's qubit is maximum i.e. $\mathcal{I}_a=\mathcal{I}_b=1~ bit$. For small values of $r\in[0,2.5]$, $I_a$ and $\mathcal{I}_b$ are remaining maximum $(1~bit)$. For larger values of $r$, the accelerated information $\mathcal{I}_a$ decreases as the acceleration $r$ increases, while Bob's information (non-accelerated information) is slightly affected due to the local WRM. Moreover, the decay rate of the accelerated information for a system prepared initially in a partial entangled state is smaller than that displayed for systems that are initially prepared in MES. This shows that, the possibility of protecting the accelerated local information encoded in PES by the WRM is much better than protecting that coded in MES. \begin{figure} \begin{center} \includegraphics[width=20pc,height=15pc]{Fig4a.eps}~\quad \put(-80,140){\Large$\mathcal{I}_{b}$} \put(-80,90){\Large$\mathcal{I}_{a}$} \put(-240,90){$\mathcal{I}_{Local}$} \put(-110,5){\Large$r$} \put(-60,150){$(a)$} \includegraphics[width=20pc,height=15pc]{Fig4b.eps} \put(-240,120){\Large$\mathcal{I}_{coh}$} \put(-240,60){\Large$\mathcal{I}_{a}$} \put(-110,5){\Large$r$} \put(-60,150){$(b)$} \caption{(a)Alice and Bob's information for a system is initially prepared in a MES or PES. The solid and dash curves represent Alice's for system is initially prepared in MES and PES, respectively, while the dash-dot and dot curves represent Bob's local information for MES and PES respectively, where we set $\alpha_q^{(1)}=\alpha_q^{(2)}= \beta_q^{(1)}=\beta_q^{(2)}=\alpha_q=0.5$ (b) The local and non local information for different values of $\alpha_q=0.5,0.8,0.9$, for the solid,dash and dot curves respectively. } \end{center} \end{figure} The effect of the local WRM on the accelerated local information, $\mathcal{I}_a$ and the coherent information $\mathcal{I}_{coh}$ is depicted in Fig.(4b), where it is assumed that the system is initially prepared in the MES. In these calculations, we consider that the strengths of the local weak measurement and the reverse measurements are equal $(\alpha_q=0.5)$. The behavior of information shows that, both types of the information are maximum at $r=0$. As $r$ increases, the accelerated local information $\mathcal{I}_a$ has maximum values for $r\in[0,2.5]$, while the coherent information decreases for any value of $r>0$. It is clear that, as the strengths of the WRM increase, both types of information decrease. However the decay rate of $\mathcal{I}_{coh}$ is much larger than that for accelerated local information, $\mathcal{I}_a$. For larger values of $r$ and $\alpha_q$, the coherent information increases at the expense of the local accelerated information. From these figures, one concludes that the accelerated information for small values of acceleration can be protected by using the weak-reverse measurement. The possibility of protecting the accelerated information which is encoded in partial entangled states is larger than that encoded in a maximum entangled states. The local measurement has a very slight effect on the non-accelerated local information. For larger values of accelerations, the coherent information increases at the expense of the accelerated information \subsection{Two-qutrit system} Fig.(5) shows the behavior of the coherent information $\mathcal{I}_{coh}$ and the accelerated information $\mathcal{I}_a$ for a system initially prepared in an entangled maximum two-qutrit system. It is clear that, both types of information decrease as the acceleration $r$ increases. The decay rate is larger for the coherent information compared with that depicted for the accelerated information. However, as the strengths of the local operations increase both types of information are improved. For $\alpha_t\in[0.8,~1]$, the coherent information increases at the expense of the accelerated information. \begin{figure} \begin{center} \includegraphics[width=19pc,height=19pc]{Fig5a.eps}~~\quad \put(-160,28){\Large$r$} \put(-255,100){$\mathcal{I}_{coh}$} \put(-20,60){\Large$\alpha_t$} \put(-60,205){$(a)$} \includegraphics[width=19pc,height=19pc]{Fig5b.eps} \put(-160,30){\Large$r$} \put(-230,98){$\mathcal{I}_a$} \put(-15,60){\Large$\alpha_t$} \put(-60,205){$(b)$} \caption{The same as Fig.(3), but the system is initially prepared in a maximum entangled 2-qutrits state. } \end{center} \end{figure} The dynamics of different types of information coded in a system initially prepared in a maximum two- qutrit entangled state is described in Fig.(6). The behavior of the accelerated local information $\mathcal{I}_a$ and the non-accelerated local information, $\mathcal{I}_b$ is displayed in Fig.(6a), where it is assumed that, the system initially prepared in MES or PES. It is clear that, the accelerated information for a system is initially prepared in a MES is almost stable and it is a maximum for any value of $r$. On the other hand, starting from PES, the initial values of the accelerated information is smaller than that depicted for MES. However, the upper bounds of $\mathcal{I}_a$ increase as the acceleration $r$ increases. These upper bounds become larger than that depicted for MES for $r\in[0,~0.8]$. For non-accelerated information the behavior is completely different from those shown for the two-qubit system, where it is always smaller than the accelerated information. Moreover, it decreases as $r$ increases. Fig(6b) shows the behavior of the coherent information and the accelerated information for different values of the local operation's strengths, with initial system prepared in a MES of the two-qutrit system. It is clear that, $\mathcal{I}_{coh}$ decreases as $\alpha_t$ and $r$ increase. For smaller values of $\alpha_t$, the coherent information decreases faster than that displayed for larger values of $\alpha_t$. On the other hand, for larger values of $\alpha_t$ and $r$, the coherent information increases to reach its upper bounds as $r\to \infty$. Also, the general behavior of the accelerate information shows that, $\mathcal{I}_a$ decreases as the strengths of the local operations increases. However, the decay rate is smaller than the coherent information. This explain the decay of the non-accelerated information displayed in Fig.(6a). \begin{figure}[t!] \begin{center} \includegraphics[width=20pc,height=15pc]{Fig6a.eps} \put(-240,120){\Large$\mathcal{I}_{a}$} \put(-240,60){\Large$\mathcal{I}_{b}$} \put(-110,5){\Large$r$} \put(-60,150){$(a)$} \includegraphics[width=20pc,height=15pc]{Fig6b.eps} \put(-240,120){\Large$\mathcal{I}_{coh}$} \put(-240,60){\Large$\mathcal{I}_{a}$} \put(-110,5){\Large$r$} \put(-60,150){$(b)$} \caption{ Local and non-local information of a MES of two-qutrit system where $\alpha_t=0.5,0.8,0.9$. } \end{center} \end{figure} From these figures one can extract some important facts. The local information which is encoded in the accelerated part of the composite system can be protect by using the WRM. The local accelerated information which is encoded in a partial entangled state can be improved even for larger accelerations. The non-accelerated local information is very sensitive to the local operations (WRM), compared with that shown for the two-qubit system. The decay rate of accelerated information is much smaller than that shown for the coherent information. For larger values of accelerations, the coherent information increases at the expense of the non-accelerated local information. \section{Conclusion} In this contribution, we assume that two users, Alice and Bob, share 2-qubit or 2-qutrit states as a communication channels to perform some quantum information tasks. Alice is allowed to accelerate her qubit/qutrit. Subsequently, the coherence of the communication channel between the partners decreases. Thus, we try to investigate the possibility of utilizing the weak-reverse technique to improve the quantum correlations between the users. It is shown that, the initial state settings, the values of the acceleration and the strengths of the weak-reverse operations represent control keys to the behavior of entanglement, local and non-local information. It is demonstrated that, for small values of acceleration the improving rate of the quantum correlation of the communication channels increases as the strengths of the weak-reverse operation increase. However, as one increases the local operations' strengths further the correlation decreases. Our results show that, starting from less entangled state the improving rate of correlations is better than that depicted for maximum entangled states. On the other hand, the increasing rate of the correlation for the two-qutrit system is larger for large values of the acceleration comparing with that displayed for the two-qubit-system. Additionally, the behavior of the local information which includes the accelerated and non-accelerated information is assayed. It is shown that, for the two-qubit system, the non-accelerated information is almost stable and slightly decreases for systems initially prepared in maximum entangled state. But the accelerated information decreases as the acceleration increases. Moreover, the encoded information in a system initially prepared in partial entangled state is more robust than that encoded in a system initially prepared in maximum entangled states. Theses results are dramatically changed for the two-qutrit systems, where the non- accelerated information is more fragile than the accelerated information. Also, the encoded information in the partial entangled state of the two-qutrit system is improved as the acceleration increases under weak-reverse measurements. Finally, the general behavior of the coherent information displays that. it decreases as the acceleration increases. For a fixed value of acceleration, the coherent information increases as the strength of the weak-reverse measurement increases. However, for the two-qubit system, larger values of these strengths cause a sudden decay of information for small values of the accelerations and gradually decay for larger values of these accelerations. Meanwhile, for the two-qutrit system, the coherent information slightly increases at larger values of the weak-reverse measurements' strengths. The changes happened at the expense of the local information. {\it In conclusion}, it is possible to improve and protect the accelerated communication channel by using weak-reverse measurements. The possibility of protecting the local encoded information on an accelerated 2-qutrit system is larger than that displayed for 2-qubit system. The restrained decay rate of the coded information on accelerated partial entangled state is better than that displayed for maximum entangled state. The encoded information in the accelerated part of the two-qutrit system can be improved by using the local operations, while it decays for the two-qubit system. Therefore, to perform quantum key distribution or quantum coding protocols using accelerated system, it is better to use partial entangled 2-qutrit systems with larger accelerations.
cf3200fa3ec0112e1743b2c6e5b289cc793b4c2e
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction. The Burgers equation} The main topics in this paper are directed polymers and the Burgers equation. We begin with the Burgers equation which is one of the basic nonlinear evolutionary PDEs. It has various interpretations and applications. Although Burgers himself introduced it as a fluid dynamics model in an attempt to create a simplified picture of turbulence (see \cite{Burgers:MR0001147}, \cite{burgers1973nonlinear}), the equation along with its modifications has been used to model diverse real world phenomena such as interface growth, traffic, or the distribution of matter in the Universe. Various aspects of the Burgers turbulence are discussed in~\cite{Bec-Khanin:MR2318582}. The Burgers equation is also directly related to the KPZ equation that has been intensively studied lately. For an introduction to KPZ, see~\cite{Corwin:MR2930377} or~\cite{Quastel:MR3098078}. In one dimension, the Burgers equation is \begin{equation} \label{eq:Burgers} \partial_t u + u \partial_x u=\varkappa\partial_{xx} u + f. \end{equation} Under the fluid dynamics interpretation, the equation describes the evolution of a velocity profile $u$ of particles moving along the real line. The velocity of the particle located at time $t\in{\mathbb R}$ at point $x\in{\mathbb R}$ is denoted by $u(t,x)\in{\mathbb R}$. The left-hand side of~\eqref{eq:Burgers} represents the acceleration of the particle, so the right-hand side must contain all the forces acting on the particle. In this pressureless model, particles are subject to external forcing $f=f(t,x)$ and viscous friction forces represented by the term $\varkappa\partial_{xx}u(t,x)$, where $\varkappa\ge0$ is the viscosity constant. The following (viscous) Hamilton--Jacobi--Bellman (HJB) equation: \begin{equation} \label{eq:HJB} \partial_t U + \frac{(\partial_x U)^2}{2}=\varkappa\partial_{xx} U + F, \end{equation} is tightly connected to the Burgers equation. Namely, if $U$ is a solution of~\eqref{eq:HJB}, then $u=\partial_x U$ solves~\eqref{eq:Burgers} with $f=\partial_x F$. Under mild assumptions on $f$ and the initial conditions, the Cauchy problem for~\eqref{eq:Burgers} has smooth classical solutions for~$\varkappa>0$. In fact, the Hopf--Cole logarithmic transformation reduces the problem to the linear heat equation with multiplicative potential. This linear equation can be solved using the classical Feynman--Kac formula. The works~\cite{Hopf:MR0047234},~\cite{Cole:MR0042889} by Hopf and Cole were, in fact, preceded by \cite{Florin:MR0029605}, see the note~\cite{Biryuk} for some historical comments. Another way to represent solutions of viscous HJB equations is via stochastic control, see~\cite{Fleming-Soner:MR2179357} for systematic treatment of stochastic control. If~$\varkappa=0$, then even smooth initial velocity profiles result in formation of discontinuities called shock waves. In this important case, one has to work with appropriately defined generalized solutions that can be obtained from the smooth solutions via a limiting $(\varkappa\to0)$ procedure. The solutions can be characterized through a variational principle. In this paper we continue the study of the case where the forcing $f=f_{\omega}(t,x)$ is a stationary random field, the argument $\omega$ being an element of a probability space $(\Omega,\mathcal{F},\mathsf{P})$. Ergodic properties of the Burgers equation (and its generalizations) with random forcing have been studied in various settings in \cite{Sinai:MR1117645}, \cite{ekms:MR1779561}, \cite{Iturriaga:MR1952472}, \cite{Khanin-Hoang:MR1975784}, \cite{Suidan:MR2141893}, \cite{Gomes-Iturriaga-Khanin:MR2241814}, \cite{Dirr-Souganidis:MR2191776}, \cite{yb:MR2299503}, \cite{Bakhtin-quasicompact}, \cite{BCK:MR3110798}, \cite{Debussche-Vovelle:MR3418750}, \cite{kickb:bakhtin2016}. The details of the systems considered in those papers vary. In one dimension, the equation may be considered on a circle, a compact segment, or the entire real line. The kinds of time-stationary forcing that have been considered are: white noise, kick (applied at a discrete sequence of times), Poissonian (concentrated at configuration points of a Poisson point process). It is usual to assume the zero-range dependence in time (this ensures the Markov property of the solutions but is not absolutely necessary). Several kinds of spatial dependence structure and behavior at infinity have been considered. Also, the role of the external forcing can be played by random boundary conditions. Each setting calls for its own toolbox, but despite the variety of approaches and methods that have been employed, there are several general features that we proceed to describe informally. In dissipative systems, statistically steady states often emerge due to some form of energy balance. For the Burgers equation, the energy is pumped into the system by the external forcing and dissipated either due to the Laplacian viscosity friction term (in the case of positive viscosity), or at shocks (in the zero viscosity case). So it is natural to expect existence of stationary regimes. The question of ergodicity of the invariant measures for Markov processes including those associated with stochastic PDEs can often be approached by studying regularity properties of the transition probabilities or controllability properties, see, e.g.,~\cite{H-M:MR2478676} for the stochastic Navier--Stokes case. In contrast with turbulence described by the Navier--Stokes system, the dynamics generated by Burgers equation is dominated by contraction. So the random dynamical system approach turns out to be more fruitful and gives more detailed information about the pathwise behavior of the system. Namely, it is natural and beneficial to study the stochastic flow, i.e., the self-consistent (satisfying the so called {\it cocycle} property) family of random operators $\Phi^{st}_\omega$ constructing the solution $\Phi^{st}_\omega u$ at time $t$ given the initial condition $u$ at time~$s$. For various settings, it turns out that one can describe ergodic components in the following terms: for two velocity profiles $u^1$ and $u^2$ in the same ergodic component, $\Phi_\omega^{st}u^1$ and $\Phi_\omega^{st}u^2$ get close to each other as $t-s\to\infty$. In other words, the evolution over long time intervals depends mostly on the random forcing while the dependence on the initial condition vanishes, which can be interpreted as loss of memory in the system. Moreover, with probability one, there is a limit \begin{equation} \label{eq:general-pullback} u_{t,\omega}=\Phi^{-\infty t}_\omega u^0=\lim_{s\to-\infty} \Phi^{s t}_\omega u^0, \end{equation} and it does not depend on the initial condition $u^0$ within an ergodic component. The resulting family $(u_{t,\omega})$ of velocity profiles forms a global solution, i.e., \[ u_{t,\omega}= \Phi_\omega^{st}u_{s,\omega},\quad s<t, \] and is non-anticipating, i.e., $u_t$ depends only on the history of the forcing up to time~$t$. Moreover, for almost every $\omega$, $(u_{t,\omega})$ is a unique global solution with values in the given ergodic component. This statement along with the pullback attraction property~\eqref{eq:general-pullback} is often called One Force --- One Solution Principle (1F1S). The study of ergodic properties of solutions of~\eqref{eq:Burgers} with random forcing began in \cite{Sinai:MR1117645} where the evolution was considered on the circle (or one-dimensional torus) ${\mathbb T}^1={\mathbb R}^1/ {\mathbb Z}^1$ (i.e., all the functions involved were assumed or required to be space-periodic), the forcing was assumed to be white in time and smooth in the space variable, and a mixing statement showing loss of memory in the system was proved. The key consideration in this paper is the view at the iterative application of the Feynman--Kac formula as the product of positive operators. In~\cite{Kifer:MR1452549} the connection with the directed polymers in random environments was noticed and used for the first time. With the help of the Hopf--Cole transform and Feynman--Kac formula it was shown that for the high-dimensional version of~\eqref{eq:Burgers} and sufficiently small forcing (this situation is known as {\it weak disorder} in the studies of directed polymers in random environments), certain series in the spirit of perturbation theory converge and can be used to define global attracting solutions of the Burgers equation. In~\cite{ekms:MR1779561}, the zero viscosity case on the circle was considered. Solutions of the Burgers equation with zero viscosity admit a variational Hamilton -- Jacobi -- Bellman -- Hopf -- Lax -- Oleinik representation. The minimizing paths in the variational principle can be identified with particle trajectories, and the analysis of solutions over long time intervals involves the study of asymptotic properties of those minimizers. Since the mean velocity is preserved by the Burgers system, all velocity profiles in one ergodic component have the same mean. One of the main results of~\cite{ekms:MR1779561} is that all functions with the same mean form one ergodic component, i.e., there is a unique invariant measure for the corresponding Markov dynamics on this set. Moreover, for each mean velocity, 1F1S holds on the associated ergodic component. The global solution is defined by a family of one-sided infinite action minimizers stretching into the infinite past. Also, {\it hyperbolicity} holds, i.e., all these minimizers are exponentially asymptotic to each other. In \cite{Iturriaga:MR1952472}, this program was repeated for the multi-dimensional version of the inviscid Burgers equation on the torus ${\mathbb T}^d={\mathbb R}^d/{\mathbb Z}^d$, $d\in{\mathbb N}$, and in~\cite{Gomes-Iturriaga-Khanin:MR2241814}, it was extended to the positive viscosity case. Unlike \cite{Sinai:MR1117645}, the approach of~\cite{Gomes-Iturriaga-Khanin:MR2241814} was based on stochastic control. In fact, for a fixed mean velocity, a unique global solution is constructed using optimal control of diffusions on semi-infinite time intervals stretching to the infinite past. The variational character of the stochastic control approach allowed to show that as $\varkappa\to 0$, the optimally controlled diffusions converge to the one-sided action minimizers. This also allowed to deduce the convergence of invariant distributions as $\varkappa\to 0$. In~\cite{yb:MR2299503}, 1F1S was established for the Burgers equation with random boundary conditions. Given an appropriate notion of generalized solutions and the associated variational characterization, the argument is very simple. It turns out that it takes a finite random time to erase all the memory about the initial condition, so the system exhibits an extreme form of contraction. In all the results discussed above (see also \cite{Dirr-Souganidis:MR2191776} and \cite{Debussche-Vovelle:MR3418750} that do not use variational or stochastic representations and use PDE tools instead), the space was assumed to be compact, being a torus or a segment, except \cite{Kifer:MR1452549}. Extending those results to noncompact situations turned out to be a nontrivial task. Quasi-compact settings where the system is considered on the entire real line but the forcing is mostly concentrated in a compact part, were studied in~\cite{Khanin-Hoang:MR1975784}, \cite{Suidan:MR2141893}, \cite{Bakhtin-quasicompact}. However, truly noncompact situations with space-time homogeneous random forcing in one dimension for positive or zero viscosity presented serious difficulties. In the noncompact case, there is much less rigidity in the behavior of optimal paths or diffusions used in the representation of solutions, and they are much harder to control. Also, the approach of~\cite{Kifer:MR1452549} is useful only in the weak disorder case and fails in dimension~$1$. In the zero viscosity case, the ergodic theory of the Burgers equation on the real line without compactness or periodicity assumptions was constructed in~\cite{BCK:MR3110798} for forcing given by space-time Poisson point process, and in~\cite{kickb:bakhtin2016} for kick forcing. Similarly to the compact case, the ergodic components are essentially formed by velocity profiles with common mean, but establishing 1F1S on each ergodic component required using methods originating from studies of long geodesics in the last-passage percolation theory. In the Poissonian forcing case, due to the discrete character of the forcing, all the one-sided minimizers giving rise to the global solution coalesce, strengthening the hyperbolicity property for the spatially smooth periodic forcing case. However, the behavior of minimizers in the kick forcing case is more complicated. Although they are expected to be asymptotic to each other, only a much weaker {\it liminf} substitute of hyperbolicity was proved in~\cite{kickb:bakhtin2016}. The main goal of the present paper is to extend this program to the Burgers equation with positive viscosity, space-time homogeneous random kick forcing, and without any compactness assumptions. Our main results at a formal level look similar to the ergodic results for the inviscid case with kick forcing: ergodic components are formed by velocity profiles with common mean value, and on each ergodic component 1F1S holds. This extension to the viscous case is natural to expect because positive viscosity means stronger dissipation and contraction. However, a very important feature of our work is that in order to analyze the Burgers equation we rely on the Feynman--Kac formula and the associated directed polymer model. For that directed polymer model in $1+1$ dimensional random environment, we prove a whole series of new results that are of independent interest. These results include: the existence of finite nonrandom quenched free energy density for every fixed asymptotic slope and its quadratic dependence on the slope; a concentration inequality for free energy and a bound on the fluctuation exponent; straightness estimates and a bound on the transversal fluctuation exponent. We use those results to construct thermodynamic limits satisfying the DLR conditions and prove their uniqueness for every fixed asymptotic slope. We also prove results on limiting ratios of partition functions playing the role of Busemann functions and show that a version of hyperbolicity property holds true. Namely, we show that the marginals of any two infinite volume polymer measures with the same slope are asymptotic to each other in the total variation distance. We state and discuss our main results on directed polymers in section~\ref{sec:polymers} that also has an introductory character. Before that, we explain the Burgers equation setting in Section~\ref{sec:setting} and state the main 1F1S results in Section~\ref{sec:main_results}. \bigskip {\bf Acknowledgments.} The authors are grateful to Konstantin Khanin for stimulating discussions. They also thank Firas Rassoul-Agha for useful bibliographical suggestions. Yuri Bakhtin gratefully acknowledges partial support by NSF through grant DMS-1460595. \section{The setting. Kick forcing} \label{sec:setting} The main model that we study in this paper is the Burgers equation with kick forcing of the following form: \begin{equation*} f(t,x)=\sum_{n\in{\mathbb Z}} f_{n}(x)\delta_{n}(t). \end{equation*} This means that the additive forcing is applied only at integer times. At each time $n\in{\mathbb Z}$, the entire velocity profile receives an instantaneous macroscopic increment equal to $f_n$: \begin{equation} \label{eq:effect-of-one-kick} u(n,x)=u(n-0,x)+f_{n}(x),\quad x\in {\mathbb R}, \end{equation} and between the integer times the velocity field evolves according to the unforced viscous Burgers equation \begin{equation} \label{eq:unforced-Burgers} \partial_t u + u \partial_x u=\varkappa\partial_{xx} u. \end{equation} We assume that the potential $F=F_{n,\omega}(x)$ of the forcing \[ f_n(x)=f_{n,\omega}(x)=\partial_x F_{n,\omega}(x), \quad n\in{\mathbb Z},\ x\in{\mathbb R},\ \omega\in \Omega, \] is a stationary random field (i.e., its distribution is invariant under space-time shifts) defined on some probability space $(\Omega,\mathcal{F},\mathsf{P})$. We will describe all conditions that we impose on $F$ in the end of this section. At this point we need only the following consequence of those conditions: for every $\omega\in\Omega$, and every $n\in{\mathbb Z}$, the function $F_{n,\omega}(\cdot)$ is measurable and satisfies \begin{equation} \label{eq:forcing-averages-to-0} \lim_{|x|\to\infty} \frac{F_{n,\omega}(x)}{|x|}=0. \end{equation} \bigskip Let us now define the Burgers dynamics with kick forcing. First, we remind that if $U$ is a $C^3$ function solving the unforced ($F\equiv 0$) equation~\eqref{eq:HJB}, then, differentiating~\eqref{eq:HJB}, we obtain that $u=U_x$ solves~\eqref{eq:unforced-Burgers}. Furthermore, we can introduce a new variable $\varphi$ via the Hopf--Cole transformation: \begin{equation*} U(t,x)=- 2 \varkappa \ln \varphi(t,x), \end{equation*} or, equivalently, \begin{equation*} \varphi(t,x)=e^{- \frac{U(t,x)}{2\varkappa}}, \end{equation*} and directly check that if $\varphi$ is a $C^3$ positive solution of the heat equation \begin{equation} \label{eq:HE} \partial_t \varphi =\varkappa \partial_{xx} \varphi, \end{equation} then the function $u$ obtained from $\varphi$ by \begin{equation} \label{eq:Hopf-Cole-1} u(t,x)=\partial_x U(t,x)=- 2\varkappa \partial_x \ln \varphi(t,x) = - 2\varkappa\frac{\partial_x \varphi(t,x)}{\varphi(t,x)} \end{equation} is a solution of~\eqref{eq:unforced-Burgers}. The solution for the Cauchy problem for~\eqref{eq:HE} under very broad assumptions (measurability and moderate growth of the initial condition) is given by \begin{equation} \label{eq:heat-equation-solution} \varphi(t,x)=\int_{\mathbb R} \varphi(s,y) g_{2\varkappa(t-s) }(x-y)dy,\quad x\in{\mathbb R},\ t>s, \end{equation} where \begin{equation*} g_{D}(x)=\frac{1}{\sqrt{2\pi D}}e^{-\frac{x^2}{2D}}, \quad x\in{\mathbb R}, \end{equation*} is the centered Gaussian density corresponding to variance $D>0$. So, the evolution between the kicks is governed by \eqref{eq:Hopf-Cole-1} and \eqref{eq:heat-equation-solution}. To see what happens at the kick, we rewrite~\eqref{eq:effect-of-one-kick} at the level of potentials as \begin{equation*} U(n,x)=U(n-0,x)+F_{n}(x),\quad x\in {\mathbb R}, \end{equation*} and in the multiplicative form as \begin{equation*} \varphi(n,x)=\varphi(n-0,x)e^{-\frac{F_{n}(x)}{2\varkappa}}. \end{equation*} Combining the instantaneous kick at a time $n$ and unforced evolution on the time interval $[n,n+1)$ leads to the following random linear operator: \begin{align*} \Xi^{n,n+1}_\omega \varphi (y)&=\int_{{\mathbb R}} g_{2\varkappa}(y-x)e^{-\frac{F_{n,\omega}(x)}{2\varkappa}}\varphi(x)dx,\quad y\in{\mathbb R}. \end{align*} Iterating this, we obtain that the (linear) solution operator over an interval $[m,n)$ (producing the solution just before the next instantaneous kick is applied at time~$n$) is given by \begin{equation} \label{eq:F-K} \Xi^{m,n}_\omega \varphi (y)=\int_{{\mathbb R}} Z^{m,n}_{x,y,\omega}\varphi(x)dx,\quad y\in{\mathbb R}, \end{equation} where \begin{equation} \label{eq:Z} Z^{m,n}_{x,y,\omega}=\int_{{\mathbb R}}\dots\int_{\mathbb R} \prod_{k=m}^{n-1} \left[g_{2\varkappa}(x_{k+1}-x_k)e^{-\frac{F_k(x_k)}{2\varkappa}}\right] \delta_x(dx_m) dx_{m+1}\ldots dx_{n-1}\delta_y(dx_n). \end{equation} We can interpret~\eqref{eq:F-K} as a time-discrete version of the Feynman--Kac formula for the following kicked parabolic model \begin{equation*} \partial_t \varphi(t,x)=\varkappa \partial_{xx} \varphi(t,x)+ \varphi(t,x)\cdot\sum_{n\in{\mathbb Z}} (e^{-\frac{F_n(x)}{2\varkappa}}-1)\delta_n(t). \end{equation*} Invoking the change of variables~\eqref{eq:Hopf-Cole-1}, we can define evolution on potentials by \[ \Phi_\omega^{m,n} U = -2\varkappa\ln \Xi_\omega^{m,n} e^{- \frac{U}{2\varkappa}}. \] Let us now be more precise about the domains of the operators we have introduced. The $C^3$~requirement that we started with is superfluous since our Feynman--Kac formula makes sense and produces generalized solutions under much weaker requirements. The space of velocity potentials that we will consider will be~${\mathbb H}$, the space of all locally Lipschitz functions $W:{\mathbb R}\to{\mathbb R}$ satisfying \begin{align*} \liminf_{x\to\pm\infty}\frac{W(x)}{|x|}&>-\infty. \end{align*} We will also need a family of spaces \[ {\mathbb H}(v_-,v_+)=\left\{W\in{\mathbb H}:\ \lim_{x\to \pm\infty} \frac{W(x)}{x}=v_\pm \right\},\quad v_-,v_+\in{\mathbb R}. \] \begin{lemma} \label{lem:invariant_spaces} For any $\omega\in\Omega$, for any $l,n,m\in{\mathbb Z}$ with $l<n<m$ and $W\in{\mathbb H}$, \begin{enumerate} \item\label{it:inv-of-HH} $\Phi^{n,m}_\omega W$ is well-defined and belongs to ${\mathbb H}$; \item\label{it:inv-of-HH-vv} if $W\in{\mathbb H}(v_-,v_+)$ for some $v_-,v_+$, then $\Phi^{n,m}_\omega W\in{\mathbb H}(v_-,v_+)$; \item\label{it:cocycle} $\Phi^{l,m}_\omega W=\Phi^{n,m}_\omega\Phi^{l,n}_\omega W$. \end{enumerate} \end{lemma} We give a proof of this lemma in Section~\ref{sec:aux}. Due to the last statement of the lemma (called the cocycle property) we can view the Burgers evolution as a random dynamical system (see, e.g.,~\cite[Section~1.1]{Arnold:MR1723992}). Potentials are naturally defined up to an additive constant. It is thus convenient to work with $\hat {\mathbb H}$, the space of equivalence classes of potentials from ${\mathbb H}$. We can also introduce spaces $\hat{\mathbb H}(v_-,v_+)$ as classes of potentials in ${\mathbb H}(v_-,v_+)$ coinciding up to an additive constant. The cocycle $\Phi$ can be projected on $\hat{\mathbb H}$ in a natural way. We denote the resulting cocycle on $\hat{\mathbb H}$ by~$\hat\Phi$. We can also introduce the Burgers dynamics on the space ${\mathbb H}'$ of velocities $w$ (actually, classes of equivalence of functions since we do not distinguish two functions coinciding almost everywhere) such that for some function $W\in{\mathbb H}$ and Lebesgue almost every~$x$, $w(x)=W'(x)=\partial_x W(x)$. For all $v_-,v_+\in{\mathbb R}$, ${\mathbb H}'(v_-,v_+)$ is the space of velocity profile with well-defined one-sided averages $v_-$ and $v_+$, it consists of functions $w$ such that the potential~$W$ defined by $W(x)=\int_0^x w(y)dy$ belongs to ${\mathbb H}(v_-,v_+)$. We will write $ w_1=\Psi^{n_0,n_1}_\omega w_0$ if $w_0=W'_0$, $w_1=W'_1$, and $W_1=\Phi^{n_0,n_1}_\omega W_0$ for some $W_0,W_1\in{\mathbb H}$. Of course, the maps belonging to the Burgers cocycle $(\Psi^{n_0,n_1})$ map ${\mathbb H}'$ into itself. The spaces ${\mathbb H}'(v_-,v_+)$ are also invariant. \smallskip Our choice of solution spaces follows that of~\cite{kickb:bakhtin2016} where discontinuous solutions in ${\mathbb H}'$ naturally arose in the zero viscosity setting. In our setting, in fact, for any $w\in {\mathbb H}'$, $\omega\in\Omega$, and $m,n\in{\mathbb Z}$ satisfying $m<n$, $\Psi^{m,n}w$ can be viewed as a function in $C^\infty({\mathbb R})$ as straightforward differentiation in~\eqref{eq:F-K} shows. In fact, the Burgers dynamics has the following additional regularizing property: \begin{lemma} \label{lem:monotonicity_of_x-ux} For any $w\in {\mathbb H}'$, $\omega\in\Omega$, and $m,n\in{\mathbb Z}$ satisfying $m<n$, the function $x\mapsto x-\Psi^{m,n}w(x)$ is nondecreasing. \end{lemma} We prove this lemma in Section \ref{sec:aux} using the results on monotonicity that we introduce in Sections~\ref{sec:monotonicity} and~\ref{sec:uniqueness-of-global-solutions}. \subsection{The requirements on the random forcing} \label{sec:assumption-on-the-potential} For simplicity, we will work on the canonical probability space $(\Omega_0,\mathcal{F}_0,\mathsf{P}_0)$ of realizations of the potential, although other more general settings are also possible. We assume that $\Omega_0$ is the space of continuous functions $F:{\mathbb R}\times\Z\to{\mathbb R}$ equipped with $\mathcal{F}_0$, the completion of the Borel $\sigma$-algebra with respect to local uniform topology, and $\mathsf{P}_0$ is a probability measure preserved by the group of shifts $(\theta^{n,x})_{(n,x)\in\Z\times{\mathbb R}}$ defined by \[ (\theta^{n,x}F)_m(y) = F_{n+m}(x+y),\quad (n,x),(m,y)\in\Z\times{\mathbb R}, \] i.e., $(F_n(x))_{(n,x)\in\Z\times{\mathbb R}}$ is a space-time stationary process. In this framework, $F=F_\omega=\omega$, and we will use all these notations intermittently. In addition to this, we introduce the following requirements: \begin{description} \item[{(A1)}\label{item:stationary-in-space}] The flow $(\theta^{0,x})_{x\in{\mathbb R}}$ is ergodic. In particular, for every $n\in{\mathbb Z}$, $F_n(\cdot)$ is ergodic with respect to the spatial shifts. \item[{(A2)}\label{item:indep-in-time}] The sequence of processes $\big( F_n(\cdot)\big)_{n \in {\mathbb Z}}$ is i.i.d.\ \item[(A3)\label{item:smooth}] With probability 1, for all $n\in{\mathbb Z}$, $F_n(\cdot)\in C^1({\mathbb R})$. \item[(A4)\label{item:exponential-moment-with-viscosity}] For all $(n,x) \in {\mathbb Z}\times{\mathbb R}$, \begin{equation*} \lambda:=\mathsf{E} e^{- \frac{1}{2 \varkappa} F_n(x)} < \infty. \end{equation*} \item[(A5)\label{item:exponential-moment-for-maximum}] There is $\eta>0$ such that for all $(n,j) \in {\mathbb Z} \times {\mathbb Z}$, \begin{equation*} \mathsf{E} e^{\eta F^{*}_{n}(j)} < \infty, \end{equation*} where $F^{*}_{n}(j) = \sup \{ |F_{n}(x)| : x \in [j,j+1] \}$. \end{description} We will use these standing assumptions throughout the paper. However, many of our results will hold true if one removes~\ref{item:smooth} because~\ref{item:exponential-moment-for-maximum} guarantees that $F$ is locally bounded which is sufficient for most of our results. Of course, differentiability of $F$ guarantees that $f=\partial_x F$ in the Burgers equation is defined as a function, but even this is not necessary for some of our claims on the Burgers equation. A~sufficient condition on distributional properties of~$F$ at any fixed time, say, time 0, for existence of an appropriate probability space satisfying~\ref{item:stationary-in-space} and \ref{item:indep-in-time} is mixing of~$F_0$ with respect to spatial shifts. This (along the other requirements from the list above) holds, for example, for Gaussian processes with decaying correlations and processes with finite dependence range. Also, processes obtained from Poissonian noise (or any other space-time ergodic process) via spatial smoothening are compatible with probability spaces satisfying~\ref{item:stationary-in-space}--\ref{item:indep-in-time}. So, the conditions that we impose define a very broad class of processes. We note that the entire inviscid Burgers equation program developed in~\cite{kickb:bakhtin2016} using shot-noise potential, also holds for this broad class of potentials. Stationarity and~\ref{item:exponential-moment-for-maximum} imply that~\eqref{eq:forcing-averages-to-0} holds with probability~$1$ on $\Omega_0$. It will be convenient in this paper to work on a modified probability space \begin{equation} \label{eq:def-of-Omega} \Omega=\left\{F\in\Omega_0: \lim_{|x|\to\infty}\frac{F_{n}(x)}{|x|}=0,\quad n\in{\mathbb Z}\right\}\in\mathcal{F}_0. \end{equation} of probability $1$ instead of $\Omega_0$. As we will see, on this set, the Burgers evolution possesses some nice properties. Moreover, $\Omega$ is invariant under space-time shifts~$\theta^{n,x}$ and under Galilean space-time shear transformations~$L^{v},$ $v\in{\mathbb R}$, defined by \begin{equation} \label{eq:def-of-shear} (L^v F)_n (x)=F_n(x+vn),\quad (n,x)\in\Z\times{\mathbb R}. \end{equation} We denote the restrictions of $\mathcal{F}_0$ and $\mathsf{P}_0$ onto $\Omega$ by $\mathcal{F}$ and $\mathsf{P}$. From now on we work with the probability space $(\Omega,\mathcal{F},\mathsf{P})$. Under this modification, all the distributional properties are preserved. Having introduced the shifts $\theta^{n,x}$, we can also rewrite the cocycle property as \[ \Phi^{n+m}_{\omega}W=\Phi^{m}_{\theta^{n}\omega} \Phi^{n}_\omega W,\quad n,m\in{\mathbb N},\quad\omega\in\Omega, \] where $\theta^{n}=\theta^{n,0}$ and $\Phi^n_\omega=\Phi^{0,n}_\omega$. The cocycle property of $\Psi$ and $\Xi$ can also be expressed similarly. \section{Main results}\label{sec:main_results} Our main results for the positive viscosity Burgers equation are parallel to those of~\cite{kickb:bakhtin2016} for the inviscid case. We say that $u(n,x)=u_\omega(n,x)$, $(n,x)\in{\mathbb Z}\times{\mathbb R}$ is a global solution for the cocycle~$\Psi$ if there is a set $\Omega'\in\mathcal{F}$ with $\mathsf{P}(\Omega')=1$ such that for all $\omega\in\Omega'$, all $m$ and $n$ with $m<n$, we have $\Psi^{m,n}_\omega u_\omega(m,\cdot)= u_\omega(n,\cdot)$. A function $u_\omega(x)$, $\omega\in\Omega$, $x\in{\mathbb R}$ is called skew-invariant if there is a set $\Omega'\in\mathcal{F}$ with $\mathsf{P}(\Omega')=1$ such that for any $n\in{\mathbb Z}$, $\theta^n\Omega'=\Omega'$, and for any $n\in{\mathbb N}$ and $\omega\in\Omega'$, $\Psi^n_\omega u_\omega =u_{\theta^n\omega}$. If $u_\omega(x)$ is a skew-invariant function, then $u_\omega(n,x)=u_{\theta^n\omega}(x)$ is a stationary global solution. One can naturally view the potentials of $u_\omega(x)$ and $u_\omega(n,x)$ as a skew-invariant function and global solution for the cocycle $\hat\Phi$. To state our first result, a description of stationary global solutions, we need more notation. For a subset $A$ of ${\mathbb Z}\times{\mathbb R}$, we denote by $\mathcal{F}_A$ the $\sigma$-sub-algebra of $\mathcal{F}$ generated by random variables $(F_{n}(x))_{(n,x)\in A}$. \begin{theorem}\label{thm:global_solutions} For every $v\in{\mathbb R}$ there is a unique skew-invariant function $u_v:\Omega\to{\mathbb H}'$ such that for almost every $\omega\in\Omega$, $u_{v,\omega}\in {\mathbb H}'(v,v)$. The process $u_{v,\omega}(n,\cdot)=u_{v,\theta^n\omega}(\cdot)$ is a unique stationary global solution in ${\mathbb H}'(v,v)$. The potential $U_{v,\omega}$ defined by $U_{v,\omega}(x)=\int^x u_{v,\omega}(y)dy$ is a unique skew-invariant function for $\hat\Phi$ in~$\hat{\mathbb H}(v,v)$. It defines a unique stationary global solution $U_{v,\omega}(n,\cdot)=U_{v,\theta^n\omega}(\cdot)$ for $\hat\Phi$ in $\hat{\mathbb H}(v,v)$. The skew-invariant functions $U_{v,\omega}$ and $u_{v,\omega}$ are measurable w.r.t.\ $\mathcal{F}|_{(-{\mathbb N})\times{\mathbb R}}$, i.e., they depend only on the history of the forcing. The spatial random process $(u_{v,\omega}(x))_{x\in{\mathbb R}}$ is stationary and ergodic with respect to space shifts. \end{theorem} \begin{remark}\rm All uniqueness statements in this theorem are understood {\it up to zero-measure modifications}. We say that a process $u$ is a unique (up to a zero-measure modification) process with certain properties if for every process $\tilde u$ defined on the same probability space and possessing these properties, $u$ and $\tilde u$ coincide with probability~$1$. \end{remark} This theorem can be interpreted as a 1F1S Principle: for any velocity value $v$, the solution at time $0$ with mean velocity $v$ is uniquely determined by the history of the forcing: $u_{v,\omega}\stackrel{\rm a.s.}{=}\chi_v(F|_{(-{\mathbb N})\times{\mathbb R} })$ for some deterministic functional $\chi_v$ of the forcing in the past, i.e., in $(-{\mathbb N})\times{\mathbb R}$. We actually describe $\chi_v$ in the proof, it is constructed via infinite volume polymer measures on one-sided paths. Since the forcing is stationary in time, we obtain that $u_{v,\theta^n\omega}$ is a stationary process in $n$, and the distribution of $u_{v,\omega}$ is an invariant distribution for the corresponding Markov semi-group, concentrated on ${\mathbb H}'(v,v)$. \medskip The next result shows that each of the global solutions constructed in Theorem~\ref{thm:global_solutions} plays the role of a one-point pullback attractor. To describe the domains of attraction we need to introduce several assumptions on initial potentials $W\in{\mathbb H}$. Namely, we will assume that there is $v\in{\mathbb R}$ such that $W$ and $v$ satisfy one of the following sets of conditions: \begin{align} v&=0,\notag\\ \liminf_{x\to+\infty} \frac{W(x)}{x}&\ge 0, \label{eq:no_flux_from_infinity}\\ \limsup_{x\to-\infty} \frac{W(x)}{x}&\le 0,\notag \end{align} or \begin{align} v&> 0,\notag \\ \lim_{x\to-\infty} \frac{W(x)}{x}&= v,\label{eq:flux_from_the_left_wins}\\ \liminf_{x\to+\infty} \frac{W(x)}{x}&> -v,\notag \end{align} or \begin{align} v&< 0,\notag\\ \lim_{x\to+\infty} \frac{W(x)}{x}&= v,\label{eq:flux_from_the_right_wins}\\ \limsup_{x\to-\infty} \frac{W(x)}{x}&< -v.\notag \end{align} Condition~\eqref{eq:no_flux_from_infinity} means that there is no macroscopic flux of particles from infinity toward the origin for the initial velocity profile $W'$. In particular, any $W\in{\mathbb H}(0,0)$ or any $W\in{\mathbb H}(v_-,v_+)$ with $v_-\le 0$ and $v_+\ge 0$ satisfies~\eqref{eq:no_flux_from_infinity}. If, additionally, $v_+>0$ and $v_-<0$, then it is natural to call this situation a rarefaction fan. We will see that in this case the long-term behavior is described by the global solution $u_0$ with mean velocity $v=0$. Condition~\eqref{eq:flux_from_the_left_wins} means that the initial velocity profile $W'$ creates an influx of particles from $-\infty$ with effective velocity $v\ge 0$, and the influence of the particles at $+\infty$ is not as strong. In particular, any $W\in{\mathbb H}(v,v_+)$ with $v\ge 0$ and $v_+> -v$ (e.g., $v_+=v$) satisfies \eqref{eq:flux_from_the_left_wins}. We will see that in this case the long-term behavior is described by the global solution $u_v$. Condition~\eqref{eq:flux_from_the_right_wins} describes a situation symmetric to~\eqref{eq:flux_from_the_left_wins}, where in the long run the system is dominated by the flux of particles from $+\infty$. The following precise statement supplements Theorem~\ref{thm:global_solutions} and describes the basins of attraction of the global solutions $u_v$ in terms of conditions~\eqref{eq:no_flux_from_infinity}--\eqref{eq:flux_from_the_right_wins}. \begin{theorem}\label{thm:pullback_attraction} For every $v\in{\mathbb R}$, there is a set $\hat \Omega\in\mathcal{F}$ with $\mathsf{P}(\hat \Omega)=1$ such that if $\omega\in\hat \Omega$, $W\in {\mathbb H}$, and one of conditions~\eqref{eq:no_flux_from_infinity},\eqref{eq:flux_from_the_left_wins},\eqref{eq:flux_from_the_right_wins} holds, then $w=W'$ belongs to the domain of pullback attraction of $u_v$: for any $n\in{\mathbb R}$ and any $x\in{\mathbb R}$, \[ \lim_{m\to-\infty} \Psi^{m,n}_\omega w(x) = u_{v,\omega}(n,x), \] and the convergence is uniform on compact sets. \end{theorem} The last statement of the theorem implies that for every $v\in{\mathbb R}$, the invariant measure on ${\mathbb H}'(v,v)$ described after Theorem~\ref{thm:global_solutions} is unique and for any initial condition $w=W'\in{\mathbb H}'$ satisfying one of conditions~\eqref{eq:no_flux_from_infinity},\eqref{eq:flux_from_the_left_wins}, and~\eqref{eq:flux_from_the_right_wins}, the distribution of the random velocity profile at time $n$ weakly converges to the unique stationary distribution on ${\mathbb H}'(v,v)$ as $n\to\infty$, in the local uniform topology. However, our approach does not produce any convergence rate estimates. We also note that, due to Lemma~\ref{lem:monotonicity_of_x-ux}, proving uniform convergence in this theorem amounts to proving pointwise convergence. \medskip We prove Theorems~\ref{thm:global_solutions} and~\ref{thm:pullback_attraction} in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions} using thermodynamic limits for directed polymers that we introduce next. \section{Directed polymers}\label{sec:polymers} Our analysis of the Burgers equation is based on several new results concerning directed polymers and their thermodynamic limits. We provide the necessary background and state those results in this section. Directed polymers in random environment are a class of random media models given by random Boltzmann--Gibbs distributions on paths with (i) free measure describing classical random walks and (ii) the energy function given by the potential accumulated from the random environment by the random walk. In the Burgers equation context, the directed polymers emerge naturally through the Feynman--Kac formula~\eqref{eq:F-K}. It can be understood as integration over the space of paths endowed with appropriate polymer measures. Namely, for any $m,n\in{\mathbb Z}$ with $m<n$ and any $x,y\in{\mathbb R}$, we can introduce the point-to-point polymer measure $\mu^{m,n}_{x,y,\omega}$ via its density \begin{align*} p^{m,n}_{x,y,\omega}(x_m,\ldots,x_n) &=\frac{\prod_{k=m}^{n-1} \left[g_{2\varkappa}(x_{k+1}-x_k)e^{-\frac{F_k(x_k)}{2\varkappa}}\right] }{Z^{m,n}_{x,y,\omega}}, \end{align*} with respect to $\delta_x\times \mathrm{Leb}^{n-m-1} \times \delta_y$, where $\mathrm{Leb}$ is the Lebesgue measure on ${\mathbb R}$. The random normalizing quantity $Z^{m,n}_{x,y,\omega}$ in this formula has been defined in~\eqref{eq:Z}. It is called the point-to-point partition function. The random measure $\mu^{m,n}_{x,y,\omega}$ can be viewed as the Gibbs measure obtained from the Gaussian random walk connecting $x$ to $y$ over time interval~$[m,n]_{\mathbb Z}=[m,n]\cap{\mathbb Z}$, by reweighting paths using the potential energy function \begin{equation} \label{eq:path-energy} H_\omega^{m,n}(x_m,\ldots,x_n)=\sum_{k=m}^{n-1}F_{k,\omega}(x_k), \quad (x_m,\ldots, x_n)\in {\mathbb R}^{n-m+1}, \end{equation} and temperature $2\varkappa$. Note the asymmetry in the definition of $H_\omega^{m,n}$: we have to include $k=m$, but exclude $k=n$. All our results on polymer measures are proved for this choice of path energy, but it is straightforward to obtain their counterparts for the version of energy where $k=m$ is excluded and $k=n$ is included. In our notation for partition functions, we omit the dependence on~$\varkappa$. In fact, in the remaining part of the paper, we will always assume $\varkappa=1/2$ to make the notation lighter. Adaptation of all the statements and proofs to the general $\varkappa>0$ is straightforward. We also often omit the $\omega$ argument in all the notations used above and often write $Z^{m,n}(x,y)$ for $Z^{m,n}_{x,y}$. The polymer density can also be expressed as \[ p^{m,n}_{x,y,\omega}(x_m,\ldots,x_n)= \frac{(2\pi)^{-(n-m)/2}e^{-A_\omega^{m,n}(x_m,\ldots,x_n)}}{Z^{m,n}_{x,y,\omega}}, \] where the total action \begin{equation} \label{eq:action} A^{m,n}_\omega(x_m,\ldots,x_n)=H_\omega^{m,n}(x_m,\ldots,x_n)+I^{m,n}(x_m,\ldots,x_n) \end{equation} is defined via the potential energy introduced in~\eqref{eq:path-energy} and the kinetic energy: \[ I^{m,n}(x_m,\ldots,x_n)=\frac{1}{2}\sum_{k=m}^{n-1}(x_{k+1}-x_k)^2. \] Asymptotic properties of directed polymer models similar to ours have been extensively studied in the literature, see, e.g., surveys \cite{Comets-Sh-Y-survey:MR2073332}, \cite{Hollander:MR2504175}, \cite{Giacomin:MR2380992}. Here, we will mention only results most tightly related to ours. One of our first results is the existence of the infinite volume quenched density of the free energy: \begin{theorem} \label{thm:shape-function} There is a constant $\alpha_0\in{\mathbb R}$ such that for any $v\in{\mathbb R}$, \[ \frac{\ln Z^{0,n}_{0,nv}}{n}\ {\stackrel{\mathrm{a.s.}}{\longrightarrow}}\ \alpha_0-\frac{v^2}{2}, \quad {n\to\infty}. \] \end{theorem} We prove it in Section~\ref{sec:shape-function} after providing some useful properties of partition functions in Section~\ref{sec:properties-of-partition-function}. The quadratic term $-\frac{v^2}{2}$ comes from the shear-invariance symmetry (see~(\ref{eq:def-of-shear})) of our model. The existence of infinite volume normalized quenched free energy has been obtained for a variety of polymer models using subadditivity arguments, see~\cite{Carmona-Hu:MR1939654}, \cite{Comets-Shiga-Yoshida:MR1996276}, \cite{Vargas:MR2299713}, \cite{Comets-Fukushima-Nakajima-Yoshida:MR3406700}, for lattice polymers under various assumptions and \cite{Comets-Yoshida:MR2117626}, \cite{Comets-Cranston:MR3038513}, \cite{Comets-Yoshida:MR3085670} for some continuous models. Variational characterizations of the free energy in terms of auxiliary skew-invariant functions (cocycles) were developed in \cite{Yilmaz-2009:MR2531552}, \cite{Rassoul-Agha--Seppalainen:MR3176363}, \cite{Rassoul-Agha--Seppalainen--Yilmaz:MR2999296}, \cite{Rassoul-Agha--Seppalainen--Yilmaz:2013arXiv1311.3016G}, \cite{Rassoul-Agha--Seppalainen--Yilmaz:2014arXiv1410.4474R}. Concentration inequalities for the finite volume free energy have been obtained for several polymer models, see, e.g., \cite{Mejane:MR2060455}, \cite{Carmona-Hu-2004:MR2073415}, \cite{Rovira-Tindel:MR2129770}. These estimates also imply that the fluctuation of the quenched energy for polymer of length $n$ are (roughly) bounded by $n^{1/2}$ and the typical transversal fluctuations for polymer paths themselves in those settings are smaller than (roughly) $n^{3/4}$, although it is believed that for a large class of models including ours (KPZ universality class, see, e.g.,~\cite{Corwin:MR2930377}), the true scalings are $n^{1/3}$ and $n^{2/3}$, respectively. We prove new concentration inequalities for our model. Their statements, related results, and proofs are given in Section~\ref{sec:concentration}. In the proof of concentration inequalities and throughout the rest of the paper we heavily use the 1-dimensional character of the model. More specifically, in Section~\ref{sec:monotonicity} we prove monotone dependence of point-to-point polymer measures on the endpoints, along with several other monotonicity results. In Section~\ref{sec:delta-straightness-and-tightness}, we use the concentration inequalities obtained in Section~\ref{sec:concentration} to derive so-called $\delta$-straightness which in turn is then used to prove tightness of sequences of polymer measures. In Section~\ref{sec:infinite-volume-polymer-measure}, we study the thermodynamic limit for polymer measures. Namely, we use tightness to construct infinite volume polymer measures as limits of finite volume ones and prove a uniqueness statement. These infinite volume polymer measures will be used in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions} to construct global solutions of the Burgers equation and prove their uniqueness and pullback attraction property. The notion of $\delta$-straightness goes back to~ \cite{Ne}. It can be derived from the concentration of finite volume free energy and the uniform curvature assumption on the shape function, introduced in~\cite{Ne}. It was later used in~\cite{Licea1996}, \cite{HoNe}, \cite{Wu}, \cite{ferrari2005}, \cite{CaPi}, \cite{BCK:MR3110798} and~\cite{kickb:bakhtin2016} in the context of optimal paths in random environments. In these papers, either the curvature assumption was assumed (as in \cite{Licea1996}) or the shape functions were explicitly known so that the curvature assumption was satisfied. To state the thermodynamic limit results precisely, we need some notation. For every $m,n\in{\mathbb Z}$ satisfying $m<n$, we denote the set of all paths $\gamma:\{m,m+1,\ldots,n\}\to{\mathbb R}$ by $S^{m,n}$. If in addition a point $x\in{\mathbb R}$ is given, then $S_{x,*}^{m,n}$ denotes the set of all such paths that satisfy $\gamma_m=x$. If $n=\infty$, then we understand the above spaces as the spaces of one-sided semi-infinite paths. If points $x,y\in{\mathbb R}$ are given, then $S_{x,y}^{m,n}$ denotes the set of all such paths that satisfy $\gamma_m=x$ and $\gamma_{n}=y$. The random point-to-point polymer measure $\mu_{x,y}^{m,n}=\mu_{x,y,\omega}^{m,n}$ is concentrated on $S_{x,y}^{m,n}$ for all~$\omega\in\Omega$. We call a measure $\mu$ on $S^{m,n}_{x,*}$ a polymer measure if there is a probability measure~$\nu$ on~${\mathbb R}$ such that $\mu=\mu_{x,\nu}^{m,n}$, where \[ \mu_{x,\nu}^{m,n}=\int_{{\mathbb R}}\mu_{x,y}^{m,n}\nu(dy). \] We call $\nu$ the terminal measure for~$\mu=\mu_{x,\nu}^{m,n}$. It is also natural to call $\mu$ a point-to-measure polymer measure associated to~$x$ and~$\nu$. Later we will also introduce and study a related notion of point-to-line polymer measure. For every $(m,x)\in\Z\times{\mathbb R}$, we denote by $S_{x}^{m,+\infty}$ the set of paths $y:\{m,m+1,m+2,\ldots\}\to{\mathbb R}$ such that $y_{m}=x$. A measure $\mu$ on $S_{x}^{m,+\infty}$ is called an infinite volume polymer measure if for any $n\ge m$ the projection of $\mu$ on $S^{m,n}_{x,*}$ is a polymer measure. This condition is equivalent to the Dobrushin--Lanford--Ruelle (DLR) condition on the measure $\mu$. We will need the notions of the law of large numbers (LLN) and the strong law of large numbers (SLLN) for polymer measures. We begin with SLLN. For every $(m,x)\in\Z\times{\mathbb R}$ and every $v\in{\mathbb R}$, we denote \[ S_x^{m,+\infty}(v)=\left\{\gamma\in S_x^{m,+\infty}:\ \lim_{n\to\infty}\frac{\gamma_n}{n}=v\right\} \] and say that the strong law of large numbers (SLLN) with slope $v\in{\mathbb R}$ holds for a measure $\mu$ on $S_x^{m,+\infty}$ if $\mu(S_x^{m,+\infty}(v))=1$. We say that LLN with slope $v\in{\mathbb R}$ holds for a sequence of Borel measures $(\nu_n)$ on ${\mathbb R}$ if for all $\delta>0$, \[ \lim_{n\to\infty} \nu_n([(v-\delta)n,(v+\delta) n])=1. \] Finally, for any $(m,x)\in\Z\times{\mathbb R}$, we say that a measure $\mu$ on $S_x^{m,+\infty}$ satisfies LLN with slope $v$ if the sequence of its marginals $\nu_k(\cdot)=\mu\{\gamma:\ \gamma_k\in\cdot\}$ does. We denote by $\mathcal{P}_x^{m,+\infty}(v)$ the set of all polymer measures on $S_x^{m,+\infty}$ satisfying SLLN with slope $v$. The set of all polymer measures on $S_x^{m,+\infty}$ satisfying LLN with slope $v$ is denoted by $\widetilde \mathcal{P}_x^{m,+\infty}(v)$. These sets are random since they depend on the realization of the environment, but we suppress the dependence on $\omega\in\Omega$ in this notation. \begin{theorem}\label{thm:thermodynamic-limit} Let $v\in{\mathbb R}$. Then there is a set $\Omega_v\in\mathcal{F}$ and such that \begin{enumerate}[1.] \item $\mathsf{P}(\Omega_v)=1$. \item \label{it:existence-uniqueness-of-infinite-volume} For all $\omega\in\Omega_v$ and all $(m,x)\in{\mathbb Z}\times{\mathbb R}$, there is a polymer measure $\mu_{x}^{m,+\infty}(v)$ such that \[\mathcal{P}_x^{m,+\infty}(v)=\widetilde\mathcal{P}_x^{m,+\infty}(v)=\{\mu_{x}^{m,+\infty}(v)\}.\] The finite-dimensional distributions of $\mu_{x}^{m,+\infty}(v)$ are absolutely continuous. \item \label{it:thermodynamic-limit} For all $\omega\in\Omega_v$, all $(m,x)\in{\mathbb Z}\times{\mathbb R}$, and for every sequence of measures $(\nu_n)$ satisfying LLN with slope $v$, finite-dimensional distributions of $\mu_{x,\nu_n}^{m,n}$ converge to $\mu_{x}^{m,+\infty}(v)$ in total variation. \end{enumerate} \end{theorem} In other words, with probability one, there is a unique infinite volume polymer measure with prescribed endpoint and slope. Moreover, this infinite volume measure can be obtained via a thermodynamic limit, i.e., as a limit of finite volume polymer measures. A similar result was obtained in \cite{Georgiou--Rassoul-Agha--Seppalainen--Yilmaz:MR3395462} for a model called log-gamma polymer. Log-gamma polymer describes a random walk in a certain random potential on the lattice ${\mathbb Z}^2$. Compared to that model, the one that we study has several features that make the analysis harder. Namely, in our model, the space is continuous and the increments of the polymer paths are not uniformly bounded. Moreover, our model does not give rise to explicit computations that are possible for the log-gamma polymer, so we have to rely only on estimates. Of course, a very useful feature of our model is that the free energy function is exactly computed in Theorem~\ref{thm:shape-function} (except an unknown additive constant), it is quadratic and thus strongly convex. Note that we prove the thermodynamic limit not just for point-to-point polymers, but also for more general point-to-measure polymers. This can be done for gamma-polymers as well. In~\cite{Georgiou--Rassoul-Agha--Seppalainen--Yilmaz:MR3395462} similar results on point-to-line polymers are established for terminal conditions on the line given by a linear tilt function. Our results on pullback attraction in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions} allow to state a version of such a result in our setting, with more general tilt functions that are required to be only asymptotically linear. Tightly connected to the thermodynamic limit results in~\cite{Georgiou--Rassoul-Agha--Seppalainen--Yilmaz:MR3395462} are results on the limits of ratios of partition functions. Logarithms of these limiting ratios are polymer counterparts of Busemann functions that compare actions of infinite geodesics to each other in zero temperature models such as first passage percolation (FPP), last passage percolation, or zero-viscosity Burgers equation, see~\cite{HoNe}, \cite{CaPi-ptrf}, \cite{BCK:MR3110798},\cite{kickb:bakhtin2016}, \cite{Rassoul-Agha--Seppalainen--Yilmaz:2013arXiv1311.3016G}, \cite{Rassoul-Agha--Seppalainen--Yilmaz:2015arXiv151000859G}, \cite{DamronHanson2014}, \cite{Damron2017} and~\cite{AHD:2015arXiv151103262A}, which is a recent survey on FPP. In~\cite{Rassoul-Agha--Seppalainen--Yilmaz:2013arXiv1311.3016G} and \cite{Rassoul-Agha--Seppalainen--Yilmaz:2014arXiv1410.4474R} a variational approach to ratios of partition functions is described. It should be noted that in~\cite{Rassoul-Agha--Seppalainen--Yilmaz:2015arXiv151000859G} and~\cite{DamronHanson2014}, \cite{Damron2017}, some differentiablity assumptions on the shape function were used to study the semi-infinite geodesic and the Busemann function. We also prove a result on limits of partition function ratios for our model: \begin{theorem} \label{thm:ratio-partition-function-intro} For every $v \in {\mathbb R}$, there is a full measure set $\Omega'_{v}$ such that for all $\omega\in\Omega'_v$, for all $(n_1,x_1),(n_2,x_2) \in {\mathbb Z}\times{\mathbb R}\ $, and for every sequence of numbers $(y_N)$ with $\lim\limits_{N \to \infty} y_N/N = v$, we have \begin{equation*} \lim\limits_{N\to \infty} \frac{Z_{x_1, y_N}^{n_1, N}}{Z_{ x_2, y_N}^{n_2, N}} = G, \end{equation*} where $G=G_{v,\omega}(n_1,x_1,n_2,x_1) \in(0,\infty)$ does not depend on the sequence $(y_N)$. \end{theorem} This theorem along with some related and more general results is proved in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions}. In fact, we also prove convergence of logarithmic derivatives of partition function ratios. This extension is nontrivial but useful in our space-continuous setting since, due to the Hopf--Cole transformation, solutions of the Burgers equation can be expressed through these logarithmic derivatives. In particular, this enables us to obtain in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions} our main 1F1S results on global solutions of the Burgers equation stated in Section~\ref{sec:main_results}. For these results, we need to work with time-reversed polymer measures on paths that are defined down to $-\infty$ in time. The subtlety here is that our results on forward polymer measures are not absolutely equivalent to those for backward polymer measures due to the slight asymmetry in the definition of the energy function~$H^{m,n}_\omega$. However, these results (and their proofs) are straightforward modifications of each other, and we often find it practical to ignore the difference between them. \medskip In Section~\ref{sec:overlap}, we use the result on convergence of partition function ratios to derive a version of hyperbolicity property for the positive temperature polymer case. Namely, we show that the marginals of any two polymer measures with the same slope are asymptotic to each other: \begin{theorem} \label{thm:convergence-total-variation} Let $v \in {\mathbb R}$. On a full measure event $\Omega'_{v}$, for any $(n_1,x_1), (n_2,x_2) \in {\mathbb Z} \times{\mathbb R}$, we have \begin{equation*} \lim_{N \to \infty} \| \mu_{x_1}^{n_1,+\infty}(v) \pi^{-1}_N - \mu_{x_2}^{n_2,+\infty}(v) \pi^{-1}_N \|_{TV} = 0, \end{equation*} where $\mu \pi^{-1}_N$ is the projection of a measure $\mu$ on $N$-th coordinate and $\| \cdot \|_{TV}$ denotes the total variation distance. \end{theorem} In fact, we prove a time-reversed version of this result in Section~\ref{sec:overlap}. Since the marginals $\mu_{x_i}^{n_i,+\infty}(v) \pi^{-1}_N$ define the entire measure $\mu_{x_i}^{n_i,+\infty}(v)$ uniquely due to the Markovian character of nearest neighbor interactions encoded in the action functional, a stronger statement on overlap of measures on paths also follows immediately. \section{Properties of the partition function}\label{sec:properties-of-partition-function} We recall that throughout the paper, the reasoning does not depend on the viscosity value, so from now on we set $\varkappa=1/2$ for brevity. Under this convention, $2\varkappa=1$, so we will often work with Gaussian kernel $g(\cdot)=g_1(\cdot)$. Let us denote $Z^n(y)=Z^{0,n}(0,y)$ for brevity. The following lemma states how distributional properties of partition functions behave under shear transformations of space-time. We write $\stackrel{d}{=}$ to denote identity in distribution. \begin{lemma} For any $m,n\in{\mathbb Z}$ satisfying $m<n$ and any points $x,y\in{\mathbb R}$, \[ Z^{n+l,m+l}(x+\Delta,y+\Delta)\stackrel{d}{=} Z^{n,m}(x,y). \] Also, for any $v\in{\mathbb R}$, \[ Z^n(vn)\stackrel{d}{=} e^{-\frac{v^2}{2}n} Z^{n}(0). \] \end{lemma} \bpf The first statement of the lemma follows from the space-time stationarity of~$F$. For the second claim, let us make a change of variables $x_k=x'_k+kv$ for $k=0,\ldots, n$ in \eqref{eq:Z}, to obtain the following integral ($x_0=0$ and $x_n=vn$ are fixed, i.e., $x'=0$ and $x'_n=0$ are fixed): \begin{align}\notag Z^{n}(vn)&=\int_{{\mathbb R}}\dots\int_{\mathbb R} \prod_{k=0}^{n-1} [g(x'_{k+1}-x'_k+v)e^{-F_k(x'_k+kv)}] dx'_{1}\ldots dx'_{n-1}\\ &\stackrel{d}{=}\int_{{\mathbb R}}\dots\int_{\mathbb R} \prod_{k=0}^{n-1} [g(x'_{k+1}-x'_k+v)e^{-F_k(x'_k)}] dx'_{1}\ldots dx'_{n-1} \label{eq:applying-shear} \end{align} due to the i.i.d.\ property and the spatial stationarity of~$F$. Now notice that \begin{align*} \prod_{k=0}^{n-1} g(x'_{k+1}-x'_k+v) &=\frac{1}{(2\pi)^{n/2}} e^{-\frac{1}{2}\sum_{k=0}^{n-1} (x'_{k+1}-x'_k+v)^2} \\&=\frac{1}{(2\pi)^{n/2}} e^{-\frac{1}{2}\sum_{k=0}^{n-1} (x'_{k+1}-x'_k)^2 -v\sum_{k=0}^{n-1}(x'_{k+1}-x'_k) -\frac{n}{2}v^2} \\&= e^{-\frac{v^2}{2}n}\prod_{k=0}^{n-1} g(x'_{k+1}-x'_k) \end{align*} since \[ \sum_{k=0}^{n-1}(x'_{k+1}-x'_k)=x'_n-x'_0=0. \] Plugging this into~\eqref{eq:applying-shear}, we obtain \[ Z^{n}(vn)\stackrel{d}{=}e^{-\frac{v^2}{2}n} \int_{{\mathbb R}}\dots\int_{\mathbb R} \prod_{k=0}^{n-1} [g(x'_{k+1}-x'_k)e^{-F_k(x'_k)}] dx'_{1}\ldots dx'_{n-1} = e^{-\frac{v^2}{2}n} Z^{n}(0), \] and the proof is completed. {{\hfill $\Box$ \smallskip}} It is easy to extend this proof and obtain a more general lemma on shift-invariance and shear-invariance. To state this result, in addition to the existing notation $\theta^{n,x}$ and $L^v$ for transformations of the probability space, we define transformations of $\Z\times{\mathbb R}$: \[ \theta_{n,x}(m,y)=(m-n,y-x),\quad\quad L_v(m,y)=(m,y-mv). \] Throughout the paper we write $Pf^{-1}$ to denote the pushforward of a measure~$P$ under a map $f$. \begin{lemma} The following holds true: \label{lem:shear-for-Z_v} \begin{enumerate} \item Let $(n,x),(m,y)\in \Z\times{\mathbb R}$. Then, for all $\omega\in \Omega$, \[ \mu^{n,m+n}_{x,y+x,\omega}\theta_{n,x}^{-1}=\mu^{0,m}_{0,y,\theta^{n,x}\omega}. \] \item Let $Z_v(n)=e^{\frac{v^2}{2}n}Z^{n}(vn)$, $n\in{\mathbb N}$, $v\in{\mathbb R}$. Then the distribution of the process $Z_v(\cdot)$ does not depend on $v$. \item For every $n\in{\mathbb N}$, the process $\bar Z_n(x)=e^{\frac{x^2}{2n}}Z^{n}(x)$, $x\in{\mathbb R}$, is stationary in $x$. \item Let $n\in{\mathbb N}$, $v\in{\mathbb R}$. Then, for all $\omega\in\Omega$, \[ \mu^{0,n}_{0,nv,\omega}L_{v}^{-1}=\mu^{0,0}_{0,0,L^v\omega}. \] \end{enumerate} \end{lemma} \begin{lemma}\label{lem:expectation-of-Z} For any $m,n\in{\mathbb Z}$ satisfying $m<n$ and any $x,y\in{\mathbb R}$, \[\mathsf{E} Z^{m,n}(x,y)=\lambda^{n-m}g_{n-m}(y-x),\] where $\lambda=\mathsf{E} e^{-F_0(0)}<\infty$ according to~\ref{item:exponential-moment-with-viscosity}. \end{lemma} \bpf We can use Fubini's theorem and the i.i.d.\ property of $(F_k)$ to write \begin{align*} \mathsf{E} Z^{m,n}(x,y)&=\int_{{\mathbb R}}\dots\int_{\mathbb R} \prod_{k=m}^{n-1} [g(x_{k+1}-x_k)\mathsf{E} e^{-F_k(x_k)}] dx_{m+1}\ldots dx_{n-1}\\&=\lambda^{n-m}g_{n-m}(y-x), \end{align*} where we also used the convolution property of Gaussian densities.{{\hfill $\Box$ \smallskip}} \begin{lemma} \label{lem:smoothness-of-partition-function} Let $m < n$. For all~$\omega$, the point-to-point partition function $Z^{m,n}(x,y)$ is $C^{\infty}$ in $y$ and as smooth as $F_m(x)$ in $x$. Moreover, partial derivatives of $Z^{m,n}(x,y)$ can be obtained by differentiation under the integral in~\eqref{eq:Z}. \end{lemma} \begin{proof} If $n-m=1$, the claim is obvious. If $n -m \ge 2$, it suffices to show that \begin{align*} e^{F_m(x)} Z^{m,n}(x,y) = \int f(x,y,x_{m+1},\dotsc, x_{n-1}) \, dx_{m+1} \dots dx_{n-1} \end{align*} is smooth in $x$ and $y$, where \[ f(x,y,x_{m+1},\dotsc, x_{n-1})=e^{-\frac{1}{2}(x-x_{m+1})^2 - \frac{1}{2}(x_{n-1}-y)^2}\prod_{k=m+1}^{n-1} e^{-\frac{1}{2}(x_{k+1}-x_k)^2- F_k(x_k)}. \] By (\ref{eq:def-of-Omega}), we can find a constant $c$ such that if $m+1 \le k \le n-1$, then $|F_k(z)| \le c(|z|+1)$ for all~$z$. The lemma follows from \begin{align*} &\int \bigg| \frac{\partial^i}{\partial x^i} \frac{\partial^j}{ \partial y^j} f(x,y,x_{m+1}, \dotsc, x_{n-1}) \bigg| \, dx_{m+1} \dots dx_{n-1} \\ \le& \int c_i c_j \big(|x-x_{m+1}|^i + 1\big) \big(|y-x_{n-1}|^j + 1\big) \\ & \cdot e^{-\frac{1}{2}(x-x_{m+1})^2 - \frac{1}{2}(x_{n-1}-y)^2} \prod_{k=m+1}^{n-1} e^{-\frac{1}{2}(x_{k+1}-x_k)^2 + c(|x_k| + 1)} \, dx_{m+1}\dots dx_{n-1} < \infty, \end{align*} where $c_i$ are absolute constants. \end{proof} As a corollary, we have \begin{lemma} \label{lem:smoothness-of-partition-ratio} Let $m,n > k$. Then $Z^{k,m}(x,y) /Z^{k,n}(x,z)$ is $C^{\infty}$ in $x$, $y$ and $z$. \end{lemma} \section{The shape function or free energy density }\label{sec:shape-function} The goal of this section is to study the linear growth of $\ln Z^{m,n}(x,y)$ over long time intervals. It is useful to introduce an auxiliary function \[ Z_*^{m,n}(x,y)=\min_{|\Delta x|,|\Delta y|<1/2}Z^{m,n}(x+\Delta x, y+\Delta y). \] \begin{lemma}\label{lem:super-additive} The process $Z_*$ is super-multiplicative, i.e., \[ Z_*^{n_1,n_3}(x,z)\ge Z_*^{n_1,n_2}(x,y)Z_*^{n_2,n_3}(y,z). \] Equivalently, $\ln Z_*$ is super-additive, i.e., \[ \ln Z_*^{n_1,n_3}(x,z)\ge \ln Z_*^{n_1,n_2}(x,y) + \ln Z_*^{n_2,n_3}(y,z). \] \end{lemma} \bpf Let $|x'-x|,|z'-z|<1/2$. Then \begin{align*} Z^{n_1,n_3}(x',z')&= \int_{y'\in{\mathbb R}} Z^{n_1,n_2}(x',y') Z^{n_2,n_3}(y',z')dy'\\ &\ge \int_{y':|y'-y|<1/2} Z^{n_1,n_2}(x',y') Z^{n_2,n_3}(y',z')dy' \\ &\ge Z^{n_1,n_2}_*(x,y)Z_*^{n_2,n_3}(y,z), \end{align*} and the lemma follows. {{\hfill $\Box$ \smallskip}} \begin{lemma}\label{lem:shape-function-for-aux} For any $v\in{\mathbb R}$, there is $\alpha(v)\in{\mathbb R}$ such that \[ \frac{\ln Z_*^{0,n}(0,nv)}{n}{\stackrel{\mathrm{a.s.}}{\longrightarrow}} \alpha(v),\quad n\to\infty. \] \end{lemma} \begin{proof} Due to Lemma~\ref{lem:super-additive} and Kingman's sub-additive ergodic theorem, it suffices to check that for every $v\in{\mathbb R}$, there is $C(v)>0$ such that \[ \mathsf{E} \ln Z_*^{0,n}(0,nv)< C(v)n,\quad n \in{\mathbb N}. \] This follows from Jensen's inequality and Lemma~\ref{lem:expectation-of-Z}: \[ \mathsf{E} \ln Z^{0,n}_*(0,nv)\le \ln \mathsf{E} Z^{0,n}(0,nv)\le n\ln \lambda -\frac{(nv)^2}{2n}-\frac{1}{2}\ln(2\pi)- \frac{1}{2}\ln n, \] and the proof is completed. \end{proof} \begin{lemma}\label{lem:same-shape-function-for-true-Z} For any $v\in{\mathbb R}$, \[ \frac{\ln Z^{0,n}(0,nv)}{n}{\stackrel{\mathrm{a.s.}}{\longrightarrow}}\alpha(v),\quad n\to\infty, \] where $\alpha(v)$ has been introduced in Lemma~\ref{lem:shape-function-for-aux}. \end{lemma} \begin{proof} Due to Lemma~\ref{lem:shear-for-Z_v}, it is sufficient to prove the lemma for $v=0$. Lemma~\ref{lem:shape-function-for-aux} and the inequality $Z^{0,n}(0,0)\ge Z^{0,n}_*(0,0)$ imply that it suffices to check \begin{equation} \label{eq:discrepancy-between-Z-and-min-Z} \limsup_{n\to\infty}\left(\frac{\ln Z^{0,n}(0,0)}{n}-\frac{\ln Z^{0,n}_*(0,0)}{n}\right)\le 0. \end{equation} For this, we need to see that $Z^{0,n}(0,0)/Z^{0,n}_*(0,0)$ is bounded by a function that grows sub-exponentially in $n$. First we note that there is $q>0$ such that \begin{equation} \liminf_{n\to\infty} \frac{Z^{0,n}_*(0,0)}{q^n}\stackrel{\rm a.s.}{>}0. \label{eq:exp-growth-guaranteed} \end{equation} To see this, it is sufficient to notice that for every $x,y\in[-1/2,1/2]$, \[ Z^{0,n}(x,y)\ge \int_{[-1/2,1/2]}\ldots\int_{[-1/2,1/2]} \bar{g}^n e^{-\sum_{k=0}^{n-1}\bar{F}_{k}}dx_1\ldots dx_{n-1}, \] where \[ \bar{g}=g(1)=\min_{|z_1|,|z_2|<1/2} g(z_1-z_2), \] \[ \bar{F}_{k}=\max_{|z|<1/2} F_{k}(z),\quad k\ge 0, \] and apply the SLLN to the partial sums of i.i.d.\ sequence $(\bar{F}_{k})_{k\ge 0}$. To compare $Z^{0,n}(0,0)$ to $Z^{0,n}_*(0,0)$, let us take $r_n=n^{3/4}$, introduce sets $A_{-1}=A_{-1}(n)=(-\infty,r_n]$, $A_{0}=A_0(n)=[-r_n,r_n]$, $A_{1}=A_1(n)=[r_n,\infty)$, and write \[ Z^{0,n}(0,0)=\sum_{i,j\in\{-1,0,1\}}B_{ij}^n(0,0), \] where \begin{align*} B^n_{ij}(x,y)&=\int_{x_1\in A_i}\int_{x_{n-1}\in A_j} Z^{0,1}(x,x_1)Z^{1,n-1}(x_1,x_{n-1})Z^{n-1,n}(x_{n-1},y)dx_1dx_{n-1}. \end{align*} We need to estimate $B_{ij}^n(0,0)/Z^{0,n}_*(0,0) = B_{ij}^n(0,0)/Z^{0,n}(x_*,y_*)$, where points $x_*$ and~$y_*$ provide minimum in the definition of $Z^{0,n}_*(0,0)$. Let us estimate $B_{11}^n(0,0)$ and $B_{10}^n(0,0)$ first. By the Fubini theorem and the convolution property of Gaussian densities, \[ \mathsf{E} [B_{11}^n(0,0)+B_{10}^n(0,0)]\le \lambda^n \int_{A_1}\int_{A_{1}\cup A_0} g(x_1)g_{n-2}(x_{n-1}-x_1)g(-x_{n-1})\, dx_1\,dx_{n-1}. \] Since $g_{n-2}(z)\le 1$ for all $z\in{\mathbb R}$ and $g$ is a probability density, we conclude that \begin{align*} \mathsf{E} [B_{11}^n(0,0)+B_{10}^n(0,0)] &\le \lambda^n \int_{A_1}\int_{A_0\cup A_{1}} g(x_1)g(-x_{n-1})\, dx_1\,dx_{n-1}\\ &\le \lambda^n \int_{A_1} g(x)\, dx \le \lambda^n \mathsf{P}\{\mathcal{N}>r_n\} \le \lambda^n \frac{1}{(2\pi)^{1/2} r_n } e^{-r_n^2/2}, \end{align*} where $\mathcal{N}$ is a standard Gaussian random variable. So, for any $\rho>0$, \[ \mathsf{P}\{B_{11}^n(0,0)+B_{10}^n(0,0)>\rho^n\}\le \rho^{-n}\mathsf{E} [B_{11}^n(0,0)+B_{10}^n(0,0)]\le \frac{\lambda^n}{\rho^n} \frac{1}{(2\pi)^{1/2} r_n } e^{-r_n^2/2}. \] Here, the last factor decays super-exponentially, and the Borel--Cantelli Lemma implies that for any $\rho>0$, \begin{equation} \label{eq:super-esp-decay} \lim_{n\to\infty}\frac{B_{11}^n(0,0)+B_{01}^n(0,0)}{\rho^n}\stackrel{\rm a.s.}{=}0. \end{equation} Combining \eqref{eq:super-esp-decay} with \eqref{eq:exp-growth-guaranteed} and applying the same reasoning to all terms $B_{ij}^n$ with $|i|+|j|\ne 0$, we obtain \begin{equation} \label{eq:B_11} \lim_{n\to\infty}\frac{\sum_{|i|+|j|\ne 0}B_{ij}^n(0,0)}{Z_*^{0,n}(0,0)}\stackrel{\rm a.s.}{=}0. \end{equation} It remains to estimate $B_{00}(0,0)$: \begin{align}\notag \frac{B_{00}^n(0,0)}{Z^{0,n}(x_*,y_*)} &\le \frac{B_{00}^n(0,0)}{B_{00}^n(x_*,y_*)}\le \max_{x_1,x_{n-1}\in A_0(n)} \frac{Z^{0,1}(0,x_1)Z^{n-1,n}(x_{n-1},0)}{Z^{0,1}(x_*,x_1)Z^{n-1,n}(x_{n-1},y_*)} \\ \notag&\le \max_{x_1,x_{n-1}\in A_0(n)} \frac{g(x_1)e^{-F_0(0)} g(-x_{n-1})e^{-F_{n-1}(x_{n-1})}}{g(x_1-x_*)e^{-F_0(x_*)} g(y_*-x_{n-1})e^{-F_{n-1}(x_{n-1})}} \\ \notag &\le \max_{x_1,x_{n-1}\in A_0(n)}e^{-F_0(0)+F_0(x_*)} e^{(x_*^2+y_*^2)/2} e^{-x_*x_1-y_*x_2} \\&\le C_1(\omega)e^{r_n} \label{eq:random-ratio-of-densities} \end{align} for some random constant $C_1(\omega)$ and all $n\ge 2$. Combining~\eqref{eq:B_11} and \eqref{eq:random-ratio-of-densities}, we obtain~\eqref{eq:discrepancy-between-Z-and-min-Z} and finish the proof of Lemma~\ref{lem:same-shape-function-for-true-Z}.{{\hfill $\Box$ \smallskip}} \begin{theorem} There is a constant $\alpha_0\in{\mathbb R}$ such that for any $v\in{\mathbb R}$, \[ \lim_{n\to\infty} \frac{\ln Z^n(vn)}{n}{\stackrel{\mathrm{a.s.}}{=}}\alpha_0-\frac{v^2}{2}, \] i.e., the shape function $\alpha(\cdot)$ introduced in Lemmas~\ref{lem:shape-function-for-aux} and~\ref{lem:same-shape-function-for-true-Z}, satisfies \[ \alpha(v)=\alpha_0-\frac{v^2}{2}. \] \end{theorem} \bpf The theorem follows directly from Lemmas~\ref{lem:same-shape-function-for-true-Z} and~\ref{lem:shear-for-Z_v}. \end{proof} \section{Monotonicity} \label{sec:monotonicity} The order on the real line plays an important role in our analysis. The goal of this section is to establish monotonicity of polymer measures with respect to endpoints, along with some related results. We begin with an auxiliary lemma on a monotonicity property of the Gaussian kernel. \begin{lemma} \label{lem:abstract-monotonicity} Suppose $\nu$ is a Borel $\sigma$-finite measure such that \[ Z(x)=\int_{{\mathbb R}}g(z-x)\nu(dz) \] is finite for all $x\in{\mathbb R}$, and let \[ G(x,y)=\frac{\int_{(-\infty,y]}g(z-x)\nu(dz)}{Z(x)},\quad x,y\in{\mathbb R}. \] Then $G(x,y)$ is nondecreasing in~$y$. If $\nu\{(y,\infty)\}>0$ and $\nu\{(-\infty,y]\}>0$, then $G(x,y)$ is strictly decreasing in $x$. \end{lemma} \begin{proof} The monotonicity in $y$ is obvious. Due to \[ \frac{1}{G(x,y)}= \frac{\int_{{\mathbb R}}g(z-x)\nu(dz)}{\int_{(-\infty,y]}g(z-x)\nu(dz)}=1+\frac{\int_{(y,\infty)}g(z-x)\nu(dz)}{\int_{(-\infty,y]}g(z-x)\nu(dz)}, \] it remains to prove that for all $z\in(y,\infty)$, \[ H(x,y,z)=\frac{\int_{(-\infty,y]}g(z'-x)\nu(dz')}{g(z-x)} \] decreases in $x$. We rewrite \[ H(x,y,z)=\int_{(-\infty,y]}e^{\frac{-(x-z')^2+(x-z)^2}{2}}\nu(dz')=e^{z^2/2}\int_{(-\infty,y]}e^{x(z'-z)-\frac{z'^2}{2} }\nu(dz'). \] Since $z'-z<0$, the integrand $e^{x(z'-z)-\frac{z'^2}{2} }$ decreases in $x$ and so does the integral on the right-hand side. \end{proof} For any $d\in{\mathbb N}$, we denote by ${\preceq}$ the natural partial order on ${\mathbb R}^d$, i.e., we write $x\preceq y$ iff $x_k\le y_k$ for all $k=1,\ldots,d$. A function $f:{\mathbb R}^d\to {\mathbb R}$ is coordinatewise nondecreasing if $x\preceq y$ implies $f(x)\le f(y)$. For two Borel probability measures $\nu_1,\nu_2$ on ${\mathbb R}^d$, we write $\nu_1\preceq \nu_2$ (and say that $\nu_1$ is {\it stochastically dominated} by $\nu_2$) iff for any bounded coordinatewise nondecreasing function $f:{\mathbb R}^d\to{\mathbb R}$ \[ \int_{{\mathbb R}^d} f(x)\nu_1(dx)\le \int_{{\mathbb R}^d} f(x)\nu_2(dx). \] For $d=1$, $\nu_1\preceq\nu_2$ is equivalent to $\nu_1\{(-\infty,x]\}\ge \nu_2\{(-\infty,x]\}$ for all $x\in{\mathbb R}$. There is also a coupling characterization of stochastic dominance usually called Strassen monotone coupling theorem (see Theorems~7 and~11 in~\cite{Strassen:MR0177430} and a discussion in~\cite{Lindvall:MR1711599}). To state this theorem and our results on stochastic dominance, we introduce notation that will be used in various contexts throughout the paper: we use $\pi_k x$ to denote the $k$-th coordinate of $x$, where $x$ is either a vector or an infinite sequence. We also use $\pi_{m,n}x=(x_{m},\ldots,x_{n})$. \begin{lemma}[Monotone coupling]\label{lem:dominance-by-coupling} Borel measures $\nu_1,\ldots,\nu_n$ on ${\mathbb R}^d$ satisfy $\nu_1\preceq\ldots\preceq \nu_n$ iff there is a measure $\nu$ on $({\mathbb R}^d)^n$ such that $\nu_k$ is the $k$-th marginal of $\nu$, i.e., $\nu_k=\nu \pi_k^{-1}$, $k=1,\ldots,n$, and \[ \nu\{(x^{(1)},\ldots,x^{(n)})\in({\mathbb R}^d)^n:\ x^{(1)}\preceq\ldots\preceq x^{(n)}\}=1. \] \end{lemma} \begin{lemma}\label{lem:dominance-1} Let $x\le x'$. Then for any $m,n$ with $m<n$, any $y\in{\mathbb R}$, and all $\omega$, the polymer measure $\mu_{x,y}^{m,n}$ is stochastically dominated by $\mu_{x',y}^{m,n}$. \end{lemma} \begin{proof} The reasoning does not depend on $m$, so we set $m=0$ for brevity. We prove by induction in $k$ that for all $x < x'$ and for any $k\in (0,n)\cap{\mathbb N}$, there is a measure $\nu_{k}$ on $({\mathbb R}^k)^2$ such that \begin{align*} \nu_k(\cdot \times {\mathbb R}^k) &= \mu_{x,y}^{0,n} \pi_{1,k}^{-1},\\ \nu_k({\mathbb R}^k \times \cdot) &= \mu_{x',y}^{0,n} \pi_{1,k}^{-1}, \end{align*} and \begin{equation} \label{eq:dominance-coupling} \nu_{k} \{ (x, x'): x \preceq x'\} = 1. \end{equation} In particular, taking $k = n-1$ we obtain the conclusion of the lemma. Let us check the case $k=1$ first. \begin{align*} \mu_{x,y}^{0,n}\pi^{-1}_{1} \bigl( (-\infty, r] \bigr) &=\frac{1}{Z_{x,y}^{0,n}}\int_{(-\infty,r]}Z_{x,{s}}^{0,1}Z_{{s},y}^{1,n} d{s} =\frac{\int_{(-\infty,r]}g({s}-x)e^{-F_0(x)}Z_{{s},y}^{1,n} d{s}}{\int_{{\mathbb R}}g({s}-x)e^{-F_0(x)}Z_{{s},y}^{1,n} d{s}} \\ &=\frac{\int_{(-\infty,r]}g({s}-x)Z_{{s},y}^{1,n} d{s}}{\int_{{\mathbb R}}g({s}-x)Z_{{s},y}^{1,n} d{s}}. \end{align*} Introducing $\nu(d{s})=Z_{{s},y}^{1,n} d{s}$, we can apply Lemma~\ref{lem:abstract-monotonicity} to see that $\mu_{x,y}^{0,n} \pi_{1}^{-1}\bigl( (-\infty,r] \bigr)$ is decreasing in $x$. Therefore $\mu_{x,y}^{0,n}\pi_{1}^{-1}\preceq\mu_{x',y}^{0,n}\pi_{1}^{-1}$ for $x < x'$, which finishes the argument for the basis of induction. Suppose for $k \ge 1$ the desired $\nu_k$ have been constructed. We will construct $\nu_{k+1}$ using $\nu_k$. The basis of induction (the claim for 1-dimensional marginals) implies that for any $z,z'\in{\mathbb R}$ satisfying $z \le z'$, there is a measure $\nu_{z,z'}$ on ${\mathbb R} \times {\mathbb R}$ such that \begin{align*} \nu_{z,z'}(\cdot \times {\mathbb R}) &= \mu_{z,y}^{k,n}\pi_{k+1}^{-1}(\cdot),\\ \nu_{z,z'}({\mathbb R} \times \cdot) &= \mu_{z',y}^{k,n}\pi_{k+1}^{-1}(\cdot), \end{align*} and $\nu_{z,z'}\{(w,w'): w \le w' \} = 1$. Then the measure $\nu_{k+1}$ defined by \begin{multline*} \nu_{k+1}\bigl( (A_{1}\times\dotsm \times A_{k+1})\times (A_{1}' \times \dotsm \times A_{k+1}') \bigr) \\ = \int_{x_i \in A_i, x_i' \in A_i',i \le k} \nu_k \bigl( dx_{1}, ..., dx_{k}, dx_{1}', ..., dx_{k}' \bigr) \nu_{x_{k}, x_{k}'} \bigl( A_{k+1}\times A_{k+1}' \bigr) \end{multline*} satisfies (\ref{eq:dominance-coupling}) with $k$ replaced by $k+1$. To see that $\nu_{k+1}$ has correct marginals, it suffices to notice that from the definition of polymer measures, we have \[ \mu_{x,y}^{0,n}(A_{1}\times \dotsm \times A_{n-1}) = \int_{x_i \in A_i, i\le k} \mu_{x,y}^{0,n}\pi_{1,k}^{-1} (dx_{1}, ..., dx_{k}) \mu_{x_{k}, y}^{k,n} ( A_{k+1} \times \dotsm \times A_{n-1}) \] for any $x, y$ and $k \le n-1$. \end{proof} One can also easily obtain a time-reversed version of Lemma~\ref{lem:dominance-1}: \begin{lemma}\label{lem:dominance-2} Let $y\le y'$. Then for any $m,n$ with $m<n$, any $x\in{\mathbb R}$, and all $\omega$, the polymer measure $\mu_{x,y}^{m,n}$ is stochastically dominated by $\mu_{x,y'}^{m,n}$. \end{lemma} We can now state the main result of this section. It easily follows from Lemmas~\ref{lem:dominance-by-coupling},~\ref{lem:dominance-1}, and~\ref{lem:dominance-2}. \begin{lemma}[Main monotonicity lemma]\label{lem:main-monotonicity} The following holds for all $\omega\in\Omega$: \begin{enumerate}[1.] \item Let $x\le x'$ and $y\le y'$. Then for any $m,n$ with $m<n$, the polymer measure $\mu_{x,y}^{m,n}$ is stochastically dominated by $\mu_{x',y'}^{m,n}$. \item If two distributions $\nu_1,\nu_2$ on ${\mathbb R}$ satisfy $\nu_1\preceq\nu_2$, then, for any $x\in{\mathbb R}$ and any $m,n\in{\mathbb Z}$ satisfying $m\le n$, we have $\mu_{x,\nu_1}^{m,n}\preceq \mu_{x,\nu_2}^{m,n}$. \item If $x\le x'$, then for any distribution $\nu$ on ${\mathbb R}$ and any $m,n\in{\mathbb Z}$ satisfying $m\le n$, we have $\mu_{x,\nu}^{m,n}\preceq \mu_{x',\nu}^{m,n}$. \end{enumerate} \end{lemma} \section{Concentration inequality for free energy}\label{sec:concentration} The aim of this section is to prove a concentration inequality of the free energy around its asymptotic value. It is natural to expect that our model belongs to the KPZ universality class. In particular, the fluctuations of $\ln Z^n$ around the mean are expected to be of order~$n^{\chi}$, where~$\chi=1/3$. The main result of this section may be interpreted as~$\chi\le 1/2$. \begin{theorem} \label{thm:concentration-of-free-energy} There are positive constants $c_0, c_1, c_2, c_3$ such that for all $v \in {\mathbb R}$, all $n > c_0$ and all $u \in (c_3n^{1/2} \ln^{3/2}n, n \ln n]$, \begin{equation*} \mathsf{P} \{ |\ln Z^n(0,vn) - \alpha(v)n | > u \} \le c_1 \exp \left\{ -c_2 \frac{u^2}{n \ln^2n} \right\}. \end{equation*} \end{theorem} The right endpoint of the interval of applicability of this inequality is chosen to be $ n \ln n$. In fact, similar inequalities with appropriate adjustments hold for different choices of the right endpoint. Due to the invariance under shear transformations (Lemma~\ref{lem:shear-for-Z_v}), it suffices to prove this theorem for $v = 0$. The proof can be divided into two steps: the first step is Lemma~\ref{lem:free-energy-concentration-around-expectation}, where we obtain a concentration expectation of $\ln Z^n$ around its expectation $\mathsf{E} \ln Z^n$ using Azuma's inequality after certain truncation; the second step is to estimate the speed of convergence, i.e., how far $\mathsf{E} \ln Z^n$ deviates from the linear function $\alpha_0n$, using Lemmas~\ref{lem:doubling-argument-inequality} and~\ref{lem:doubling-argument-lemma}. Such concentration inequalities have been proved for various FPP/LPP and polymer models with different tails. The first such result appeared in~\cite{Kesten:MR1221154} on FPP, with a tail of $e^{-cu/\sqrt{n}}$. Using Talagrand's inequality, this can be improved to~$e^{-cu^2/n}$. In \cite{benjamini2003}, the authors proved that for FPP with edge weight distribution~$\mathsf{P}(w_e = a ) =\mathsf{P}(w_e = b) = 1/2$, the variance of $\ln Z^n$ is $O(\frac{n}{\log n})$, which is sublinear. The result was later strengthened to a concentration inequality with a tail $e^{-cu \sqrt{\ln n/n}}$ for more general distributions, see~\cite{benaim2008} and \cite{damron2014}. In~\cite{Alexander2013}, similar concentration inequality was obtained for a polymer model. Our method in this section is more elementary and will not lead to a sharper subgaussian concentration as mentioned above, but it is sufficient, in conjunction with the convexity of the shape function, to help us to establish straightness estimates in section~\ref{sec:delta-straightness-and-tightness}. Moreover, in a forthcoming paper on the zero-temperature limit of polymer measures, we strengthen our estimates proving a functional concentration inequality on the free energy process indexed by temperature (or viscosity) in uniform topology. \begin{lemma} \label{lem:free-energy-concentration-around-expectation} There are positive constants $b_0,b_1,b_2,b_3$ such that for all $n \ge b_0$ and all $u\in (b_3, n\ln n]$, \begin{equation*} \mathsf{P}\{ |\ln Z^n - \mathsf{E} \ln Z^n| > u \} \le b_1 \exp \left\{ -b_2 \frac{u^2}{n\ln^2n} \right\}. \end{equation*} \end{lemma} Here are some notational conventions. For a Borel set $B \subset {\mathbb R}^{n-1}$, we define \begin{equation*} Z^n(B) = \int_{{\mathbb R}\times B\times {\mathbb R}} \prod_{j=0}^{n-1} g(x_{j+1}-x_j)e^{-F_j(x_j)} \delta_0(dx_0)dx_{1}...dx_{n-1}\delta_0(dx_n). \end{equation*} It is often practical to use the same description for events that are technically defined on paths over different time intervals. To that end we introduce more notation. We recall that $\pi_{m,n}$ denotes the restriction of a vector or sequence onto the time interval $[m,n]_{{\mathbb Z}}$. For a Borel set $D \subset {\mathbb R}^{\infty} = S^{-\infty,\infty}_{*,*}$, we define \begin{equation*} \mu_{x,y}^{m,n}(D) = \mu_{x,y}^{m,n}(\pi_{m,n} D), \quad Z^{m,n}_{x,y}(D) = Z^{m,n}(x,y,D) = Z^{m,n}_{x,y} \mu_{x,y}^{m,n}(D). \end{equation*} For example, if $m \le k \le l \le n$ and \begin{equation*} D = \{ \gamma: \gamma_k \in A_k, \dotsc, \gamma_l \in A_l\}, \end{equation*} then \begin{equation*} Z^{m,n}_{x,y}(D) = \int_{x_k \in A_k, \dotsc, x_l \in A_l} \prod_{j=m}^{n-1} g(x_{j+1}-x_j)e^{-F_j(x_j)} \delta_x(dx_m)dx_{m+1}...dx_{n-1}\delta_y(dx_n). \end{equation*} If $x=y=0$, then we write $Z^{m,n}_{x,y}(D) = Z^{m,n}(D)$. \begin{lemma} \label{lem:partition-function-not-too-small} There are constants $\beta, r_0, \rho_0 \in (0,1)$ such that for all $n\in{\mathbb N}$ and any Borel set $B$ that satisfies $[0,1]^{n-1} \subset B \subset {\mathbb R}^{n-1}$, \begin{equation*} \mathsf{P} \left\{ Z^n(B) \le \rho^n \right\} \le \beta^n \left(\frac{\rho}{\rho_0}\right)^{nr_0},\quad \rho>0, \end{equation*} or, equivalently, \begin{equation*} \mathsf{P} \left\{ Z^n(B) \le z \right\} \le \beta^n \frac{z^{r_0}}{\rho_0^{nr_0}},\quad z>0. \end{equation*} \end{lemma} \begin{proof} For any $\rho>0$, \begin{align*} &\mathsf{P}\{Z^n(B)<\rho^n\} \\ \le & \mathsf{P}\left\{\int_{{\mathbb R}\times [0,1]^{n-1}\times {\mathbb R}} \prod_{k=0}^{n-1} [g(x_{k+1}-x_k)e^{-F_k(x_k)}] \delta_0(dx_0)dx_{1}\ldots dx_{n-1}\delta_0(dx_n)<\rho^n \right\} \\ \le & \mathsf{P}\left\{ g^{n}(1)e^{-\sum_{k=0}^{n-1} F^{*}_k(0)} <\rho^n\right\} \le \mathsf{P}\left\{- \sum_{k=0}^{n-1} F^{*}_k(0)<n\ln(\rho g^{-1}(1))\right\}. \end{align*} Let us use the standard Cram\'er large deviation approach. Taking any number $r>0$ and using the Markov inequality, we can write \begin{align*} & \quad\mathsf{P}\bigg\{ -\sum_{k=0}^{n-1} F^*_k(0) <n\ln\big(\rho g^{-1}(1)\big)\bigg\} \\ &= \mathsf{P}\bigg\{ -\sum_{k=0}^{n-1} \Big(F^*_k(0)-\mathsf{E} F^*_k(0) \Big) <n\ln \Big(\rho g^{-1}(1)e^{\mathsf{E} F^*_k(0) } \Big) \bigg\} \\ &\le \exp \Big( rn\ln\big(\rho g^{-1}(1)e^{\mathsf{E} F^*_k(0) }\big) \Big) \Big(\mathsf{E} e^{r \big(F^{*}_0(0)-\mathsf{E} F^{*}_0(0)\big)}\Big)^n \\ &\le \exp \Big( n\big(r\ln(\rho g^{-1}(1)e^{\mathsf{E} F^*_k(0) }) + \Lambda(r) \big) \Big) \\ &\le \exp\Big( n\big(r\ln(\rho_0 g^{-1}(1)e^{\mathsf{E} F^*_k(0)}) + \Lambda(r) \big) \Big) \Big(\frac{\rho}{\rho_0}\Big)^{nr} \le e^{nH(r,\rho_0)}\left(\frac{\rho}{\rho_0}\right)^{nr}, \end{align*} where \[ \Lambda(r)=\ln \mathsf{E} e^{r (F^{*}_0(0)-\mathsf{E} F^{*}_0(0))},\quad r \le \eta, \] and \[ H(r,\rho_0)= r\ln(\rho_0 g^{-1}(1)e^{\mathsf{E} F^*_k(0)}) + \Lambda(r) ,\quad r,\rho_0>0. \] Here we use the assumption \ref{item:exponential-moment-for-maximum}. There is $\sigma>0$ such that \[ \Lambda(r)=\frac{\sigma^2}{2} r^2+o(r^2),\quad r\to 0. \] Therefore, choosing $\rho_0\in(0,1)$ small enough to guarantee that \[ \rho_0 g^{-1}(1) e^{\mathsf{E} F^*_k(0)}<1, \] and choosing $r \in (0,1)$ sufficiently small we guarantee that $H(r,\rho_0)<0$, and the lemma follows. \end{proof} We recall the standard tail estimate for normal distribution: \begin{lemma} \label{lem:tail-estimate-for-normal} For any $a, b > 0$, \begin{equation*} \int_{x \ge b } e^{-\frac{x^2}{a}} dx \le \frac{a}{2b} e^{-\frac{b^2}{a}};\quad\quad\quad \int_{|x| \ge b } e^{-\frac{x^2}{a}} dx \le \frac{a}{b} e^{-\frac{b^2}{a}}. \end{equation*} \end{lemma} Let \begin{equation*} \Gamma(R) = \{ (x_1, ..., x_{n-1}) \in {\mathbb R}^{n-1}: |x_i| \le Rn, i=1,...,n-1 \}. \end{equation*} \begin{lemma}\label{lem:large-deviation-for-Z-bar} There are constants $R, r_1 > 0$ such that for sufficiently large $n$, \begin{equation*} \mathsf{P} \left\{ \frac{Z^n(\Gamma(R)^c)}{Z^n} > 2^{-n} \right\} \le e^{-r_1 n}, \end{equation*} or, in terms of polymer measures, \begin{equation*} \mathsf{P} \Big\{ \mu_{0,0}^{0,n} \big\{ \gamma: \max_{1\le k \le n-1}|\gamma_k| \ge Rn \big\} \ge 2^{-n} \Big\} \le e^{-r_1n}. \end{equation*} \end{lemma} \begin{proof} Suppose $1 \le k \le n-1$. By \ref{item:exponential-moment-with-viscosity} and the Fubini theorem, \begin{equation*} \begin{split} \mathsf{E} \int_{|y| \ge Rn} Z^{0,k}(0,y) Z^{k,n}(y,0) dy &= \lambda^n \int_{|y| \ge Rn} g_k(y)g_{n-k}(y) dy \\ &= \lambda^n \int_{|y| \ge Rn} \frac{1}{2\pi \sqrt{k(n-k)}} e^{-\frac{y^2}{2k} - \frac{y^2}{2(n-k)}} dy. \\ \end{split} \end{equation*} Since $k(n-k) \ge n-1$ and \begin{equation*} \frac{y^2}{2k} + \frac{y^2}{2(n-k)} \ge \frac{y^2}{n}, \end{equation*} we can continue the computation to find \begin{equation*} \mathsf{E} \int_{|y| \ge Rn} Z^{0,k}(0,y) Z^{k,n}(y,0) dy \le \frac{\lambda^n}{2\pi \sqrt{n-1}} \int_{|y| \ge Rn} e^{- \frac{y^2}{n}} dy \le \frac{\lambda^n}{2\pi R\sqrt{n-1}} e^{-R^2n}. \end{equation*} Then by Markov inequality, \begin{equation} \label{eq:nominator-not-too-big} \begin{split} \mathsf{P} \left\{ Z^n\big(\Gamma(R)^c\big) \ge \left(\frac{\rho_0}{2}\right)^n \right\} &\le \frac{\mathsf{E} Z^n \bigl( \Gamma(R)^c \bigr)}{(\rho_0/2)^{n}} \\ & \le \frac{1}{(\rho_0/2)^n} \sum_{k=1}^{n-1} \mathsf{E} \int_{|y| \ge Rn} Z^{0,k}(0,y) Z^{k,n}(y,0) dy \\ & \le \frac{n\lambda^n e^{- R^2n}}{2\pi R \sqrt{n-1} (\rho_0/2)^n} \le e^{-\theta_1n} \end{split} \end{equation} for some $\theta_1>0$ if $R$ is chosen large enough. Now, taking $\rho=\rho_0$ in Lemma \ref{lem:partition-function-not-too-small}, we can find $\theta_2>0$ such that \begin{equation*} \mathsf{P} \{ Z^n \le \rho_0^n \} \le e^{-\theta_2n}, \end{equation*} and the lemma follows. \end{proof} \begin{lemma} \label{lem:second-moment-growth-of-partition-function} There is a constant $d_1$ such that for any Borel set $B$ that satisfies $[0,1]^{n-1} \subset B \subset {\mathbb R}^{n-1}$, \begin{equation*} \mathsf{E} \ln^2 Z^n(B) \le d_1 n^2. \end{equation*} In particular, $\mathsf{E} \ln^2 Z^n \le d_1 n^2$ for all n. \end{lemma} \begin{proof} We will apply Lemma \ref{lem:partition-function-not-too-small} to control $\ln^2 Z^n(B)$ when $Z^n(B)$ is small, and when $Z^n(B)$ is large we will use Markov inequality. Taking any $\mu > \lambda \vee 1$, we split $\mathsf{E} \ln^{2}Z^n(B)$ into three parts \begin{align*} \mathsf{E} \ln^2Z^n(B) &= A_1 +A_2 + A_{3 } \\ &= \mathsf{E} \ln^2Z^n(B){\mathbf{1}}_{\{Z^n(B) \le \rho_0^n\}} + \mathsf{E} \ln^2Z^n(B) {\mathbf{1}}_{\{ \rho_0^n < Z^n(B) < \mu^n \}}\\ &\quad + \mathsf{E} \ln^{2 } Z^n(B) {\mathbf{1}}_{\{ Z^n(B) \ge \mu^n \}}. \end{align*} Clearly $A_2 \le n^2 (\ln^2\rho_0 \vee \ln^{2}\mu)$. To estimate $A_1$, we apply Lemma \ref{lem:partition-function-not-too-small}: \begin{align*} \mathsf{E} \ln^2 Z^2(B){\mathbf{1}}_{\{ Z^n(B) \le \rho_0^n \}} &= \sum_{m=0}^{\infty} \mathsf{E} \ln^2 Z^n(B) {\mathbf{1}}_{\{ \rho_0^n2^{-m-1} < Z^n(B) \le \rho_0^n2^{-m} \}} \\ &\le \sum_{m=0}^{\infty} \ln^2(\rho_0^n2^{-m-1}) \mathsf{P} \{ Z^n(B) \le \rho_0^n2^{-m} \}\\ &\le \sum_{m=0}^{\infty}(n \ln \rho_0 - (m+1)\ln 2)^2 \beta^n 2^{-mr_0}\\ &\le \beta^{n} \sum_{m=0}^{\infty} 2(n^2 \ln^2\rho_0 + (m+1)^2 \ln^2 2) 2^{-mr_0}\le Cn^2\beta^n, \end{align*} for some $C>0$ and all $n\in{\mathbb N}$. For $A_3$, noting that $Z^{n}(B) \le Z^{n}$, we have \begin{align*} A_3 &\le \sum_{k=1}^{\infty} \mathsf{E} \left[ \ln^2 Z^n(B) {\mathbf{1}}_{\{ \mu^{nk} \le Z^n(B) < \mu^{n(k+1)}\}} \right] \le \sum_{k=1}^{\infty} \ln^2\mu^{n(k+1)}\, \mathsf{P} \{ Z^n > \mu^{nk}\} \\ & \le n^2 \sum_{k=1}^{\infty} (k+1)^2 \ln^2\mu \cdot \frac{\lambda^n}{\mu^{nk}} \le Cn^2, \end{align*} for some $C>0$ and all $n\in{\mathbb N}$, which completes the proof. \end{proof} Let us fix $R \in {\mathbb N}$ such that Lemma \ref{lem:large-deviation-for-Z-bar} holds and denote $Z^n(\Gamma(R))$ by $\tilde{Z}^n$. \begin{lemma} \label{lem:truncated-log-partition-function-expectation} For some constant $d_2 > 0$, \[ 0 \le \mathsf{E} \ln Z^n - \mathsf{E} \ln \tilde{Z}^n \le d_2,\quad n\in{\mathbb N}. \] \end{lemma} \begin{proof} The first inequality is obvious since $\tilde{Z}^n \le Z^n$. Let $\Lambda = \{ \tilde{Z}^{n}/Z^n \le 1 - 2^{-n} \}$. By Lemma~\ref{lem:large-deviation-for-Z-bar}, $\mathsf{P}(\Lambda) \le e^{-r_1n}$. We also have $\mathsf{E} \ln^2 \tilde{Z}^n \le d_1n^2$ and $\mathsf{E} \ln^2 Z^n \le d_1n^2$ by Lemma \ref{lem:second-moment-growth-of-partition-function}. Therefore, \begin{align*} \mathsf{E} \ln Z^n - \mathsf{E}\ln \tilde{Z}^{n} &\le \mathsf{E} \big(-\ln(\tilde{Z}^n/Z^n) {\mathbf{1}}_{\Lambda^c} \big) + \mathsf{E} (|\ln Z^n| + |\ln \tilde{Z}^n|){\mathbf{1}}_{\Lambda} \\ &\le -\ln(1-2^{-n}) + \sqrt{2(\mathsf{E} \ln^2 Z^n + \mathsf{E} \ln^2 \tilde{Z}^n) } \sqrt{\mathsf{P} (\Lambda)} \\ & \le \ln 2 + \sqrt{4d_1n^2 e^{-r_1n}} \le d_2, \end{align*} and the lemma is proved. \end{proof} To obtain a concentration inequality for $\ln \tilde{Z}^n$, we will apply Azuma's inequality: \begin{lemma} \label{lem:azuma} Let $(M_k)_{0\le k\le N}$ be a martingale with respect to a filtration $(\mathcal{F}_k)_{k=0}^{N}$. Assume there are constants $c_k$, $1\le k\le N$ such that the increments $\Delta_k = M_k - M_{k-1}$ satisfy \begin{equation*} | \Delta_k | \le c,\quad k=1,\dots, N. \end{equation*} Then \begin{equation*} \mathsf{P} \left\{ |M_N - M_0| \ge x \right\} \le 2 \exp \left( \frac{-x^2}{ 2Nc^2} \right). \end{equation*} \end{lemma} To apply the Azuma inequality, we need to introduce an appropriate martingale with bounded increments. The modified partition function $\tilde{Z}^{n}$ depends only on the potential process on $\{0,1,\dotsc,n-1\}\times [-Rn, Rn]$, and we need an additional truncation of the potential on this set. Let $b>4/\eta$, where $\eta$ is taken from condition \ref{item:exponential-moment-for-maximum}. For $0 \le k \le n-1$ and $x \in [-Rn, Rn]$, we define (suppressing the dependence on $n$ for brevity) \[\xi_k = \max\{F^{*}_k(j): j=-Rn, -Rn+1, ..., Rn-1 \},\quad k=0,\ldots,n,\] \begin{equation*} \bar{F}_k(x) = \begin{cases} 0, & \xi_k \ge b\ln n,\\ F_k(x), & \text{otherwise}, \end{cases} \end{equation*} and, setting $x_0=x_n=0$, \begin{equation*} \tilde{Z}^n(\bar{F}) = \int_{|x_i| \le Rn} \prod_{j=0}^{n-1} g(x_{j+1}-x_j) e^{-\bar{F}_j(x_j)} dx_1\ldots dx_{n-1}. \end{equation*} \begin{lemma} \label{lem:partition_function_of_omega_bar_and_omega} For sufficiently large $n\in{\mathbb N}$, the following holds true: \begin{gather} \label{eq:item:3} \mathsf{E} \exp \Bigl( \frac{\eta}{2} \xi_k {\mathbf{1}}_{\{\xi_k \ge b\ln n\}} \Bigr) \le 2,\\ \label{eq:item:4} \mathsf{E} \xi_k \le b\ln n + 4/\eta,\\ \label{eq:item:8} \mathsf{P} \{ | \ln \tilde{Z}^{n} - \ln \tilde{Z}^n(\bar{F}) | > x \} \le 2e^{-\eta x/2},\quad x > 0,\\ \label{eq:item:9} |\mathsf{E} \ln \tilde{Z}^n - \mathsf{E}\ln \tilde{Z}^n(\bar{F})| \le 4/\eta. \end{gather} \end{lemma} \begin{proof} Since $\xi_k$ is the maximum of $2Rn$ random variables with the same distribution, we have \begin{align}\notag \mathsf{E} \exp \Bigl( \frac{\eta}{2}\xi_k {\mathbf{1}}_{\{\xi_k > b\ln n\}}\Bigr) &\le 1 + \mathsf{E} e^{\frac{\eta}{2}\xi_k} {\mathbf{1}}_{\{\xi_k > b\ln n\}} \le 1 + \mathsf{E} \sum_{j=-Rn}^{Rn-1} e^{\frac{\eta}{2} F^{*}_k(j)} {\mathbf{1}}_{\{F^{*}_k(j) > b\ln n\}} \\ \notag &\le 1 + 2Rn \mathsf{E} e^{ \frac{\eta}{2} F^{*}_k(0)} {\mathbf{1}}_{\{ F^{*}_k(0) > b\ln n\}} \le 1 + 2Rn \frac{\mathsf{E} e^{\eta F^{*}_k(0)}}{e^{ \frac{b\eta}{2} \ln n}} \\ & \le 1 + \frac{c}{n^{\frac{b\eta}{2}-1}}, \label{eq:exp-moment-above-bln} \end{align} where $c = 2R \mathsf{E} e^{\eta F^{*}_k(0)}$ is a constant. Now \eqref{eq:item:3} follows from $b>4/\eta$. If $x > b\ln n$, then by Markov inequality and~\eqref{eq:item:3}, we have \[ \mathsf{P} \{\xi_k \ge x \} \le \mathsf{P} \{\xi_k {\mathbf{1}}_{\{ \xi_k \ge b\ln n\}} \ge x\} \le e^{-\eta x/2} \mathsf{E} \exp \Bigl( \frac{\eta}{2} \xi_k {\mathbf{1}}_{\{ \xi_k \ge b \ln n\}} \Bigr) \le 2e^{-\eta x/2} \] for sufficiently large $n$. This implies~\eqref{eq:item:4}: \[ \mathsf{E} \xi_k \le b \ln n + \mathsf{E} \xi_k {\mathbf{1}}_{\{\xi_k \ge b\ln n\}} \le b \ln n + \int_{b\ln n}^{\infty} \mathsf{P} \{\xi_k \ge x\} \, dx \le b\ln n + \frac{4}{\eta}. \] It follows from the definition of $\tilde{Z}(\bar{F})$ that \begin{equation*} |\ln \tilde{Z}^{n} - \ln \tilde{Z}^{n}( \bar{F})| \le \sum_{k=0}^{n-1} \xi_k{\mathbf{1}}_{\{\xi_k > b\ln n\}}. \end{equation*} By Markov inequality, the i.i.d.\ property of $(\xi_k)$, and~\eqref{eq:exp-moment-above-bln}, we have \begin{align*} \mathsf{P} \left\{ | \ln \tilde{Z}^{n} - \ln \tilde{Z}^n(\bar{F}) | > x \right\} & \le \mathsf{P} \Bigl\{ \frac{\eta}{2}\sum_{k=0}^{n-1} \xi_k {\mathbf{1}}_{\{\xi_k > b\ln n\}} > \frac{\eta x}{2} \Bigr\} \\ & \le e^{-\eta x/2} \mathsf{E} \exp \Bigl(\frac{\eta}{2} \sum_{k=0}^{n-1} \xi_k {\mathbf{1}}_{\{\xi_k> b\ln n\}} \Bigr)\\ & = e^{-\eta x/2} \Big(\mathsf{E} \exp \big( \frac{\eta}{2}\xi_0 {\mathbf{1}}_{\{\xi_0> b\ln n\}}\big) \Big)^{n} \\ & \le e^{-\eta x/2} (1+ c/n^{\eta b/2-1})^{n}. \end{align*} Since $b>4/\eta$,~\eqref{eq:item:8} follows. It immediately implies \begin{equation*} \begin{split} |\mathsf{E} \ln \tilde{Z}^{n} - \mathsf{E}\ln \tilde{Z}^n(\bar{F}) | &\le \mathsf{E} | \ln \tilde{Z}^n - \ln \tilde{Z}^n(\bar{F})|\\ &= \int_0^{\infty} \mathsf{P} \{ | \ln \tilde{Z}^{n} - \ln \tilde{Z}^n(\bar{F}) | > x \} \, dx \le 4/\eta, \end{split} \end{equation*} so \eqref{eq:item:9} is also proved. \end{proof} \begin{lemma} \label{lem:martingale-concentration-for-ZF-bar} For all $n\in{\mathbb N}$ and $x>0$, \begin{equation*} \mathsf{P} \left\{ |\ln \tilde{Z}^{n}(\bar{F}) - \mathsf{E} \ln \tilde{Z}^n(\bar{F}) | > x \right\} \le 2\exp \left\{ -\frac{x^2}{8nb^2\ln^2 n} \right\}. \end{equation*} \end{lemma} \begin{proof} Let us introduce the following martingale $(M_k,\mathcal{F}_k)_{-1\le k \le n-1}$: \begin{equation*} M_k = \mathsf{E} ( \ln \tilde{Z}^n(\bar{F})\,|\, \mathcal{F}_k ), \quad -1 \le k \le n-1, \end{equation*} where \begin{equation*} \mathcal{F}_{-1}= \{\emptyset,\Omega \}, \quad \mathcal{F}_k = \sigma \big( F_{i}(x) : 0 \le i \le k \big), \quad k=0,\dotsc,n-1. \end{equation*} If we can show that $|M_{k} - M_{k-1}| \le 2b\ln n$, $ 0 \le k \le n-1$, then the conclusion of the lemma follows immediately from Azuma's inequality (Lemma~\ref{lem:azuma}). For a process $\bar G$, an independent distributional copy of $\bar F$, we define \begin{multline*} \tilde{Z}^n ([\bar{F},\bar{G}]_k) = \int_{|x_i|\le Rn} \prod_{i=0}^k g(x_{i+1}-x_i) e^{-\bar{F}_i(x_i)} \\ \cdot \prod_{i=k+1}^{n-1} g(x_{i+1}-x_i) e^{-\bar{G}_i(x_i)} \delta_0(d x_0) dx_1\dotsm dx_{n-1} \delta_0(d x_n). \end{multline*} Denoting by $P_k$ the distribution of $\bar{F}_{k}(\cdot)$, we obtain \begin{equation*} \begin{split} &|M_{k} - M_{k-1}| \\=& \Bigg| \int \ln \tilde{Z}^n([\bar{F},\bar{G}]_k) \prod_{i=k+1}^{n-1} P_i\big( d \bar{G}_i \big) - \int \ln \tilde{Z}^n([\bar{F},\bar{G}]_{k-1}) \prod_{i=k}^{n-1} P_i\big( d \bar{G}_i \big) \Bigg| \\ \le& \int \Big| \ln \tilde{Z}^n([\bar{F},\bar{G}]_k) - \ln \tilde{Z}^n([\bar{F},\bar{G}]_{k-1}) \Big| \prod_{i=k}^{n-1} P_i\big( d \bar{G}_i \big) \\ \le& \int \Big( \sup_{|x| \le Rn} |\bar{F}_k(x)| + \sup_{|x| \le Rn} | \bar{G}_k(x)| \Big) \prod_{i=k}^{n-1} P_i\big( d \bar{G}_i \big) \le 2b\ln n, \\ \end{split} \end{equation*} since $|\bar{F}_k(x)|$ and $|\bar{G}_k(x)|$ are bounded by $b \ln n$. This completes the proof. \end{proof} \begin{proof}[Proof of Lemma \ref{lem:free-energy-concentration-around-expectation}] Suppose $u \in \big( 3(d_2+4/\eta + 3), n \ln n \big] $, where $d_2$ is introduced in Lemma \ref{lem:truncated-log-partition-function-expectation}. Then \begin{align*} \mathsf{P} \big\{ |\ln Z^n -& \mathsf{E} \ln Z^n| \ge u \big\} \le \mathsf{P} \bigl\{ |\ln Z^n - \ln \tilde{Z}^{n}| \ge 1 \bigr\} \\ &+ \mathsf{P} \bigl\{ | \ln \tilde{Z}^n - \ln \tilde{Z}^n(\bar{F}) | \ge \frac{u}{3} \bigr\} + \mathsf{P} \bigl\{ |\ln \tilde{Z}^n(\bar{F}) - \mathsf{E} \tilde{Z}^n(\bar{F}) | \ge \frac{u}{3}\bigr\}\\ &+ \mathsf{P} \bigl\{ |\mathsf{E} \ln \tilde{Z}^n(\bar{F}) - \mathsf{E} \ln \tilde{Z}^n| \ge 4/\eta +1 \bigr\}\\ &+ \mathsf{P} \bigl\{ |\mathsf{E} \ln \tilde{Z}^n - \mathsf{E} \ln Z^n| \ge d_2 +1 \bigr\}. \end{align*} By \eqref{eq:item:9} and Lemma \ref{lem:truncated-log-partition-function-expectation}, the last two terms equal 0. By Lemma \ref{lem:large-deviation-for-Z-bar}, we have \begin{equation*} \mathsf{P} \big\{ |\ln Z^n - \ln \tilde{Z}^n | \ge |\ln (1-2^{-n}) | \; \big\} \le 2^{-r_1n}, \end{equation*} which bounds the first term. The second term is bounded by $2e^{-\eta u/6}$ by \eqref{eq:item:8} and the third term is bounded by $2e^{-u^2/72b^2n\ln n}$ by Lemma \ref{lem:martingale-concentration-for-ZF-bar}. Combining all these estimates together, we obtain \begin{equation*} \mathsf{P} \big\{ |\ln Z^n - \mathsf{E} \ln Z^n| \ge u \big\} \le 2^{-r_1n} + 2e^{-\frac{\eta u}{6}} + 2e^{ - \frac{u^2}{72b^2n \ln^2 n}} \le b_1 e^{-b_2 \frac{u^2}{n\ln^2 n}}, \end{equation*} for some constants $b_1,b_2 > 0$, where in the last inequality we use $u \le n\ln n$. \end{proof} To prove Theorem~\ref{thm:concentration-of-free-energy}, we need to estimate $|\mathsf{E} \ln Z^n - \alpha_0n|$. \begin{lemma} \label{lem:small-probability-for-single-square-root-jump} There are constants $R', r_2 > 0$ such that for sufficiently large $n$, \begin{equation*} \mathsf{P} \Bigl\{ \mu_{0,0}^{0,n} \{ \gamma: | \gamma_k - \gamma_{k+1}| \ge R' \sqrt{n} \} \ge 2^{-n} \Bigr\} \le e^{-r_2n},\quad 0 \le k \le n-1. \end{equation*} \end{lemma} \begin{proof} Let $A_k = \{ \gamma: | \gamma_k - \gamma_{k+1}| \ge R' \sqrt{n} \}$. The random variables $\mu_{0,0}^{0,n}(A_{n-1})$ and $\mu_{0,0}^{0,0}(A_0)$ are identically distributed. So let us assume $k \neq 0$. By Markov inequality, for $\rho_0$ defined in Lemma \ref{lem:partition-function-not-too-small}, \begin{equation*} \begin{split} &\quad \mathsf{P} \Big\{ Z^{0,n}_{0,0}(A_k) \ge \Big(\frac{\rho_0}{2} \Big)^n \Big\} \le \Big( \frac{2}{\rho_0} \Big)^n \mathsf{E} Z^{0,n}_{0,0}(A_k) \\ &\le \Big( \frac{2}{\rho_0}\Big)^n \lambda^n \int_{|x_k-x_{k+1}| \ge R' \sqrt{n}} \prod_{j=0}^{n-1} g(x_j-x_{j+1}) \delta_0(dx_0)dx_1\cdots dx_{n-1} \delta_0(dx_n) \\ &\le \Big( \frac{2}{\rho_0}\Big)^n \lambda^n \frac{1}{\sqrt{2\pi}} \int_{|y_k| \ge R' \sqrt{n}} \prod_{j=1}^{n-1} g(y_j) dy_1 \cdots dy_{n-1} \\ &= \Big( \frac{2}{\rho_0}\Big)^n \lambda^n \frac{1}{\sqrt{2\pi}} \int_{|y_k| \ge R' \sqrt{n}} g(y_k) dy_k \le \Big( \frac{2}{\rho_0}\Big)^n \lambda^n \frac{1}{2\pi R' \sqrt{n}}e^{-R'^2n/2} \le e^{-Cn} \end{split} \end{equation*} for some constant $C>0$ and all $n$ if $R'$ is chosen sufficiently large. Here, in the third line, we use $g(x_0-x_1) \le 1/\sqrt{2\pi}$ and then a change of variables $y_i = x_i-x_{i+1}$. The lemma now follows from \begin{equation*} \mathsf{P} \big\{ \mu_{0,0}^{0,n}(A_k) \ge 2^{-n} \} \le \mathsf{P} \big\{ Z^{0,n}_{0,0} \le \rho_0^n \big\} + \mathsf{P} \Big\{ Z^{0,n}_{0,0}(A_k) \ge \Big( \frac{\rho_0}{2} \Big)^n \Big\} \end{equation*} and Lemma \ref{lem:partition-function-not-too-small}. \end{proof} \begin{lemma} \label{lem:doubling-argument-inequality} There is positive constant $b_4$ such that for sufficiently large $n$, \begin{equation*} |\mathsf{E} \ln Z^{0,2n} - 2\mathsf{E} \ln Z^{0,n} | \le b_4 n^{1/2} \ln^2 n. \end{equation*} \end{lemma} \begin{proof} For $R, R'$ in Lemma \ref{lem:large-deviation-for-Z-bar} and Lemma \ref{lem:small-probability-for-single-square-root-jump}, define \begin{align*} B &= \{ \gamma: \max_{1 \le i \le 2n-1}|\gamma_i| \le 2Rn \}, \\ C &= \{ \gamma : |\gamma_n-\gamma_{n+1}| \le R'\sqrt{2n},\ |\gamma_n-\gamma_{n-1}| \le R' \sqrt{2n} \}. \end{align*} Using an argument similar to the proof of Lemma \ref{lem:truncated-log-partition-function-expectation}, we can obtain \begin{equation} \label{eq:truncation_does-not-change-much} |\mathsf{E} \ln Z^{0,2n} (B\cap C) - \mathsf{E} \ln Z^{0,2n}| \le d_3, \end{equation} for some constant $d_3$. In fact, introducing $ \Lambda = \left\{ \frac{Z^{0,2n}(B\cap C)}{Z^{0,2n}} \le 1-3\cdot 2^{-n} \right\}$, we can write for large $n$: \begin{align*} & \mathsf{E} \ln Z^{0,2n} - \mathsf{E} \ln Z^{0,2n}(B\cap C) \\ \le & \mathsf{E} \big( - \ln (Z^{0,2n}(B \cap C) / Z^{0,2n}) {\mathbf{1}}_{\Lambda^c} \big) + \mathsf{E} \big(|\ln Z^{0,2n}(B\cap C)| + | \ln Z^{0,2n}|\big) {\mathbf{1}}_{\Lambda} \\ \le & -\ln(1-3\cdot 2^{-n}) + \sqrt{2 \big( \ln^2 Z^{0,2n}(B \cap C) + \ln^2 Z^{0,2n} \big) } \sqrt{\mathsf{P}(\Lambda)} \\ \le & -\ln(1-3\cdot 2^{-n}) + \sqrt{4d_1n^2 e^{-K_2n}} \le d_3, \end{align*} where in the last line we used Lemma \ref{lem:second-moment-growth-of-partition-function} and the estimate $\mathsf{P}(\Lambda) \le e^{-K_2n}$ for some constant $K_2$ implied by Lemmas~\ref{lem:large-deviation-for-Z-bar} and~\ref{lem:small-probability-for-single-square-root-jump}. Clearly, $\mathsf{E} \ln Z^{0,2n} > \mathsf{E} \ln Z^{0,2n}(B \cap C)$, so (\ref{eq:truncation_does-not-change-much}) follows. To prove the lemma, it is sufficient now to obtain upper and lower bounds on $\mathsf{E} \ln Z^{0,2n} (B\cap C)$. First, we bound $\mathsf{E} \ln Z^{0,2n}(B\cap C)$ from below. Since $|\gamma_n| \le Rn$ on~$B$, we have \begin{align*} Z^{0,2n}(B\cap C) \ge Z^{0,2n}(B\cap C \cap \{\gamma_n \in [0,1) \}). \end{align*} Let us now compare the action of every path $\gamma$ in $B\cap C \cap \{\gamma_n \in [0,1) \}$ to the action of the modified path $\bar \gamma$ defined by $\bar{\gamma}_n = 0$ and $\bar{\gamma}_j = \gamma_j$ for $j \neq n$. We recall that action was defined in~\eqref{eq:action}. Since $|\gamma_{n+1} - \gamma_{n}| \le R' \sqrt{2n}$, $|\gamma_n-\gamma_{n-1}| \le R' \sqrt{2n}$, and $|\gamma_n| \le 1$, we get \begin{equation*} \begin{split} |A^{0,2n}(\gamma) - A^{0,2n}(\bar \gamma)| &\le \frac{1}{2}\left| (\gamma_{n+1}-\gamma_n)^2-\gamma_{n+1}^2 + (\gamma_{n-1}-\gamma_n)^2 - \gamma_{n-1}^2 \right| + 2F_{n}^{*}(0) \\ &\le 2R' \sqrt{2n} + 1 + 2F_{n}^{*}(0). \end{split} \end{equation*} So, there is a constant $K_3>0$ such that \begin{equation} \label{eq:Z-almost-a-product} Z^{0,2n}(B\cap C)\ge Z^{0,n}(D^-)Z^{n,2n}(D^+) e^{-K_3 \sqrt{n} - 2F_{n}^{*}(0)}, \end{equation} where \begin{align*} D^+ &= \{\gamma: |\gamma_{n+1}| \le R' \sqrt{2n} + 1,\ |\gamma_i| \le 2Rn,\ n+1 \le i \le 2n-1 \}, \\ D^- &= \{ \gamma: |\gamma_{n-1}| \le R' \sqrt{2n} + 1,\ |\gamma_i| \le 2Rn,\ 1 \le i \le n-1\}. \end{align*} Similarly to~\eqref{eq:truncation_does-not-change-much}, there is a constant $d_4$ such that \begin{equation*} |\mathsf{E} \ln Z^{n}(D^{+}) - \mathsf{E} \ln Z^{n} | \le d_4, \quad |\mathsf{E} \ln Z^n(D^-) - \mathsf{E} \ln Z^n | \le d_4. \end{equation*} Combining this with~\eqref{eq:Z-almost-a-product}, we obtain \begin{align*} \mathsf{E} \ln Z^{0,2n}(B \cap C) &\ge \mathsf{E} \ln Z^{0,n}(D^-) + \mathsf{E} \ln Z^{n,2n}(D^+) -K_3 \sqrt{n} - 2\mathsf{E} F_{n}^{*}(0) \\ &\ge 2\mathsf{E} \ln Z^{0,n} - 2d_4-K_3 \sqrt{n} - 2\mathsf{E} F_{n}^{*}(0), \end{align*} the desired lower bound. Next, we bound $\mathsf{E} \ln Z^{0,2n}(B\cap C)$ from above. Similarly to~\eqref{eq:Z-almost-a-product}, we compare actions of generic paths in $B\cap C$ to the actions of the modified paths that take integer values at time $n$: \begin{align*} Z^{0,2n}(B\cap C) &= \sum_{k=-2Rn}^{2Rn-1} Z^{0,2n}(B\cap C\cap \{\gamma_n \in [k,k+1) \} ) \\ &\le \sum_{k=-2Rn}^{2Rn-1} Z^{0,n}(0,k)Z^{n,2n}(k,0) e^{K_3 \sqrt{n} + 2F_{n}^{*}(k)} \\ &\le 4Rn \max_k [Z^{0,n}(0,k)Z^{n,2n}(k,0)] e^{K_3 \sqrt{n} + 2\max_k F_{n}^{*}(k)}, \end{align*} where the maxima are taken over $-2Rn \le k \le 2Rn-1$. Taking logarithm and then expectation of both sides, we obtain \begin{align*} &\quad \mathsf{E} \ln Z^{0,2n}(B\cap C) \\ &\le \mathsf{E} \max_k \ln Z^{0,n}(0,k) + \mathsf{E} \max_k \ln Z^{n,2n}(k,0) + \ln(4Rn) + K_3 \sqrt{n} + 2 \mathsf{E} \max_k F_{n}^{*}(k) \\ &\le \max_k \mathsf{E} \ln Z^{0,n}(0,k) + \mathsf{E} \max_k X_k + \max_k \mathsf{E} \ln Z^{n,2n}(k,0) + \mathsf{E} \max_k Y_k \\ & \hspace{9cm}+ K_{4}(\ln n + \sqrt{n} + 1) \\ &\le 2\mathsf{E}\ln Z^{0,n} + \mathsf{E} \bigl[\max_k X_k+\max_k Y_k\bigr] + K_{4} (\ln n + \sqrt{n} + 1), \end{align*} for some constant $K_4>0$, where \begin{equation*} X_k = \ln Z^{0,n}(0,k) - \mathsf{E}\ln Z^{0,n}(0,k), \quad Y_k = \ln Z^{n,2n}(k,0) - \mathsf{E}\ln Z^{n,2n}(k,0). \end{equation*} In the second inequality, we used~\eqref{eq:item:4} to conclude \begin{equation*} \mathsf{E} \max_{-2Rn\le k \le 2Rn-1}F_{n}^{*}(k) \le b \ln (2n) + 4/\eta, \end{equation*} and in the third inequality, we used the fact that \begin{equation*} \mathsf{E} \ln Z^{0,n}(0,k) \le \mathsf{E} \ln Z^{0,n}, \quad \mathsf{E} \ln Z^{n,2n}(k,0) \le \mathsf{E} \ln Z^{n,2n} = \mathsf{E} \ln Z^{0,n}. \end{equation*} It remains to bound $\mathsf{E} \max_k X_k$ and $\mathsf{E} \max_k Y_k$. By the shear invariance, all $X_k$ and~$Y_k$ have the same distribution, so \begin{equation*} \mathsf{E} X_n^2 = \mathsf{E} Y_n^2 = \mathsf{E} \ln^2 Z^n \le K_5n^2. \end{equation*} Let \begin{equation*} \Lambda = \left\{ \max_k X_k \le rn^{1/2}\ln^{3/2} n, \quad \max_k Y_k \le rn^{1/2}\ln^{3/2}n \right\}, \end{equation*} with $r$ to be determined. We have \begin{align*} \mathsf{E} \left[\max_k X_k + \max_k Y_k\right] &\le \mathsf{E} {\mathbf{1}}_{\Lambda} (\max_kX_k + \max_kY_k) + \mathsf{E} {\mathbf{1}}_{\Lambda^{c}} (\max_kX_k+\max_kY_k) \\ & \le 2rn^{1/2}\ln^{3/2}n + \sqrt{2\mathsf{P}(\Lambda^c) \mathsf{E}( \max_kX_k^2+\max_kY_k^2)} \\ & \le 2rn^{1/2} \ln^{3/2}n + \sqrt{16\mathsf{P}(\Lambda^c) K_5Rn^3}. \end{align*} To bound the second term by a constant, we use Lemma~\ref{lem:free-energy-concentration-around-expectation}: \begin{align*} \mathsf{P} (\Lambda^c) &\le \sum_{k=-2Rn}^{2Rn-1} \Biggl[\mathsf{P} \left\{ |\ln Z^{0,n}(0,k) - \mathsf{E}\ln Z^{0,n}(0,k)| \ge rn^{1/2}\ln^{3/2}n \right\} \\ &\quad + \mathsf{P} \left\{ |\ln Z^{n,2n}(k,0) - \mathsf{E}\ln Z^{n,2n}(k,0)| \ge rn^{1/2} \ln^{3/2}n \right\} \Biggr]\\ &\le 8Rn \mathsf{P} \left\{ |\ln Z^n - \mathsf{E}\ln Z^{n}| \ge rn^{1/2}\ln^{3/2}n \right\} \\ & \le 8Rn b_1 \exp \{-b_2 r^2 \ln n \}, \end{align*} and choose $r$ to ensure $b_2r^2>4$. This completes the proof.\end{proof} We can now use the following straightforward adaptation of~Lemma~4.2 of~\cite{HoNe} from real argument functions to sequences: \begin{lemma} \label{lem:doubling-argument-lemma} Suppose that number sequences $(a_n)$ and $(g_n)$ satisfy the following conditions: $a_n/n \to \nu$ as $n \to \infty$, $|a_{2n} - 2a_n| \le g_n$ for large $n$ and $\lim_{n\to \infty} g_{2n}/g_n = \psi < 2$. Then for any $c > 1/(2-\psi)$ and for all large $n$ (depending on $c$), \begin{equation*} |a_n - \nu n | \le c g_n. \end{equation*} \end{lemma} \begin{proof} Let $b_n = a_n/n$, $h_n = g_n/(2n)$. Then $|b_{2n} - b_n| \le h_n$ for large $n$ and $\lim_{n\to \infty } h_{2n}/h_n = \psi/2$. Since $\psi/2 < 1 - \frac{1}{2c} $, there is $N$ such that $h_{2m}/ h_m \le 1 -\frac{1}{2c}$ for all $m > N$. We can assume further that for $m > N$, the inequality $|b_{2m}-b_m| \le h_m$ holds. Let us now fix $n > N$. Then for $k \ge 0$ we have $h_{2^kn} \le \bigl( 1-\frac{1}{2c} \bigr)^k h_{n}$. Therefore, \begin{equation*} | b_n - b_{2^kn}| \le \sum_{i=0}^{k-1} |b_{2^{i+1}n} - b_{2^in} | \le \sum_{i=0}^{k-1} h_{2^in} \le 2c h_{n}. \end{equation*} We complete the proof by letting $k \to \infty$ in this estimate. \end{proof} \begin{proof}[Proof of Theorem \ref{thm:concentration-of-free-energy}] It suffices to prove the inequality for $v = 0$. Thanks to Lemma \ref{lem:doubling-argument-inequality} and Theorem \ref{thm:shape-function}, we can apply Lemma \ref{lem:doubling-argument-lemma} to $a_n = \mathsf{E} Z^n$, $g_n = b_4 n^{1/2} \ln^2n$, $\nu = \alpha_0$ and $\psi =\sqrt{2}$ to obtain \begin{equation} \label{eq:expectation-deviation-from-linear} |\mathsf{E} Z^n - \alpha_0n| \le c n^{1/2} \ln^2n \end{equation} for some constant $c$ and sufficiently large $n$. Then, for all sufficiently large $n$ and $u \in (2cn^{1/2}\ln^2n, n\ln n]$, we have \begin{equation*} \mathsf{P} \bigl\{ |Z^n - \alpha_0n| \ge u \bigr\} \le \mathsf{P} \bigl\{ |Z^n - \mathsf{E} Z^n| \ge u/2 \bigr\} \le b_1 \exp \left( - b_2 \frac{u^2}{4n\ln^2n} \right), \end{equation*} by Lemma~\ref{lem:free-energy-concentration-around-expectation} and~(\ref{eq:expectation-deviation-from-linear}). \end{proof} \section{Straightness and tightness} \label{sec:delta-straightness-and-tightness} \subsection{Straightness} \label{sec:delta-straightness} The notion of $\delta$-straightness of paths in random environments was introduced in~\cite{Ne}. In this section, we prove an analogous straightness property for the positive temperature polymer measure case. Some of the results of this section may be interpreted in terms of bounds on the transversal fluctuation exponent. It is expected that the fluctuations of a typical polymer path of length~$n$ are of order $n^\xi$ for some $\xi$, and the KPZ scalings predict $\xi=2/3$. Our results show that $\xi\le 3/4$. We begin with some notation. For a path $\gamma$ and $n\in {\mathbb Z}$, we define \[ \gamma^{\rm out}(n) = \{ (m,\gamma_m):\ m>n\}.\] For $v>0$, we define \[{\rm Co}(v) = \{ (n,x)\in{\mathbb N}\times {\mathbb R}:\ |x|\leq n v\}.\] For $(n,x)\in {\mathbb N}\times {\mathbb R}$ and $\eta>0$, we define: \begin{equation*} {\rm Co}(n,x,\eta) = \{ (m,y)\in {\mathbb N}\times {\mathbb R}\ :\ |y/m - x/n| \le\eta \}. \end{equation*} \medskip Let us fix $\delta \in (0,1/4)$. \begin{theorem} \label{thm:delta-straightness-for-polymer-measure} For any $v > 0$ and $\alpha \in (0,1-2\delta)$, there are events $\Omega_1^{(n)} = \Omega_1^{(n)}(\delta,v,\alpha)$ and a constant $Q=Q(\delta)$ with the following properties: \begin{enumerate} \item\label{item:13} For all $\beta \in (0, 1-4\delta)$ and sufficiently large $n$ (depending on $\beta$), \begin{equation*} \mathsf{P} (\Omega_1^{(n)}) \ge 1 - e^{-n^{\beta}}; \end{equation*} \item\label{item:14} on the event $\Omega_1^{(n)}$ the following is true: for any terminal measure $\nu$ and for all $N\in{\mathbb N}$ satisfying $N/2 > n$, \begin{equation} \label{eq:delta-straightness-for-polymer-measure} \mu_{0,\nu}^{0,N} \big\{ \gamma : \exists (k,\gamma_k) \in \mathrm{Co}(v),\ N/2 \ge k \ge n,\ \gamma^{\text{out}} \not\subset \mathrm{Co}(k,\gamma_k,Qk^{-\delta}) \big\} \le e^{-n^{\alpha}}. \end{equation} \end{enumerate} \end{theorem} For $(m,x), (n,y) \in {\mathbb Z} \times {\mathbb R}$ with $m<n$, we define $[(m,x),(n,y)]$ to be the constant velocity path connecting $(m,x)$ and $(n,y)$, i.e., $[(m,x),(n,y)]_k=x+\frac{k-m}{n-m}(y-x)$ for $k\in [m,n]_{{\mathbb Z}}$. For any $\alpha > 0$, let us define the event $A_{x,y}^{m,n} = A_{x,y}^{m,n}(\alpha)$: \begin{equation*} A_{x,y}^{m,n} = \Big\{ \mu_{x,y}^{m,n} \bigl\{ \max_{k \in I(m,n)} |\gamma_k - [(m,x),(n,y)]_k| \ge (n-m)^{1-\delta} \bigr\} \le \exp\big(-(n-m)^{\alpha}\big) \Big\}, \end{equation*} where $I(m,n) = [\frac{3m+n}{4}, \frac{m+3n}{4}]_{{\mathbb Z}}$. By translation and shear invariance all $A_{x,y}^{m,n}$ have the same probability for fixed $m$ and $n$. We also define the event $\bar{A}_{p,q}^{m,n} = \bar{A}_{p,q}^{m,n}(\alpha)$ for $p, q \in {\mathbb Z}$ to be \begin{equation*} \bar{A}_{p,q}^{m,n} = \bigcap \big\{ A_{x,y}^{m,n} : x \in [p,p+1], y \in [q,q+1] \big\}. \end{equation*} This is a measurable event since, by continuity, the intersection on the right-hand side may be restricted to rational points $x,y$. Also, all $\bar{A}_{p,q}^{m,n}$ have the same probability for fixed $m$ and $n$. \begin{lemma}\label{lem:concentration-of-polymer-measure-in-delta-rectangle} Let $\alpha \in (0,1-2\delta)$ and $\beta \in (0,1-4\delta)$. Then \begin{equation*} \mathsf{P} \big( \bar{A}_{0,0}^{0,n}(\alpha) \big) \ge 1- e^{-n^{\beta}} \end{equation*} for sufficiently large $n$. \end{lemma} \begin{proof} Lemma~\ref{lem:main-monotonicity} implies $\mu_{0,0}^{0,n}\preceq\mu_{x,y}^{0,n}\preceq \mu_{1,1}^{0,n}$ for $x,y\in[0,1]$. Therefore, it suffices to check that for any $\alpha' \in (0,1-2\delta)$, $\beta' \in (0,1-4\delta)$, all sufficiently large~$n$ and all~$k\in [n/4 ,3n/4]$, \begin{equation*} \mathsf{P} \Bigl\{ \mu_{0,0}^{0,n} \{ \gamma : |\gamma_k| > n^{1-\delta}-1 \} \ge e^{-n^{\alpha'}} \Bigr\} \le e^{-n^{\beta'}}. \end{equation*} So let us prove this estimate. Let \begin{align*} B &= \{ \gamma : \max_{1\le i \le n-1} |\gamma_i| \le Rn \},\\ C &= \{ \gamma : |\gamma_k - \gamma_{k+1} | \le R' \sqrt{n},\ |\gamma_k - \gamma_{k-1} | \le R' \sqrt{n} \}, \end{align*} for $R, R'$ given in Lemma \ref{lem:large-deviation-for-Z-bar} and Lemma \ref{lem:small-probability-for-single-square-root-jump}. By these two lemmas, the complement of the event $A=\{\mu_{0,0}^{0,n}(B^c\cup C^c) \le 3\cdot 2^{-n} \}$ is small: \begin{multline} \label{eq:control-for-A} \mathsf{P}(A^c) \le \mathsf{P} \big\{ \mu_{0,0}^{0,n}(B^c) \ge 2^{-n} \big\} + \mathsf{P} \big\{ \mu_{0,0}^{0,n}(C^c) \ge 2\cdot 2^{-n} \big\} \le 2^{-r_1n} + 2\cdot 2^{-r_2n}. \end{multline} Let $D = \bigl\{ \max\limits_{|j|\le Rn} F_{k}^{*}(j) \le \sqrt{n} \bigr\}$. By \ref{item:exponential-moment-for-maximum} and Markov inequality, \begin{equation} \label{eq:1} \mathsf{P} (D^c) \le \sum_{|j|\le Rn} \mathsf{P} \{ F_{k}^{*}(j) \ge \sqrt{n} \} \le (2Rn+1)e^{-\eta \sqrt{n}} \mathsf{E} e^{\eta F_{k}^{*}(0)} \le e^{-k_1 \sqrt{n}}, \end{equation} for some constant $k_1>0$ and all sufficiently large $n$. Let us also fix a number $\theta \in \Big( \frac{\beta'+1}{2}, 1-2\delta \Big)$, and for $|j|\le Rn$, define \begin{equation*} \begin{split} F^{+}_j &= \{ |\ln Z^{k,n}(j,0) - \alpha(n-k,j) | \le n^{\theta} \}, \\ F^-_j & = \{ |\ln Z^{0,k}(0,j) - \alpha(k,j) | \le n^{\theta} \}, \\ F &= \{ |\ln Z^{0,n}(0,0) - \alpha(n,0) | \le n^{\theta} \},\\ \hat F&= F\cap \bigcap_{|j|\le Rn} F^{+}_j \cap \bigcap_{|j|\le Rn} F^{-}_j, \end{split} \end{equation*} where \begin{equation} \label{eq:space-time-shape} \alpha(n,x) = \alpha_0n - \frac{x^2}{2n}. \end{equation} Let $\hat{k} = \min\{k,n-k\}$. By Theorem~\ref{thm:concentration-of-free-energy}, if $\hat{k}\ge n/4$, then for some $\theta'\in(\beta',2\theta - 1)$ and sufficiently large $n$, \begin{equation} \label{eq:3} \mathsf{P}(\hat F) \le (4Rn+3)e^{-n^{\theta'}}. \end{equation} On the event $\Lambda = A\cap D \cap \hat F$, we can control $\mu_{0,0}^{0,n}$. From (\ref{eq:control-for-A}), (\ref{eq:1}), (\ref{eq:3}) we get \begin{equation*} \mathsf{P} (\Lambda^c) \le e^{-r_1n } + 2\cdot e^{-r_2n} + e^{-k_1 \sqrt{n}} + (4Rn+3)e^{-n^{\theta'}} \le e^{-n^{\beta'}}. \end{equation*} Now it suffices to show that for sufficiently large $n$, \begin{equation} \label{eq:Lambda-means-straightness} \Lambda \subset \bigl\{ \mu_{0,0}^{0,n} \{ |\gamma_k| \ge n^{1-\delta} - 1 \} < e^{-n^{\alpha'}} \bigr\}. \end{equation} On $A$, we have \begin{equation} \label{eq:5} \mu_{0,0}^{0,n}\{ |\gamma_k| \ge n^{1-\delta}-1 \} \le \mu_{0,0}^{0,n} \big( \{ |\gamma_k| \ge n^{1-\delta}-1 \} \cap B \cap C \big) + 3\cdot 2^{-n}. \end{equation} For every path in $B \cap C$, we can move $\gamma_k$ to $[\gamma_k]$ to obtain a new path. The difference between integrating over the old paths and over the new ones can be estimated as follows. Since all the paths are in $C$, the kinetic action will not change by more than $2R'\sqrt{n}+1$, and since we are on the set $D$, the potential will change by at most $2\max_{|j| \le Rn} F_{k}^{*}(j)\le 2 \sqrt{n}$. The integral over modified paths can be viewed as a sum of partition functions over integer endpoints. Therefore, we obtain \begin{equation*} \mu^{0,n}_{0,0}\big\{ \{|\gamma_k| \ge n^{1-\delta} -1 \} \cap B \cap C \big\} \le e^{(2R'+3) \sqrt{n}} \bigl( Z^{0,n}_{0,0} \bigr)^{-1} \sum_{j \in J } Z^{0,k}(0,j) Z^{k,n}(j,0), \end{equation*} where $J = [n^{1-\delta}-1, Rn]_{\mathbb Z}$. On $\hat F$, we have \begin{align*} &\quad \mu_{0,0}^{0,n} \big\{ \{ |\gamma_k| \ge n^{1-\delta} -1\} \cap B \cap C \big\} \\ &\le e^{(2R'+3) \sqrt{n} + n^{\theta} - \alpha_0 n} \sum_{j \in J} \exp \Big(2n^{\theta} + \alpha_0k - \frac{j^2}{2k} + \alpha_0 (n-k)- \frac{j^2}{2(n-k)} \Big) \\ & \le e^{ (2R'+3) \sqrt{n} + 3n^{\theta} } \sum_{|j| \ge n^{1-\delta}-1} \exp \Big( -\frac{j^2}{2k}- \frac{j^2}{2(n-k)} \Big) \\ &\le e^{(2R'+3) \sqrt{n} + 3n^{\theta}} \sum_{|j| \ge n^{1-\delta}-1} \exp \Big(-\frac{2j^2}{n}\Big) \\ &\le k_2\exp\big( (2R'+3) \sqrt{n} + 3n^{\theta} - 2n^{1-2\delta} \big)\\ &\le \exp (-k_3 n^{1-2\delta} ) \end{align*} for some positive constants $k_2$, $k_3$ and sufficiently large $n$. In the last inequality, we use $\theta <1-2\delta$. We obtain~\eqref{eq:Lambda-means-straightness} from this and (\ref{eq:5}), and the lemma follows. \end{proof} Let us introduce \begin{equation*} \mathcal{C}(n,x,L) := \left\{ (m,y) \in \Z\times{\mathbb R}:\ m \in \{n+1,...,2n\},\ \left| y - \frac{m}{n}x \right| \le L \right\}, \end{equation*} a parallelogram of width $2L$ with one pair of sides parallel to the $x$-coordinate axis and the other one parallel to $[(n,x),(2n,2x)]$. We define the lateral sides by \begin{equation*} \partial_S^{\pm} \mathcal{C}(n,x,L) := \left\{ (m,y) \in \Z\times{\mathbb R}:\ m \in \{ n+1, ...,2n \} \text{ and } y - \frac{m}{n}x = \pm L \right\}, \end{equation*} and let $\partial_S \mathcal{C}(n,x,L) = \partial_S^{+} \mathcal{C}(n,x,L) \cup \partial_S^{-} \mathcal{C}(n,x,L)$. Let us define $H(n,x)$ to be \begin{equation*} H(n,x) = \left\{ \gamma: \gamma_n \in [x,x+1],\ \exists m \in \{ n+1, ..., 2n\},\ \left| \gamma_m - \frac{m}{n}x \right| > 6n^{1-\delta} \right\}. \end{equation*} Then for $N \ge 2n$, $\mu_{0,\nu}^{0,N}\big( H(n,x) \big)$ is the polymer measure of the paths that start at $(0,0)$, pass near $(n,x)$ but exit the parallelogram $\mathcal{C}(n,x,6n^{1-\delta})$ at a time between $n+1$ and $2n$. The constant 6 is chosen to be compatible with Lemma \ref{lem:concentration-of-polymer-measure-in-delta-rectangle}. \begin{lemma} \label{lem:polymer-measure-of-Hnx-is-small} Given $\alpha \in (0, 1-2\delta)$, if $A_{0,y}^{0,m}(\alpha)$ holds for all $(m,y) \in \mathcal{C}(n,x,6n^{1-\delta}) \cap \{ m \ge 3n/2 \}$, then for any $N \ge 2n$ and any terminal measure $\nu$, \begin{equation} \label{eq:polymer-measure-of-Hnx-is-small} \mu_{0,\nu}^{0,N} \big( H(n,x) \big) \le 2n\exp(-n^{\alpha} ). \end{equation} \end{lemma} \begin{proof} Let \begin{align*} B_k^+ &= \big\{ \gamma:\ \gamma_n \in [x,x+1],\ \gamma_k -\frac{k}{n}x > 4n^{1-\delta} \big\}, \quad 3n/2 \le k \le 2n, \\ B_k^- &= \big\{ \gamma:\ \gamma_n \in [x,x+1],\ \gamma_k -\frac{k}{n}x < -4n^{1-\delta} \big\}, \quad 3n/2 \le k \le 2n , \\ C_k^+ &= \big\{ \gamma:\ \gamma_n \in [x,x+1],\ \gamma_{2n} -2x \le 4n^{1-\delta},\ \gamma_k - \frac{k}{n}x > 6n^{1-\delta} \big\}, \quad n < k < \frac{3n}{2}, \\ C_k^{-} &= \big\{ \gamma:\,\gamma_n \in [x,x+1],\,\gamma_{2n} -2x \ge - 4n^{1-\delta},\,\gamma_k - \frac{k}{n}x <- 6n^{1-\delta} \big\},\ n< k < \frac{3n}{2}. \end{align*} Then $H(n,x) \subset \bigcup_k B_k^{\pm} \cup \bigcup_k C_k^{\pm}$. We need to estimate $\mu_{0,\nu}^{0,N}(B_k^{\pm})$ and $\mu_{0,\nu}^{0,N}(C_k^{\pm})$. First, we look at $B_k^+$. By monotonicity, if $ y > kx/n + 4n^{1-\delta} := f(k)$, then \begin{equation*} \mu_{0,y}^{0,k}(B_k^+)\le \mu_{0,y}^{0,k}\{\gamma: \gamma_n \le x+1 \} \le \mu_{0,f(k)}^{0,k}\{ \gamma: \gamma_n \le x+1\}. \end{equation*} Since \begin{equation*} \big[ (0,0),(k,f(k) ) \big]_n - (x +1) = \frac{4n^{2-\delta}}{k} - 1 \ge \frac{4n^{2-\delta}}{2n} - 1 > (2n)^{1-\delta} \ge k^{1-\delta} \end{equation*} for sufficiently large $n$, we have \begin{equation*} \{ \gamma: \gamma_n \le x+1 \} \subset \Big\{ \gamma : \max_{j \in [k/4, 3k/4]} \Big|\gamma_j - \big[(0,0),(k,f(k))\big]_{j}\Big| \ge k^{1-\delta} \Big\}. \end{equation*} By assumption, $A_{0,f(k)}^{0,k}$ holds. Therefore, if $y \ge f(k)$, then \begin{equation*} \mu_{0,y}^{0,k}(B_k^+)\le e^{-k^{\alpha}} \le e^{-n^{\alpha}}, \end{equation*} and hence \begin{equation*} Z^{0,k}(0, y,B_k^+) \le \exp( -n^{\alpha} ) Z^{0,k}(0,y). \end{equation*} Since every path in $B_k^+$ visits some $y \ge f(k)$ at time $k$, we have \begin{align*} Z^{0,N}(0, z, B_k^+) &= \int_{y \ge f(k)} Z^{0,k}(0,y,B_k^+) Z^{k,N}(y,z) d y \\ &\le \int_{y \ge f(k)} \exp(-n^{\alpha}) Z^{0,k}(0,y) Z^{k,N}(y,z) dy \\ & \le \exp(-n^{\alpha}) Z^{0,N}(0, z). \end{align*} Therefore, $\mu_{0,z}^{0,N}(B_k^+) \le \exp(-n^{\alpha})$ for all $z$. Integrating over $z$ we obtain \begin{equation*} \mu_{0,\nu}^{0,N}(B_k^+) = \int \mu_{0,z}^{0,N}(B_k^+) \nu(dz) \le \exp(-n^{\alpha}). \end{equation*} Next we look at $C_k^+$. By monotonicity, if $y \le f(2n)$, then \begin{equation*} \mu_{0,y}^{0,2n}(C_k^+) \le \mu_{0,y}^{0,2n}\{\gamma: \gamma_k \ge g(k) \} \le \mu_{0,f(2n)}^{0,2n}\{ \gamma: \gamma_k \ge g(k) \}, \end{equation*} where $g(k) = kx/n+6n^{1-\delta}$. Since \begin{equation*} \big[(0,0), (2n,f(2n))\big]_k - g(k) = 2k n^{-\delta} - 6n^{1-\delta} < -(2n)^{1-\delta}, \end{equation*} we have \begin{equation*} \{ \gamma : \gamma_k \ge g(k) \} \subset \Big\{ \gamma : \max_{j \in [n/2, 3n/2]} \Big|\gamma_j - \big[(0,0),(2n,f(2n) )\big]_j \Big| \ge (2n)^{1-\delta} \Big\}. \end{equation*} Since $A_{0,f(2n)}^{0,2n}$ holds, we have \begin{equation*} \mu_{0,y}^{0,2n}(C_k^+) \le \exp\big(-(2n)^{\alpha}\big) \le \exp( -n^{\alpha} ) \end{equation*} for all $y \le f(k)$. Similarly to the case of $B_k^+$, it follows that $\mu_{0,\nu}^{0,N}(C_k^+) \le \exp(-n^{\alpha})$. The same argument shows that \begin{equation*} \mu_{0,\nu}^{0,N}(B_k^-) \le \exp(-n^{\alpha}), \quad \mu_{0,\nu}^{0,N}(C_k^-) \le \exp( -n^{\alpha}). \end{equation*} We can now write \begin{align*} & \quad \mu_{0,\nu}^{0,N}\big(H(n,x)\big) \\ & \le \sum_{n< k < 3n/2} \left[ \mu_{0,\nu}^{0,N}(C_k^+) + \mu_{0,\nu}^{0,N}(C_k^-) \right] + \sum_{3n/2 \le k \le 2n} \left[ \mu_{0,\nu}^{0,N}(B_k^+) + \mu_{0,\nu}^{0,N}(B_k^-) \right] \\ & \le 2n \exp (-n^{\alpha}), \end{align*} and the lemma follows. \end{proof} The following lemma is a geometric fact. \begin{lemma} \label{lem:geometric-condition-of-staying-in-a-cone} Let $N/2 \ge M$. If a path $\gamma\notin H(n,[\gamma_n])$ (where $[\cdot]$ denotes the integer part) for all $n\in [M,N/2]_{\mathbb Z}$, then \begin{equation*} \gamma^{\text{out}}(n) \subset \mathrm{Co}(n, \gamma_n, Qn^{-\delta}),\quad n\in [M,N/2]\cap{\mathbb Z}, \end{equation*} for a constant $Q$ depending only on $\delta$. \end{lemma} \begin{proof} Let $k = [\log_2 N/n] $ so that $2^k n \le N < 2^{k+1}n$. Since $N \ge 2n$, we have $k \ge 1$. Let $x_l = \gamma_{2^ln}$ and $i_l = [x_l]$ for $ 0 \le l \le k$. If $0 \le l' \le k-1$, then, since $\gamma \not\in H(2^{l'}n, i_{l'})$, we have \begin{equation*} \bigg| \frac{i_{l'}}{2^{l'}n} - \frac{x_{l'+1}}{2^{l'+1}n} \bigg| \le \frac{6(2^{l'}n)^{1-\delta}}{2^{l'+1}n} = 3\cdot 2^{-l'\delta}n^{-\delta} . \end{equation*} Since $|i_{l'+1} - x_{l'+1}| < 1$, we have \begin{equation*} \left| \frac{i_{l'}}{2^{l'}n} - \frac{i_{l'+1}}{2^{l'+1}n}\right| \le 3 \cdot 2^{-l'\delta} n^{-\delta} + \frac{1}{2^{l'+1}n} < 3 \cdot 2^{-l'\delta} n^{-\delta} + 2^{-l'\delta}n^{-\delta} = 4\cdot 2^{-l'\delta}n^{-\delta}. \end{equation*} Summing over $0 \le l' \le l-1$ and using the triangle inequality we obtain \begin{equation}\label{eq:distance-between-diadic-points} \left| \frac{i_0}{n} - \frac{i_{l}}{2^{l}n} \right| \le 4 \cdot n^{-\delta}\sum_{l'=0}^{l-1} 2^{-l'\delta} \le c n^{-\delta} \end{equation} for all $0 \le l \le k$, where $c = 4\sum_{l'=0}^{\infty} 2^{-l'\delta}$. If $n \le m \le 2^kn $, let $l = [\log_2 \frac{m-1}{n}]$ so that $2^ln < m \le 2^{l+1}n$. Since $\gamma \not\in H(2^ln, i_l)$, we have \begin{equation* \left| \frac{i_l}{2^ln} - \frac{\gamma_m}{m} \right| \le \frac{6(2^ln)^{1-\delta}}{m} \le 6 n^{-\delta}. \end{equation*} Combining this with (\ref{eq:distance-between-diadic-points}) we find that \begin{equation} \label{eq:distant-n-to-m-small} \left| \frac{i_0}{n} - \frac{\gamma_m}{m} \right| \le (c+6) n^{-\delta}. \end{equation} If $2^k n < m \le N$, let $\bar{m}= [\frac{m+1}{2}]$ so that $n \le \bar{m}\le 2^kn$ and $\bar{m} < m \le 2\bar{m}$. Then $\gamma \not\in H(\bar{m}, [\gamma_{\bar{m}}])$ implies \begin{equation*} \left| \frac{[\gamma_{\bar{m}}]}{\bar{m}} - \frac{\gamma_m}{m} \right| \le \frac{6(\bar{m})^{1-\delta}}{m} \le 6\bar{m}^{-\delta} \le 6 n^{-\delta}. \end{equation*} Combining this with (\ref{eq:distant-n-to-m-small}) we obtain \begin{align}\label{eq:distance-n-to-m-big} \left| \frac{i_0}{n} - \frac{\gamma_{m}}{m} \right| &\le \left| \frac{i_0}{n} - \frac{\gamma_{\bar{m}}}{\bar{m}}\right| + \left| \frac{\gamma_{\bar{m}}}{\bar{m}} - \frac{[\gamma_{\bar{m}}]}{\bar{m}} \right| + \left| \frac{[\gamma_{\bar{m}}]}{\bar{m}} - \frac{\gamma_m}{m} \right| \\ \nonumber &\le (c+6)n^{-\delta} + \frac{1}{\bar{m}} + 6n^{-\delta} \le (c+13) n^{-\delta}. \end{align} Clearly, \begin{equation}\label{eq:move-initial-point} \left| \frac{\gamma_n}{n} - \frac{i_0}{n} \right| \le n^{-\delta}. \end{equation} Combining (\ref{eq:distant-n-to-m-small}), (\ref{eq:distance-n-to-m-big}), and (\ref{eq:move-initial-point}) we obtain that for every $m\in [n, N]_{{\mathbb Z}}$, \begin{equation*} \left| \frac{\gamma_n}{n} - \frac{\gamma_m}{m} \right| \le (c+14)n^{-\delta}. \end{equation*} This completes the proof, with $Q=c+14$. \end{proof} \bigskip \begin{proof}[Proof of Theorem \ref{thm:delta-straightness-for-polymer-measure}] We fix $\delta \in (0,1/4)$, $v > 0$, $\alpha \in (0,1-2\delta)$, and $\beta \in (0,1-4\delta)$. Let $\alpha' = (\alpha + 1-2\delta)/2$ and $\beta' = (\beta + 1-4\delta)/2$. Then Lemma~\ref{lem:concentration-of-polymer-measure-in-delta-rectangle} implies that for sufficiently large $n$, \begin{equation*} \mathsf{P} \Big( \big(\bar{A}_{0,l}^{0,n}(\alpha')\big)^c\Big) \le \exp(-n^{\beta'}), \quad \forall l \in {\mathbb Z}. \end{equation*} Let $v'= v+ 7$. We have \begin{equation*} \sum_{m \ge n} \sum_{|l| \le v'm+1} \mathsf{P} \Big( \big (\bar{A}_{0,l}^{0,m}(\alpha')\big)^c \Big) \le \sum_{m \ge n} (2v'm+3) \exp(-m^{\beta'}) < e^{-n^{\beta}} \end{equation*} for sufficiently large $n$. Therefore, if we define \begin{equation*} \Omega_1^{(n)} = \bigcap_{m \ge n} \bigcap_{|l| \le v'm+1} \bar{A}_{0,l}^{0,m}(\alpha') \end{equation*} depending only on $v$, $\delta$ and $\alpha$, then $\mathsf{P} (\Omega_1^{(n)}) \ge 1-e^{-n^{\beta}}$ for large $n$. Now let us fix any $\omega \in \Omega_1^{(n)}$. We have then $\omega \in A^{0,m}_{0,x}(\alpha')$ for all $|x| \le v'm$ and $m \ge n$. Let us also fix $N > 2n$ and a terminal measure $\nu$. We want to show that~(\ref{eq:delta-straightness-for-polymer-measure}) holds true. If $(k,z) \in {\rm Co}(v) \cap \{k \ge n\}$, then $(k, [z] ) \in {\rm Co}(v+1)$, so $\mathcal{C}(k,[z],6n^{1-\delta}) \subset {\rm Co}(v') $. Therefore, by Lemma \ref{lem:polymer-measure-of-Hnx-is-small}, for $\omega \in \Omega_1^{(n)}$ and such $(k,z)$, we have \begin{equation*} \mu_{0,\nu}^{0,N}\big( H(k,[z]) \big) \le 2k \exp(-k^{\alpha'}). \end{equation*} By Lemma \ref{lem:geometric-condition-of-staying-in-a-cone}, the set of paths \begin{equation*} \left\{ \gamma : \exists (k,\gamma_k) \in{\rm Co}(v), N/2 \ge k \ge n, \gamma^{\text{out}}(k) \not\subset {\rm Co}(k,\gamma_k,Qk^{-\delta}) \right\} \end{equation*} is contained in \begin{equation*} \bigcup_{N/2\ge k\ge n, y = [z], |z| \le kv } H(k,y). \end{equation*} Therefore, \begin{align*} &\quad \mu_{0,\nu}^{0,N} \left\{ \gamma : \exists (k,\gamma_k) \in \mathrm{Co}(v), N/2 \ge k \ge n, \gamma^{\text{out}}(k) \not\subset \mathrm{Co}(k,\gamma_k,Qk^{-\delta}) \right\} \\ &\le \sum_{N/2 \ge k\ge n, y = [z], |z| \le kv} \mu_{x,\nu}^{0,N}\big( H(k,y) \big) \\ &\le \sum_{k=n}^{\infty} (2kv+2) \cdot 2k \exp(-k^{\alpha'})\le \exp(-n^{\alpha}) \end{align*} for sufficiently large $n$. This completes the proof. \end{proof} \subsection{Tightness of polymer measures}\label{sec:compactness-of-polymer-measures} In this section, we will use Theorem~\ref{thm:delta-straightness-for-polymer-measure} on $\delta$-straightness to prove tightness of families of polymer measures constructed for a fixed endpoint and a well-behaved sequence of terminal measures. Let $(m,x)\in\Z\times{\mathbb R}$. For a sequence $(n_k)$ and a family of Borel probability measures $(\nu_k)$ on ${\mathbb R}$, we say that the family of polymer measures $ \big( \nu_{x,\nu_k}^{m,n_k} \big)$ is tight if for every $n\ge m$ and every $\varepsilon>0$, there is a compact set $K \subset {\mathbb R}^{n-m+1}$ such that for all $n_k > n$, \begin{equation*} \mu_{x,\nu_k}^{m,n_k} \pi_{m,n}^{-1} ({\mathbb R}^{n-m+1}\setminus K) \le \varepsilon. \end{equation*} \begin{theorem} \label{thm:tightness-of-polymer-measure} There is a full measure set $\Omega'$ such that for all~$\omega \in \Omega'$ the following holds: if a sequence $(n_k)$ and a family of probability measures $(\nu_k)$ satisfy \begin{equation} \label{eq:tightness-for-terminal-measures} \lim_{c \to \infty} \sup_k \nu_k \big( [-cn_k,cn_k]^c ) = 0, \end{equation} then for all $(m,x)\in{\mathbb Z}\times{\mathbb R}$, $\big(\mu_{x,\nu_k}^{m,n_k}\big)$ is tight. \end{theorem} We will derive this theorem from the following result: \begin{theorem} \label{thm:all-space-time-point-compactness} There is a full measure set $\Omega'$ such that for every $\omega \in \Omega'$ the following holds: if $(m,x) \in {\mathbb Z} \times {\mathbb R}$, $v'\in {\mathbb R}$, and $0\le u_0 <u_1$, then there is a random constant \begin{equation*} n_0= n_0\big(\omega,m,[x], [|v'|+u_1], [(u_1-u_0)^{-1}] \big) \end{equation*} (where $[\cdot]$ denotes the integer part) such that \begin{equation} \label{eq:remote-control-all-space-time-all-space-time} \mu_{x,\nu}^{m,N} \big\{ \gamma: |\gamma_{m+n} - v'n| \ge u_1n \big\} \le 4 \nu \big( [ (v'-u_0)N, (v'+u_0)N]^c \big) + 6e^{-\sqrt{n}} \end{equation} and \begin{multline} \label{eq:compactness-control-by-terminal-measure-all-space-time} \mu_{x,\nu}^{m,N} \big\{ \gamma: \max_{1 \le i \le n}|\gamma_{m+i} - v'i| \ge (u_1+R+1)n \big\} \\ \le 4 \nu \big( [ (v'-u_0)N, (v'+u_0)N]^c \big) + 8e^{-\sqrt{n}}, \end{multline} hold true for any terminal measure $\nu$ and $(N-m)/2 \ge n \ge n_0$. Here, we use~$R$ that has been introduced in Lemma \ref{lem:large-deviation-for-Z-bar}. \end{theorem} \begin{proof}[Derivation of Theorem~\ref{thm:tightness-of-polymer-measure} from Theorem~\ref{thm:all-space-time-point-compactness}] We take $\Omega'$ from the statement of Theorem~\ref{thm:all-space-time-point-compactness} and fix an arbitrary $\omega\in\Omega'$. Given any $\varepsilon>0$, by (\ref{eq:tightness-for-terminal-measures}), there is $c$ such that $\nu_{k}\big( [-cn_k,cn_k]^c \big) \le \varepsilon$ for all $k$. Choosing $v'=0$, $u_0=c$, $u_1=2c$ in Theorem~\ref{thm:all-space-time-point-compactness}, we see that if \begin{equation*} n \ge n_0\big(\omega, m,[x], [2c], [c^{-1}] \big) \vee \ln^2 \varepsilon, \end{equation*} then, due to (\ref{eq:compactness-control-by-terminal-measure-all-space-time}), \begin{equation*} \mu_{x,\nu_k}^{m,n_k} \Bigl\{ \gamma: \max_{m \le i \le m+n} |\gamma_i| \ge (2c+R+1)n \Bigr\} \le 4 \nu_k \big( [-cn_k, cn_k]^c \big) + 8 e^{-\sqrt{n}} \le 12 \varepsilon \end{equation*} for all $n_k \ge m+ 2n$, and tightness follows. \end{proof} Now we turn to the proof of Theorem \ref{thm:all-space-time-point-compactness}. Using the constant $R$ introduced in Lemma \ref{lem:large-deviation-for-Z-bar}, we define the events \begin{equation*} B_{x,y}^{m,n} = \Big\{ \mu_{x,y}^{m,n} \big\{ \gamma: \max_{m\le k \le n} \big| \gamma_k - [(m,x),(n,y)]_k\big| \ge (R+1)(n-m) \big\} \le 2^{-n+1} \Big\} \end{equation*} and \begin{equation*} \bar{B}_{p,q}^{m,n} = \bigcap \big\{ B_{x,y}^{m,n}: x \in [p,p+1], y \in [q,q+1]\}. \end{equation*} These are measurable events since, by continuity, the intersection on the right-hand side may be restricted to rational points $x,y$. Also, for fixed $m$ and $n$, all $\bar{B}_{p,q}^{m,n}$ have the same probability. \begin{lemma} \label{lem:uniform-small-probability-outside-a-linear-box} If $n$ is sufficiently large, then \begin{equation*} \mathsf{P} \Big( \big( \bar{B}_{0,0}^{0,n}\big)^c \Big) \le 2e^{-r_1n}. \end{equation*} \end{lemma} \begin{proof} The lemma follows from Lemma \ref{lem:large-deviation-for-Z-bar} and the fact that $\mu_{0,0}^{0,n}\preceq\mu_{x,y}^{0,n}\preceq \mu_{1,1}^{0,n}$ for $x,y\in[0,1]$. \end{proof} \begin{lemma} \label{lem:delta-straighness-imply-compactness} Suppose $\delta \in (0,1/4)$, $n \in {\mathbb N}$, $v' \in {\mathbb R}$, $0\le v_0<v_1$, and $v_1 > v_0 + Qn^{-\delta}$. If $\omega$ satisfies the following properties: \begin{enumerate} \item\label{item:6} $\omega \in A_{0,z}^{0,m}(\frac{1}{2})$ for all $m \ge n$ and $|z-v'm| \le v_0m$, \item\label{item:7} $\omega \in B_{0,z}^{0,m}$ for all $m \ge n$ and $|z-v'm| \le v_1m$, \item\label{item:5} $\omega \in \Omega_1^{(n)}(\delta,|v'|+v_1+R+2,\frac{1}{2})$, \end{enumerate} then for any $N \ge 2n$ and terminal measure $\nu$, \begin{equation*} \mu_{0,\nu}^{0,N} \big\{ \gamma: |\gamma_n-v'n| \ge v_1n \big\} \le 2\nu \big([(v'-v_0)N, (v'+v_0)N]^c\big) + 3e^{-\sqrt{n}}. \end{equation*} \end{lemma} \begin{proof} We will only give a proof for the case of $v'=0$. The extension to a general $v'$ is straightforward. Let $v_2 = v_1 + R + 2$ and fix $N \ge 2n$. Define $T=T(\gamma)=\inf\{ m \ge n: (m,\gamma_m) \in {\rm Co}(v_1)\}\wedge N$. Since $|\gamma_n| \ge nv_1$ implies $(T-1,\gamma_{T-1}) \not\in {\rm Co}(v_1)$, we can partition the set of paths $B = \{ \gamma: |\gamma_n| \ge nv_1 \}$ into the union of $A_i$, where \begin{align*} A_1 &= \big\{ \gamma: T \le [N/2],\ (T-1,\gamma_{T-1}) \in {\rm Co}(v_2)\setminus {\rm Co}(v_1),\ \gamma_N \in [- v_0N, v_0N] \big\}, \\ A_2 &= \big\{ \gamma: T \le [N/2],\ (T-1,\gamma_{T-1}) \in {\rm Co}(v_2)\setminus{\rm Co}(v_1),\ \gamma_N \in [-v_0N, v_0N]^{c} \big\}, \\ A_3 &= \big\{ \gamma: T \le [N/2],\ (T-1,\gamma_{T-1}) \not\in {\rm Co}(v_2) \big\}, \\ A_4 &= \big\{ \gamma: T > [N/2],\ \gamma_N \in [-v_0N, v_0N] \big\},\\ A_5 &= \big\{ \gamma: T > [N/2],\ \gamma_N \in [-v_0N, v_0N]^c \big\}. \end{align*} Let us estimate weights assigned to sets $A_i$ by the polymer measure. Inequalities \begin{equation} \label{eq:A2-and-A5} \mu_{0,\nu}^{0,N}(A_2) \le \nu \big( [-v_0N, v_0N]^c \big), \quad \mu_{0,\nu}^{0,N}(A_5) \le \nu \big( [-v_0N, v_0N]^c \big) \end{equation} are obvious. Since \begin{equation*} A_1 \subset \left\{ \gamma : \exists (k,\gamma_k) \in {\rm Co}(v_2),\ N/2 \ge k \ge n,\ \gamma^{\text{out}}(k) \not\subset {\rm Co}(k,z,Qk^{-\delta}) \right\} \end{equation*} (the condition is satisfied by $k=T-1$), and $\omega \in \Omega_1^{(n)}(\delta,v_2,\frac{1}{2})$, we have \begin{equation} \label{eq:A1} \mu_{0,\nu}^{0,N}(A_1) \le e^{-\sqrt{n}} . \end{equation} Since $A_3 \subset \big\{|\gamma_T| \le v_1T \big\}$ and for $n\ge R+2$ and $|z| \le v_1m$ we have \begin{align*} \mu_{0,z}^{0,m}(A_3 \cap \{T = m\}) &\le \mu_{0,z}^{0,m} \Big\{ \gamma: \max_{1 \le k \le m} \big| \gamma_k - [(0,0),(m,z)]_k \big| \ge (R+2)(m-1) \Big\} \\ &\le \mu_{0,z}^{0,m} \Big\{ \gamma: \max_{1 \le k \le m} \big| \gamma_k - [(0,0),(m,z)]_k \big| \ge (R+1)m \Big\} \end{align*} (the condition is satisfied for $k=T-1$), $\omega \in B_{0,z}^{0,m}$ will imply \begin{equation*} \mu_{0,z}^{0,m} \bigl( A_3 \cap \{ T=m \} \bigr) \le 2^{-m+1}, \end{equation*} and hence \begin{equation*} \mu_{0, \nu}^{0,N} \bigl( A_{3} \cap \{ T=m \} \bigr) \le 2^{-m+1}. \end{equation*} Therefore, \begin{equation} \label{eq:A3} \mu_{0,\nu}^{0,N}(A_3) \le \sum_{m=n+1}^{[N/2]} 2^{-m+1} \le 2^{-n+1}. \end{equation} Since $A_4 \subset \big\{ |\gamma_N| \le v_0N \big\}$ and $T > [N/2]$ implies that $([N/2],\gamma_{[N/2]}) \not\in {\rm Co}(v_1)$, we have \begin{equation*} \mu_{0,z}^{0,N} (A_4) \le \mu_{0,z}^{0,N} \Big\{ \gamma: \max_{N/4 \le k \le 3N/4} |\gamma_k - [(0,0),(N,z)]_k| \ge \frac{Q}{3}N^{1-\delta} \Big\} \end{equation*} for $|z| \le v_0N$ (the condition is satisfied for $k = [N/2]$ ). Therefore, $\omega \in A^{0,N}_{0,z}(\frac{1}{2})$ will imply (due to the choice of~$Q$ in the proof of Lemma~\ref{lem:geometric-condition-of-staying-in-a-cone}) \begin{equation*} \mu_{0,z}^{0,N} (A_4 ) \le e^{-\sqrt{N}} \le e^{-\sqrt{n}} \end{equation*} and hence \begin{equation} \label{eq:A4} \mu_{0,\nu}^{0,N}(A_4) = \int_{-v_0N}^{v_0N} \mu_{0,z}^{0,N}(A_4) \nu(dz) \le e^{-\sqrt{n}}. \end{equation} The conclusion of the lemma follows from (\ref{eq:A2-and-A5}), (\ref{eq:A1}), (\ref{eq:A3}), and (\ref{eq:A4}). \end{proof} \begin{lemma} \label{lem:sets-for-remote-control-to-holds} Let $\delta \in (0,1/4)$, $v > 0$ and $(m,q) \in {\mathbb Z} \times {\mathbb Z}$. Then there is an event $\Omega_2^{(n)} = \Omega_2^{(n)}(\delta,v, m, q)$ with the following properties. \begin{enumerate} \item\label{item:12} For all $\beta \in (0, 1-4\delta)$ and sufficiently large $n$ (depending on $\beta$), we have \begin{equation} \label{eq:Omega-2-small-probability} \mathsf{P} \big( (\Omega_{2}^{(n)} )^{c} \big ) \le e^{-n^{\beta}}. \end{equation} \item\label{item:15} Let $v' \in {\mathbb R}$, $0 \le v_0<v_1$. If $\omega \in \Omega^{(n)}_2$, $|v'| + v_1\le v$, $Qn^{-\delta} < v_1-v_0$, $N/2 \ge s \ge n$, then for any terminal measure $\nu$, \begin{multline} \label{eq:remote-control} \mu_{q,\nu}^{m,m+N} \big\{ \gamma: |\gamma_{m+s} - q - v'n| \ge v_1n \big\} \\ \le 2\nu \big( [ q+(v'-v_0)N, q+(v'+v_0)N]^c \big) + 3e^{-\sqrt{s}} \end{multline} and \begin{multline} \label{eq:compactness-control-by-terminal-measure} \mu_{q,\nu}^{m,m+N} \big\{ \gamma: \max_{1 \le i \le s}|\gamma_{m+i} - q - v'i| \ge (v_1+R+1)s \big\} \\ \le 2 \nu \big( [ q+(v'-v_0)N, q+(v'+v_0)N]^c \big) + 4e^{-\sqrt{s}}. \end{multline} \end{enumerate} \end{lemma} \begin{proof} Without loss of generality, we can assume $(m,q) = (0,0)$. Let \begin{equation*} \Omega^{(n)}_2 = \Omega_1^{(n)}(\delta,v+R+2,1/2) \cap \bigg[ \bigcap_{s \ge n, |p| \le vs+1} \Big( \bar{A}_{0,p}^{0,s}(1/2) \cap \bar{B}_{0,p}^{0,s} \Big) \bigg]. \end{equation*} We claim that such events have the desired properties. In fact, (\ref{eq:Omega-2-small-probability}) follows from Theorem \ref{thm:delta-straightness-for-polymer-measure}, Lemma \ref{lem:concentration-of-polymer-measure-in-delta-rectangle}, and Lemma \ref{lem:uniform-small-probability-outside-a-linear-box}; (\ref{eq:remote-control}) follows from Lemma \ref{lem:delta-straighness-imply-compactness}. It remains to show that if $\omega \in \Omega^{(n)}_2$, then (\ref{eq:compactness-control-by-terminal-measure}) holds. Fix $s \ge n$ and assume $\omega \in \Omega^{(n)}_2$. Let \begin{equation*} A = \big\{ \gamma : \ \max_{1 \le i \le s} |\gamma_i-v'i| \ge (v_1+R+1)s \big\}. \end{equation*} If $|z-v's| \le v_1s$, then \begin{equation*} \mu_{0,z}^{0,s} (A ) \le \mu_{0,z}^{0,s} \{ \gamma: \max_{1 \le i \le s} | \gamma_i - [(0,0), (s,z)]_i | \ge (R+1)s\}. \end{equation*} Since $\omega \in B_{0,z}^{0,s}$, we have $\mu_{0,z}^{0,s} (A) \le 2^{-s+1}$ for all $|z-v's| \le v_1s $. Therefore, \begin{equation*} \mu_{0,\nu}^{0,N} \bigl( A \cap \{ |\gamma_s - v's| \le v_1s \} \bigr) \le 2^{-s+1} \le e^{-\sqrt{s}}. \end{equation*} Then (\ref{eq:compactness-control-by-terminal-measure}) follows from this and (\ref{eq:remote-control}). \end{proof} \begin{proof}[Proof of Theorem \ref{thm:all-space-time-point-compactness}] We fix $\delta \in (0, \frac{1}{4})$. By (\ref{eq:Omega-2-small-probability}) and the Borel--Cantelli Lemma, with probability one, for all $(m,q) \in {\mathbb Z} \times {\mathbb Z}$ and $M \in {\mathbb N}$, $\Omega_2^{(n)}(\delta, M, m, q)$ hold for $n > n_1$, where $n_1=n_1(\omega, m,q, M)$ is a random constant depending on $m,q$ and $M$. Fix $0 \le u_0 < u_{1}$. Let \begin{equation*} v_0 = \frac{2u_0 + u_1}{3}, \quad v_1 = \frac{u_0 + 2u_1}{3}. \end{equation*} Suppose $x \in [q,q+1]$ for some $q \in {\mathbb Z}$ and $ |v'| + u_1 \le M$ for some $M \in {\mathbb N}$. Lemma~\ref{lem:main-monotonicity} implies $\mu_{q,\nu}^{m,N} \preceq \mu_{x,\nu}^{m,N} \preceq \mu_{q+1,\nu}^{m,N}$. Combined with Lemma \ref{lem:sets-for-remote-control-to-holds}, we see that if \begin{equation*} n \ge n_1(\omega,m,q,M) \vee n_1(\omega,m,q+1,M) \vee \Bigl( \frac{3Q}{|u_1-u_0|} \Bigr)^{\frac{1}{\delta}} \vee \bigg( \frac{3(|q|+1 + M|m|)}{|u_1-u_0|} + |m| \bigg), \end{equation*} then for any $N - m \ge 2n $ and any terminal measure $\nu$, we have \begin{align*} &\mu_{x,\nu}^{m,N} \big\{ \gamma: |\gamma_{m+n} - v'n| \ge u_1n \big\} \\ \le& \sum_{p=q}^{q+1} \mu_{p,\nu}^{m,N} \big\{ \gamma: |\gamma_{m+n} - p-v'n| \ge v_1n) \big\} \\ \le & 2 \sum_{p=q}^{q+1} \nu \bigl( [p+(v'-v_0)(N-m), p+(v' + v_0) (N-m)]^c \bigr) + 6e^{-\sqrt{n}} \\ \le &4 \nu \bigl( [(v'-u_0)N, (v'+u_0)N]^c \bigr) + 6e^{-\sqrt{n}}. \end{align*} Here, the first inequality uses $u_1n \ge v_{1}n + q+1$, which is implied by $n \ge \frac{|q|+1}{u_1-v_1} = \frac{3(|q|+1)}{|u_1-u_0|}$, the third inequality uses $q+1 + (v'+v_0)(N-m) \ge (v'+u_0) N$, i.e., $q+1+(v_0-u_0)N \ge m (v'+v_0)$, which is implied by $N-m \ge n \ge \frac{mM + |q|+1}{v_0-u_0} = \frac{3(mM+|q|+1)}{|u_1-u_0|}$. Similarly, \begin{align*} &\quad \mu_{x,\nu}^{m,N} \big\{ \gamma: \max_{1 \le i \le n}|\gamma_{m+i} - v'i| \ge (u_1+R+1)n \big\} \\ & \le \sum_{p=q}^{q+1} \mu_{p,\nu}^{m,N} \big\{ \gamma: \max_{1 \le i \le n}|\gamma_{m+i} - p-v'i| \ge (v_1+R+1)n \big\} \\ & \le 2\sum_{p=q}^{q+1} \nu \bigl( [p + (v'-v_0)(N-m), p+(v'+v_0)(N-m)]^c \bigr) + 8e^{-\sqrt{n}} \\ &\le 4 \nu \bigl( [(v'-u_0)N, (v'+u_0)N]^c \bigr) + 8e^{-\sqrt{n}}. \end{align*} This completes the proof. \end{proof} \section{Infinite-volume polymer measures } \label{sec:infinite-volume-polymer-measure} In this section, we prove most claims of the Theorem~\ref{thm:thermodynamic-limit} on thermodynamic limits of polymer measures. We prove the existence and uniqueness part, and we prove that finite time horizon polymer measures converge to the infinite volume polymer measures in the sense of weak convergence of finite-dimensional distributions. We will prove the stronger total variation convergence in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions}. We begin with the following useful statement: \begin{lemma}\label{lem:limit-of-polymers-is-a-polymer} For all $\omega\in\Omega$, if a sequence of polymer measures has a weak limit, then the limiting measure is also a polymer measure. \end{lemma} \bpf It is sufficient to prove the statement of the lemma for finite volume polymer measures. We need to prove that $\mu^{m,n}_{x,\nu_k}$ weakly converges to $\mu^{m,n}_{x,\nu}$ if $m,n\in{\mathbb Z}$, $x\in{\mathbb R}$, and $(\nu_k)$ is a sequence of distributions on~${\mathbb R}$, weakly convergent to a distribution $\nu$. It suffices to check that if $f(x_{m+1},\ldots,x_n)=f_{m+1}(x_{m+1})\ldots f_{n-1}(x_{n-1})f_{n}(x_{n})$ for continuous nonnegative functions $f_{m+1},\ldots,f_n$ with bounded support, then \begin{multline*} \lim_{k\to\infty}\int\mu^{m,n}_{x,\nu_k}(x_m,\ldots, dx_n)f(x_{m+1},\ldots,x_n)\\=\int\mu^{m,n}_{x,\nu}(dx_m,\ldots, dx_n)f(x_{m+1},\ldots,x_n). \end{multline*} Since \[ \int\mu^{m,n}_{x,\nu}(dx_m,\ldots, dx_n)f(x_{m+1},\ldots,x_n)=\int \nu(dx_n) G(x_n), \] where \[ G(x_n)=\int \mu_{x,x_n}^{m,n}(dx_m,\ldots, dx_n)f(x_{m+1},\ldots,x_n), \] we need to show that $G$ is a continuous function. The latter follows from the definition of $\mu_{x,x_n}^{m,n}$, continuity of $Z^{m,n}_{x,x_n}$ (see Lemma~\ref{lem:smoothness-of-partition-function}) and $g(x_{n}-x_{n-1}) f_n(x_{n})$ with respect to~$x_n$, and the bounded convergence theorem. {{\hfill $\Box$ \smallskip}} In addition to the terminology and notation from Section~\ref{sec:polymers}, we say that LLN with slope $v\in{\mathbb R}$ holds {\it for an increasing sequence of times} $(n_k)$ and a sequence of Borel measures $(\nu_k)$ on ${\mathbb R}$ if for all $\delta>0$, \[ \lim_{k\to\infty} \nu_k([(v-\delta)n_k,(v+\delta) n_k])=1. \] \begin{lemma}\label{lem:existence-of-polymer-measures-and-convergence-along-subsequences} For every $\omega \in \Omega'$ (introduced in Theorem \ref{thm:all-space-time-point-compactness}) the following holds true. For any $(m,x)\in {\mathbb Z} \times{\mathbb R}$, for any $v\in{\mathbb R}$ , any time sequence $(n_k)$ and any sequence of measures $(\nu_k)$ satisfying LLN with slope $v$, there is an increasing subsequence $(k_i)_{i\in{\mathbb N}}$ such that $\mu^{m,n_{k_i}}_{x,\nu_{k_i}}$ converges in the sense of weak convergence of finite-dimensional distributions to a measure $\mu$ on $S_x^{m,+\infty}$. The limiting measure $\mu$ is a polymer measure supported on $S_x^{m,\infty}(v)$. \end{lemma} \begin{proof} Since $(\nu_k)$ satisfies LLN with slope $v$, (\ref{eq:tightness-for-terminal-measures}) is satisfied. By Theorem \ref{thm:tightness-of-polymer-measure}, the sequence $(\mu_{x,\nu_k}^{m,n_k})$ forms a tight family, so by the Prokhorov theorem, there is a converging subsequence of this sequence. Let $\mu$ be the limiting measure of some subsequence $(\mu_{x,\nu_{k_i}}^{m,n_{k_i}})$. It is an infinite volume polymer measure due to Lemma~\ref{lem:limit-of-polymers-is-a-polymer}. Let us prove that for every $\varepsilon > 0$, \begin{equation} \label{eq:tails-summable} \sum_{n=1}^{\infty} \mu\pi^{-1}_{m+n}\big([(v-\varepsilon)n , (v+\varepsilon) n]^c\big) < \infty. \end{equation} The Borel--Cantelli lemma will imply then that $\mu$ is supported on $S_x^{m,+\infty}(v)$. Fixing $\varepsilon>0$, for sufficiently large $n$ and $n_{k_i}-m > 2n$, we derive from (\ref{eq:remote-control-all-space-time-all-space-time}): \begin{equation*} \mu_{x,\nu_{k_i}}^{m,n_{k_i}} \pi^{-1}_{m+n} \big([(v-\varepsilon)n, (v+\varepsilon)n]^c \big) \le 4\nu_{k_i} \big( [(v-\varepsilon/2)n_{k_i}, (v+\varepsilon/2) n_{k_i}]^c \big) + 6e^{-\sqrt{n}}. \end{equation*} Since $(\nu_k)$ satisfies LLN with slope $v$, taking the limit $k_i \to \infty$ and using the weak convergence of finite-dimensional distributions of $(\mu_{x,\nu_{k_i}}^{n,n_{k_i}})$, we find \begin{equation*} \mu \pi^{-1}_{m+n} \big( [ (v-\varepsilon)n, (v+\varepsilon)n]^c \big) \le 6 e^{-\sqrt{n}}. \end{equation*} Therefore \eqref{eq:tails-summable} holds, and the proof is complete. \end{proof} \bigskip Our next goal is uniqueness of a polymer measure with given endpoint and slope. Let $m\in{\mathbb Z}$ and let $\mu_1$ and $\mu_2$ be two measures on $S^{m,+\infty}$. We say that $\mu_1$ is stochastically dominated by $\mu_2$ if $\mu_1\pi_{m,n}^{-1}$ is stochastically dominated by $\mu_2\pi_{m,n}^{-1}$ for all finite $n>m$. \begin{lemma} \label{lem:infinite-volume-monotonicity-wrt-slope} Let $v_1<v_2$ and $(m,x)\in\Z\times{\mathbb R}$. If $\mu_1$ and $\mu_2$ are polymer measures on $S_x^{m,+\infty}$ satisfying LLN with slopes $v_1$ and $v_2$, respectively, then $\mu_2$ stochastically dominates $\mu_1$. \end{lemma} To prove this lemma, we need the following obvious auxiliary statement. \begin{lemma}\label{lem:convergence-preserves-dominance} Suppose $(\mu_1^{k})_{k\in{\mathbb N}}$ and $(\mu_2^{k})_{k\in{\mathbb N}}$ are sequences of probability measures converging weakly to probability measures $\mu_1$ and~$\mu_2$, respectively, and such that $\mu_1^k$ is dominated by $\mu_2^k$ for all $k\in{\mathbb N}$. Then $\mu_1$ is dominated by $\mu_2$. \end{lemma} \bpf[Proof of Lemma~\ref{lem:infinite-volume-monotonicity-wrt-slope}] Let us take any $\delta>0$ satisfying $v_1+\delta< v_2-\delta$, denote \[ \mu_{i,k}:=\mu_i\pi_{k}^{-1},\quad i=1,2,\quad k>m, \] and introduce $\mu_{i,k,\delta}$ as $\mu_{i,k}$ conditioned on $[(v_i-\delta)k,(v_i+\delta) k]$. Then $\mu_{1,k,\delta}$ is dominated by $\mu_{2,k,\delta}$. Using Lemma~\ref{lem:main-monotonicity} on monotonicity, we obtain that $\mu_{x,\mu_{1,k,\delta}}^{m,k}$ is dominated by $\mu_{x,\mu_{2,k,\delta}}^{m,k}$. Therefore, $\mu_{x,\mu_{1,k,\delta}}^{m,k}\pi_{m,r}^{-1}$ is dominated by $\mu_{x,\mu_{2,k,\delta}}^{m,k}\pi_{m,r}^{-1}$, for any $r$ between $m$ and $k$. Since, in addition, the LLN assumption implies \[ \lim_{k\to\infty}\|\mu_{i}\pi_{m,r}^{-1}-\mu_{x,\mu_{i,k,\delta}}^{m,k}\pi_{m,r}^{-1}\|_{TV}=\lim_{k\to\infty}\|\mu_{i,k}-\mu_{i,k,\delta}\|_{TV}=0,\quad i=1,2, \] Lemma~\ref{lem:convergence-preserves-dominance} implies that $\mu_{1}\pi_{m,r}^{-1}$ is dominated by $\mu_{2}\pi_{m,r}^{-1}$. {{\hfill $\Box$ \smallskip}} \begin{lemma} \label{lem:uniqueness-and-convergence-at-rationals}Let $v\in{\mathbb R}$. Then there is a set $\tilde \Omega_v$ of probability~$1$ such that the following holds on $\tilde \Omega_v$: \begin{enumerate}[1.] \item \label{it:uniqueness-at-rationals} For every point $(m,x)\in{\mathbb Z}\times{\mathbb Q}$, the set $\mathcal{P}_x^{m,+\infty}(v)$ of all polymer measures on $S_x^{m,+\infty}$ satisfying SLLN with slope $v$, contains exactly one element that we denote by~$\mu_{x}^{m,+\infty}(v)$. \item \label{it:convergence-of-polymer-measures-at-rationals} For every point $(m,x)\in{\mathbb Z}\times{\mathbb Q}$ and for every sequence of measures $(\nu_n)$ satisfying LLN with slope $v$, $\mu_{x,\nu_n}^{m,n}$ weakly converges to $\mu_{x}^{m,+\infty}(v)$. \end{enumerate} \end{lemma} This lemma is weaker than Theorem~\ref{thm:thermodynamic-limit} in two ways: its statements hold only for rational spatial locations, and only weak convergence is claimed. We study the irrational points later in this section, and prove the total variation convergence in Section~\ref{sec:global-solutions-and-ratios-of-partition-functions}. \bpf Let us fix a point $(m,x)$. By Lemma~\ref{lem:existence-of-polymer-measures-and-convergence-along-subsequences}, for each $v$, the set $\mathcal{P}_x^{m,+\infty}(v)$ is non-empty. For any $\mu\in \mathcal{P}_x^{m,+\infty}(v)$ and any $k>m$, the measure $\mu\pi_{k}^{-1}$ is equivalent to Lebesgue measure (in the sense of absolute continuity), so for any $\alpha\in(0,1)$ the quantile $q_\alpha(\mu)$ at level $\alpha$ is uniquely defined by~$\mu\pi_{k}^{-1}(-\infty, q_\alpha(\mu)]=\alpha$. So let us define \begin{align*} q_\alpha^-(v)&=\inf\{q_\alpha(\mu): \mu\in \mathcal{P}_x^m(v)\},\\ q_\alpha^+(v)&=\sup\{q_\alpha(\mu): \mu\in \mathcal{P}_x^m(v)\}. \end{align*} Let us prove that with probability 1, $q_\alpha^-=q_\alpha^+$. Due to Lemma~\ref{lem:infinite-volume-monotonicity-wrt-slope}, if $v_1<v_2$, then $q_\alpha^-(v_1)\le q_\alpha^+(v_1)\le q_\alpha^-(v_2)\le q_\alpha^+(v_2)$. Therefore, with probability~$1$, there may be at most countably many nonempty intervals $I_\alpha(v)=(q_\alpha^-(v),q_\alpha^+(v))$. On the other hand, space-time shear transformations map polymer measures into polymer measures (on finite or infinite paths), so $\mathsf{P}\{I_\alpha(v)\ne\emptyset\}=p$ does not depend on~$v$. Therefore, we can apply arguments similar to those in~\cite{kickb:bakhtin2016} and going back to Lemma~6 in~\cite{HoNe3}. We take an arbitrary probability density $f$ on ${\mathbb R}$ and write \[ p=\int_{\mathbb R} \mathsf{P}\{I_\alpha(v)\ne\emptyset\}f(v)dv =\int_{\mathbb R} \mathsf{E} {\mathbf{1}}_{\{I_\alpha(v)\ne\emptyset\}}f(v)dv =\mathsf{E} \int_{\mathbb R} {\mathbf{1}}_{\{I_\alpha(v)\ne\emptyset\}}f(v)dv =0, \] since $I_\alpha(v)\ne\emptyset$ can be true for at most countably many $v$. So, for any $v\in{\mathbb R}$, $\mathsf{P}\{I_\alpha(v)\ne\emptyset\}=0$. This immediately implies that for every $v\in{\mathbb R}$, \[ \mathsf{P}\{q_\alpha^-(v)= q_\alpha^+(v)~\text{for all}~\alpha\in{\mathbb Q} \}=1. \] So, for any $\mu_1,\mu_2\in \mathcal{P}_x^{m,+\infty}(v)$, the rational quantiles of~$\mu_1\pi_{k}^{-1}$ and~$\mu_2\pi_{k}^{-1}$ coincide. Therefore, $\mu_1\pi_{k}^{-1}=\mu_2\pi_{k}^{-1}$. In turn, this implies $\mu_1\pi_{m,k}^{-1}=\mu_1\pi_{m,k}^{-1}$. Since this is true for all $k$, we conclude that $\mu_1=\mu_2$. So we have proved that for a fixed point $(m,x)\in\Z\times{\mathbb R}$, with probability~$1$, a polymer measure with specified asymptotic slope is unique. We denote that measure by $\mu_x^{m,+\infty}(v)$. By countable additivity, this uniqueness statement holds true for all $(m,x)\in{\mathbb Z}\times{\mathbb Q}$ at once on a common set $\tilde\Omega_v$ of measure 1, and part~\ref{it:uniqueness-at-rationals} is proved. To prove the second part, we fix any $\omega\in\tilde \Omega_v$ and use a compactness argument. Lemma~\ref{lem:existence-of-polymer-measures-and-convergence-along-subsequences} implies that from any subsequence $(\mu_{x,\nu_n}^{m,n})$ one can choose a convergent subsubsequence. Part~\ref{it:uniqueness-at-rationals} of this lemma implies that all these partial limits must coincide with $\mu_x^{m,+\infty}(v)$. Therefore, the entire sequence converges to $\mu_x^{m,\infty}(v)$, which completes the proof of the lemma. {{\hfill $\Box$ \smallskip}} \begin{lemma} Let $v\in{\mathbb R}$. On $\tilde\Omega_v$, for every $m\in{\mathbb Z}$ and points $x_1,x_2\in{\mathbb Q}$ satisfying $x_1<x_2$, $\mu_{x_1}^{m,+\infty}(v)$ is dominated by $\mu_{x_2}^{m,+\infty}(v)$. \end{lemma} \bpf By Lemma~\ref{lem:uniqueness-and-convergence-at-rationals}, for $i\in\{1,2\}$, the sequence of measures $(\mu_{x_i,vn}^{m,n})_{n>m}$ converges to $\mu_{x_i}^{m,+\infty}(v)$ as $n\to\infty$. Since for every $n$, $\mu_{x_1,vn}^{m,n}$ is dominated by $\mu_{x_2,vn}^{m,n}$, the limiting measures are also related by stochastic dominance. {{\hfill $\Box$ \smallskip}} \begin{lemma}\label{lem:squeezing-any-endpoint-of-one-sided-polymer-measures-between-rational-ones} For every $v$, the following holds on $\Omega_v$. For every $m\in{\mathbb Z}$, every $x\in{\mathbb R}$ and every $x_-,x_+\in{\mathbb Q}$ such that $x_-<x<x_+$, every measure in $\mathcal{P}_x^{m,+\infty}(v)$ is dominated by the (unique) measure $\mu_{x_+}^{m,+\infty}(v)$ in $\mathcal{P}_{x_+}^{m,+\infty}(v)$ and dominates the (unique) measure $\mu_{x_-}^{m,+\infty}(v)$ in $\mathcal{P}_{x_-}^{m,+\infty}(v)$. \end{lemma} \bpf We take an arbitrary measure $\mu\in\mathcal{P}_x^{m,+\infty}(v)$ and denote $\nu_n=\mu\pi_{n}^{-1}$, $n>m$. Since $\nu_n$ satisfy LLN with slope $v$, $\mu_{x_-,\nu_n}^{m,n}$ and $\mu_{x_+,\nu_n}^{m,n}$ converge, by Lemma~\ref{lem:uniqueness-and-convergence-at-rationals}, to $\mu_{x_-}^{m,+\infty}(v)$ and $\mu_{x_+}^{m,+\infty}(v)$, respectively. Since $\mu_{x,\nu_n}^{m,n}$ coincides with $\mu\pi_{m,n}^{-1}$, the lemma follows from the dominance relation on the pre-limiting measures. {{\hfill $\Box$ \smallskip}} So now we know that for any $x$, the measures in $\mathcal{P}_x^{m,+\infty}(v)$ are squeezed between measures $\mu_{x_-}^{m,+\infty}(v)$, $x_-\in{\mathbb Q}\cap(-\infty,x)$ and $\mu_{x_+}^{m,+\infty}(v)$, $x_+\in{\mathbb Q}\cap(-\infty,x)$. Now we need to show that there is a unique measure with this property. \begin{lemma}\label{lem:equicontinuity-of-polymer-measures-wrt-endpoint} Let $v\in{\mathbb R}$, $m,k\in{\mathbb Z}$, $k>m$, $r\in{\mathbb N}$, $y\in{\mathbb Q}$, and a sequence of measures $\nu_n$ satisfying LLN with slope $v$. Then there is an event $\Omega_{v,m,k,r,y}$ of probability~$1$ such that on that event, the family of functions $f_n:[-r,r]\cap{\mathbb Q}\to{\mathbb R}$, $n>k$, defined by \begin{equation*} f_n(x)=\mu_{x,\nu_n}^{m,n}\pi_{k}^{-1}((-\infty,y]) \end{equation*} is uniformly equicontinuous on $[-r,r]\cap{\mathbb Q}$. \end{lemma} \bpf Without loss of generality, we assume that $m=0$. To prove the uniform equicontinuity, we will check that for every ${\varepsilon}\in(0,1/2)$, there is $\delta>0$ such that \begin{equation} \label{eq:equicontinuity-wrt-endpoint} f_n(x_0)-f_n(x'_0)\le 6{\varepsilon},\quad |x_0|,|x'_0|\le r,\ |x_0-x'_0|\le \delta. \end{equation} First, we use LLN for $(\nu_n)$ to find $L>0$ such that \begin{equation} \label{eq:sublinear-growth-of-nu_n} \nu_n({\mathbb R}\setminus [-Ln,Ln])<{\varepsilon},\quad n\in{\mathbb N}. \end{equation} Then we use monotonicity and tightness to find $R>|y|$ such that \begin{equation} \label{eq:uniform-compactness-in-starting-point-and-linearly-growing-endpoint} \mu_{x,y}^{0,n}\pi_{1,k}^{-1}({\mathbb R}^k\setminus B_R^k)<{\varepsilon},\quad x\in[-r,r],\ n\in{\mathbb N},\ y\in [-Ln,Ln], \end{equation} where $B_R^k=[-R,R]^k$. Inequality~\eqref{eq:sublinear-growth-of-nu_n} implies \begin{align*} f_n(x_0)&=\int_{{\mathbb R}}\nu_n(dw) \mu_{x_0,w}^{0,n}\pi^{-1}_{1,k}({\mathbb R}^{k-1}\times(-\infty,y]) \\ &\le \int_{[-Ln,Ln]}\nu_n(dw) \mu_{x_0,w}^{0,n}\pi^{-1}_{1,k}({\mathbb R}^{k-1}\times(-\infty,y])+{\varepsilon}. \end{align*} Introducing $B_R^k(y)=[-R,R]^{k-1}\times [-R,y]$ and $B_{Ln}=[-Ln,Ln]$, we can use~\eqref{eq:uniform-compactness-in-starting-point-and-linearly-growing-endpoint} to write \begin{align*} f_n(x_0)&\le \int_{B_{Ln}}\nu_n(dw) \mu_{x_0,w}^{0,n}\pi^{-1}_{1,k}(B_R^k(y))+2{\varepsilon}, \\ & \le \int_{B_{Ln}}\nu_n(dw)\frac{ \int_{B_R^k(y)} Z(x_0,\ldots,x_k,w)\,dx_1\ldots dx_k }{\int_{B_R^k} Z(x_0,\ldots,x_k,w)\,dx_1\ldots dx_k }+2{\varepsilon}, \end{align*} where \[ Z(x_0,\ldots,x_k,w)=e^{-F_0(x_0)}g(x_1-x_0)\cdot \prod_{i=1}^{k-1} e^{-F_i(x_i)}g(x_{i+1}-x_{i})\cdot Z_{x_k,w}^{k,n}. \] For every $\delta>0$, let us define \[ K_\delta=\sup\left\{\frac{e^{-F(x'_0)}g(x_1-x'_0)}{e^{-F(x_0)}g(x_1-x_0)}:\ |x_0|,|x'_0|\le r,\ |x_0-x'_0|\le \delta,\ |x_1|\le R\right\}. \] Then $\lim_{\delta\downarrow 0}K_\delta =1$ with probability $1$. Also, we can continue the above sequence of inequalities, assuming $|x_0-x'_0|\le \delta$: \begin{align*} f_n(x_0) &\le K^2_\delta \int_{B_{Ln}}\nu_n(dw)\frac{ \int_{B_R^k(y)} Z(x'_0,x_1\ldots,x_k,w)\,dx_1\ldots dx_k }{\int_{B_R^k} Z(x'_0,x_1\ldots,x_k,w) \,dx_1\ldots dx_k }+2{\varepsilon} \\ &\le K^2_\delta \int_{{\mathbb R}}\nu_n(dw)\frac \int_{{\mathbb R}^{k-1}\times (-\infty,y]} Z(x'_0,x_1\ldots,x_k,w) \,dx_1\ldots dx_k }{(1-{\varepsilon})\int_{{\mathbb R}^k} Z(x'_0,x_1\ldots,x_k,w) \,dx_1\ldots dx_k }+2{\varepsilon} \\ &\le \frac{K^2_\delta}{1-{\varepsilon}}f_n(x'_0)+2{\varepsilon}. \end{align*} Therefore, if $\delta$ is chosen so that $K^2_\delta\le 1+{\varepsilon}$, we obtain \[ f_n(x_0)-f_n(x'_0)\le \left(\frac{K^2_\delta}{1-{\varepsilon}}-1\right)f_n(x'_0)+2{\varepsilon} \le \frac{K^2_\delta}{1-{\varepsilon}}-1+2{\varepsilon}\le \frac{1+{\varepsilon}}{1-{\varepsilon}}-1+2{\varepsilon}\le 6{\varepsilon}, \] and~\eqref{eq:equicontinuity-wrt-endpoint} holds. {{\hfill $\Box$ \smallskip}} \begin{lemma}\label{lem:continuity-of-one-sided-polymer-measures-wrt-endpoint} Let $v\in{\mathbb R}$, $m,k\in{\mathbb Z}$, $k>m$, $r\in{\mathbb N}$, $y\in{\mathbb Q}$. On $\tilde\Omega_v\cap \Omega_{v,m,k,r,y}$, the function $f:[-r,r]\cap{\mathbb Q}\to{\mathbb R}$, defined by \begin{equation} f(x)=\mu_{x}^{m,+\infty}(v)\pi_{k}^{-1}((-\infty,y]), \end{equation} is uniformly continuous on $[-r,r]\cap{\mathbb Q}$. \end{lemma} \bpf Let us choose any sequence $(\nu_n)$ satisfying LLN with slope $v$ and define $f_n$ as in Lemma~\ref{lem:equicontinuity-of-polymer-measures-wrt-endpoint}. The statement follows then from that lemma since $\lim_{n\to\infty}f_n(x)=f(x)$ for $x\in [-r,r]\cap{\mathbb Q}$. {{\hfill $\Box$ \smallskip}} We can now prove the complete uniqueness and weak convergence claims of Theorem~\ref{thm:thermodynamic-limit}: \begin{lemma}\label{lem:uniqueness-of-polymer-measures} Let $v\in{\mathbb R}$. Then on $\Omega_v=\tilde\Omega_v\cap \bigcap_{m,k,r,y}\Omega_{v,m,k,r,y}$, \begin{enumerate}[1.] \item \label{it:uniqueness-at-reals} For any point $(m,x)\in{\mathbb Z}\times{\mathbb R}$, the set $\mathcal{P}_x^{m,+\infty}(v)$ of all polymer measures on $S_x^{m,+\infty}$ satisfying SLLN with slope $v$, contains exactly one element, $\mu_{x}^{m,+\infty}(v)$. \item \label{it:convergence-of-polymer-measures-at-reals} For any point $(m,x)\in{\mathbb Z}\times{\mathbb R}$ and for every sequence of measures $(\nu_n)$ satisfying LLN with slope $v$, $\mu_{x,\nu_n}^{m,n}$ converges to $\mu_{x}^{m,+\infty}(v)$ weakly. \end{enumerate} \end{lemma} \bpf The second part follows from the first one and the compactness argument explained in the proof of Lemma~\ref{lem:uniqueness-and-convergence-at-rationals}. To prove the first part, it is sufficient to fix $(m,x)\in\Z\times{\mathbb R}$ and check that for every $k>m$, the marginal measure $\nu_k=\mu\pi_{k}^{-1}$ does not depend on $\mu\in\mathcal{P}_x^{m,+\infty}(v)$. For that, it suffices to see that for every choice of $y\in{\mathbb Q}$, $\nu_k((-\infty,y])$ does not depend on $\mu\in\mathcal{P}_x^{m,+\infty}(v)$. If $x_-<x<x_+$, then $\mu_{x_-,\nu_n}^{m,n}$ is dominated by $\mu_{x,\nu_n}^{m,n}$ which is dominated by $\mu_{x_+,\nu_n}^{m,n}$. Therefore, for every $y\in{\mathbb R}$, \[ \mu_{x_-,\nu_n}^{m,n}\pi_{k}^{-1}((-\infty,y])\ge \mu_{x,\nu_n}^{m,n}\pi_{k}^{-1}((-\infty,y])\ge \mu_{x_+,\nu_n}^{m,n}\pi_{k}^{-1}((-\infty,y]). \] Since $\mu_{x,\nu_n}^{m,n}\pi_{k}^{-1}=\nu_k$, we obtain \begin{equation} \mu_{x_-,\nu_n}^{m,n}\pi_{k}^{-1}((-\infty,y])\ge \nu_k((-\infty,y])\ge \mu_{x_+,\nu_n}^{m,n}\pi_{k}^{-1}((-\infty,y]). \label{eq:nu_k-squeezed-1} \end{equation} If additionally $x_-,x_+\in{\mathbb Q}$, then f.d.d.'s of $\mu_{x_-,\nu_n}^{m,n}$ and $\mu_{x_+,\nu_n}^{m,n}$ weakly converge to those of $\mu^{m,+\infty}_{x_-}(v)$ and $\mu^{m,+\infty}_{x_+}(v)$, due to Lemma~\ref{lem:uniqueness-and-convergence-at-rationals} since $(\nu_n)_{n>m}$ satisfies LLN with slope~$v$. Since marginals of both $\mu^{m,+\infty}_{x_-}(v)$ and $\mu^{m,+\infty}_{x_+}(v)$ are absolutely continuous, \eqref{eq:nu_k-squeezed-1} implies \[ \mu^{m,+\infty}_{x_-}(v)((-\infty,y])\ge \nu_k((-\infty,y])\ge \mu^{m,+\infty}_{x_+}(v)((-\infty,y]). \] Lemma~\ref{lem:continuity-of-one-sided-polymer-measures-wrt-endpoint} implies that \[ \inf_{x_-\in {\mathbb Q}\cap (-\infty,x)} \mu^{m,+\infty}_{x_-}(v)((-\infty,y]) = \sup_{x_+\in {\mathbb Q}\cap (x,+\infty)} \mu^{m,+\infty}_{x_+}(v)((-\infty,y]). \] Denoting this common value by $c$, we conclude that the value of $\nu_k((-\infty,y])$ is uniquely defined and equals~$c$, which completes the proof. {{\hfill $\Box$ \smallskip}} \section{Existence and uniqueness of global solutions and basins of pullback attraction} \label{sec:global-solutions-and-ratios-of-partition-functions} The main goal of this section is to prove Theorems~\ref{thm:global_solutions} and~\ref{thm:pullback_attraction} on global solutions of the randomly kicked Burgers equation. These solutions will be constructed and studied via the pullback procedure with the help of backward polymer measures. We also prove the total variation convergence part in the thermodynamic limit Theorem~\ref{thm:thermodynamic-limit}. Our main result on thermodynamic limit for polymer measures, Theorem~\ref{thm:thermodynamic-limit}, was naturally stated and proved in terms of forward polymer measures defined on $S^{m,+\infty}_x(v)$ for endpoints $(m,x)\in\Z\times{\mathbb R}$ and slopes $v$. A direct counterpart of this result holds for thermodynamic limit polymer measures defined on spaces $S^{-\infty,m}_x(v)$ of backward infinite sequences $y:\{\ldots,m-2,m-1,m\}\to{\mathbb R}$ such that $y_m=x$ and $\lim_{n\to-\infty}(y_n/n)=v$. All the definitions, notation, and intermediate results can be straightforwardly adapted to this case. We do not introduce these adaptations one-by-one since all of them are unambiguously defined by the context. When dealing with backward polymer measures we will still refer to results obtained for forward polymer measures, although formally they are not equivalent due to the asymmetry in the definition of path energy near the path endpoints in~\eqref{eq:path-energy}. Moreover, at times we abuse the notation by using the same symbols for objects constructed for forward and backward versions. \smallskip A function $u(n,x)=u_{\omega}(n,x)$ is a global solution of the Burgers equation if the version of the Hopf--Cole transform defined by \begin{equation*} V(n,x) = e^{-U(n,x)} = e^{-\int_0^x u(n,y) dy},\quad (n,x)\in\Z\times{\mathbb R}, \end{equation*} satisfies, for all integers $m<n$ and all $x\in{\mathbb R}$, \begin{equation} \label{eq:evolution-for-V} V(n,x) = C_{m,n}[\Xi_{\omega}^{m,n} V(m,\cdot)](x) := C_{m,n}\int Z^{m,n}(y,x) V(m,y) dy, \end{equation} where $(C_{m,n})$ is a random family of constants such that $C_{m,n}C_{n,k}=C_{m,k}$, $m<n<k$. We need to introduce the normalizing constants $C_{m,n}$ for consistency with the identity $V(n,0)=1$ holding for all $n$, because we fix the the lower limit of integration to be zero when defining the Hopf--Cole transform. The following computation shows that, given any $v\in{\mathbb R}$ and $N\in{\mathbb Z}$, the functions \begin{equation} \label{eq:Hopf-Cole-for-prelimit} V_{v}^{N}(n,x) = Z^{N,n}_{Nv,x}/Z^{N,n}_{Nv,0},\quad n>N,\ x\in{\mathbb R}, \end{equation} and constants \begin{equation} \label{eq:Normalizing-constants-product-cocycle} C^N_{v,m,n} = Z^{N,m}_{Nv,0}/Z^{N,n}_{Nv,0},\quad m<n, \end{equation} satisfy (\ref{eq:evolution-for-V}) for $n> m >N$: \begin{equation* \begin{split} C_{v,m,n}^N [\Xi_{\omega}^{m,n}V^N_{v}(m,\cdot)](x) &= C_{v,m,n}^{N}\int Z^{m,n}(y,x) V_{v}^N(m,y) dy \\ &= Z^{N,m}_{Nv,0}/Z^{N,n}_{Nv,0} \int Z^{N,m}_{Nv,y} Z^{m,n}_{y,x} /Z^{N,m}_{Nv,0}\, dy\\ &= (Z^{N,n}_{Nv,0})^{-1} \int Z^{N,m}_{Nv,y} Z^{m,n}_{y,x}\, dy \\ &= (Z^{N,n}_{Nv,0})^{-1}Z^{N,n}_{Nv,x} = V^N_{v}(n,x). \end{split} \end{equation*} Therefore, a natural guess for the Hopf--Cole transform of global solutions will be $V(n,x)=V_v(n,x)=\lim\limits_{N\to -\infty}V^N_v(n,x)$, along with normalizing constants given by $C_{m,n}= C_{v,m,n}=\lim\limits_{N \to -\infty}C^N_{v,m,n}$. This leads to the study of the limits of partition function ratios. On the other hand, let $u^N_{v}(n,x) = -\frac{\partial}{\partial x}\ln V^N_v(n,x)$ be the inverse Hope--Cole transform of $V^{N}_v(n,x)$. We find that \begin{equation} \label{eq:approx-solution-via-polymer-measures} \begin{split} u^N_v(n,x) &= -\frac{\partial}{\partial x} \ln \int_{{\mathbb R}} \frac{1}{\sqrt{2\pi}}e^{-(x-y)^2/2 - F_{n-1}(y)} Z^{N,n-1}(Nv,y) dy \\ &= \frac{ \int_{{\mathbb R}} (x-y)\frac{1}{\sqrt{2\pi}}e^{-(x-y)^2/2-F_{n-1}(y)} Z^{N,n-1}(Nv,y) dy }{ Z^{N,n}(Nv,x)} \\ &= \int_{{\mathbb R}} (x-y) \mu_{Nv,x}^{N,n} \pi_{n-1}^{-1}(dy). \end{split} \end{equation} Taking the limit $N \to -\infty$, we expect the global solution to be \begin{equation} \label{eq:global-solution-via-marginal-polymer-measures} u_v(n,x) = \int_{{\mathbb R}} (x-y) \mu_x^{-\infty,n}(v) \pi_{n-1}^{-1}(dy), \end{equation} where $\mu_x^{-\infty,n}(v)$ is the backward polymer measure, the weak limit of $\mu_{Nv,x}^{N,n}\pi_{n-1}^{-1}$ as $N\to-\infty$. To justify this answer, we actually need a stronger statement than weak convergence, namely, a statement on convergence of the associated densities. The convergence of densities is closely related to convergence of partition function ratios, since the density of $\mu_{Nv,x}^{N,n}\pi_m^{-1}$ is precisely \begin{equation*} \frac{d \mu_{\nu_N,x}^{N,n}\pi_m^{-1} }{ d\mathrm{Leb} }(y) = \frac{Z_{Nv, y}^{N,m}}{Z_{Nv,x}^{N,n}}Z_{y,x}^{m,n}. \end{equation*} In Section~\ref{sec:limits-of-partition-function-ratios}, we will show that both convergences are uniform on compact sets. The existence of global solutions is then established in Section~\ref{sec:existence-of-global-solution}. The uniqueness of global solutions relies on the uniqueness of infinite volume polymer measures with any given slope $v$. Suppose $u_v(n,x) \in \mathbb{H'}(v,v)$ is a global solution and $V_v(n,x)$ is its Hopf--Cole transform. For fixed $(n,x) \in {\mathbb Z} \times {\mathbb R}$, we can define a ``backward'' point-to-line polymer measure $\bar{\mu}^{-\infty,n}_{x}$ on the set $S_x^{-\infty,n}$ of paths $\gamma:\{...,n-2,n-1,n\} \to {\mathbb R}$ with $\gamma(n)=x$ : \begin{align} \label{eq:def-of-mu-bar}\nonumber &\quad \bar{\mu}_x^{-\infty,n}(A_{n-k} \times ... \times A_{n-1} \times A_n) \\ &= \frac{ \int_{A_{n-k}} dx_{n-k} \cdots \int_{A_{n-1}} dx_{n-1} \int_{A_n} \delta_x(dx_n) \quad V_v(n-k,x_{n-k}) \prod\limits_{i=n-k}^{n-1} Z^{i,i+1}_{x_i, x_{i+1}} }{\int_{{\mathbb R}} V_v(n-k, x_{n-k}) Z^{n-k,n}_{x_{n-k},x} dx_{n-k} }. \end{align} This definition is consistent for different choices of $k$ since $V_v(n,x)$ satisfies (\ref{eq:evolution-for-V}). Then the global solution $u_{v}(n,x)$ is uniquely determined by $\bar{\mu}_x^{-\infty,n}$ through \begin{equation}\label{eq:solution-through-pt-to-line-measure} u_{v}(n,x) = \int_{{\mathbb R}} (x-y) \bar{\mu}_x^{-\infty,n} \pi_{n-1}^{-1}(dy). \end{equation} We will show that the measures $\bar{\mu}_x^{-\infty,n}$ satisfy LLN with slope~$v$. This will allow us to conclude that they are are uniquely defined by the potential and coincide with~$\mu_x^{-\infty,n}(v)$, so the global solution in ${\mathbb H}'(v,v)$ is also uniquely defined by the potential and coincides with $u_v$, see~(\ref{eq:global-solution-via-marginal-polymer-measures}). This is done in Section~\ref{sec:uniqueness-of-global-solutions}. In Section~\ref{sec:pullback} we show that global solutions are also pullback attractors. We also generalize the result on convergence of density functions to certain point-to-line polymer measures. \subsection{Limits of partition function ratios} \label{sec:limits-of-partition-function-ratios} Let $\Omega'_v = \Omega' \cap \Omega_v$, where $\Omega'$ and $\Omega_v$ have been introduced in Theorems~\ref{thm:all-space-time-point-compactness} and~\ref{thm:thermodynamic-limit}. In this section, we will prove the following two theorems on $\Omega'_v$: \begin{theorem} \label{thm:convergence-of-partition-function-ratios} For all $\omega \in \Omega_v'$, there is a function $G_v\bigl( (n_1,x_1), (n_2,x_2) \bigr)>0$ such that for any sequence $(y_N)$ satisfying $\lim\limits_{N \to -\infty} y_N/N = v$, we have \begin{equation*} \lim\limits_{N\to -\infty} Z_{y_N,x_1}^{N,n_1} /Z_ {y_N, x_2}^{N, n_2} = G_{v}\bigl( (n_1,x_1), (n_2,x_2) \bigr). \end{equation*} For $n_1,n_2$ fixed and $x_1,x_2$ restricted to a compact set, the convergence is uniform. \end{theorem} \begin{theorem} \label{thm:convergence-of-density-functions} For all $\omega \in \Omega_v'$ the following is true. Suppose a family of probability measures $\bigl( \nu_N \bigr)$ satisfies the following conditions: \begin{enumerate} \item\label{item:23} for some $c>0$ and all sufficiently large $|N|$, \begin{equation} \label{eq:finite-support-assumption} \nu_N \big( [-c|N|, c|N|]^c \big)=0; \end{equation} \item\label{item:24} $\bigl( \nu_N \bigr)$ satisfies LLN with slope $v$ as $N\to -\infty$. \end{enumerate} Let $N<m<n$ and let $f^N_{n,m}(x,\cdot)$ be the density of $\mu_{\nu_N,x}^{N,n} \pi_m^{-1}$, namely, \begin{equation*} f^N_{n,m}(x,y) = \int_{-c|N|}^{c|N|} \frac{Z^{N,m}(z,y)Z^{m,n}(y,x)}{Z^{N,n}(z,x)} \nu_N(dz). \end{equation*} Then $f_{n,m}^N(x,y)$ converges uniformly in $x$ and $y$ on compact sets to $f_{v,n,m}(x,y)$ as $N \to -\infty$, where $f_{v,n,m}(x,\cdot)$ is the density of $\mu_x^{-\infty,n}(v) \pi_m^{-1}$ and can be expressed as \begin{equation} \label{eq:expression-of-polymer-measure-density} f_{v,n,m}(x,y) = Z^{m,n}_{y,x} G_v \bigl( (m,y), (n,x) \bigr). \end{equation} \end{theorem} We can derive part~\ref{it:thermodynamic-limit} of Theorem~\ref{thm:thermodynamic-limit} from Theorem~\ref{thm:convergence-of-density-functions}. Combined with Section~\ref{sec:infinite-volume-polymer-measure}, this completes the proof of Theorem~\ref{thm:thermodynamic-limit}. \begin{proof}[Proof of (the time-reversed version of) part~\ref{it:thermodynamic-limit} of Theorem~\ref{thm:thermodynamic-limit} ] Let us take the full measure set $\Omega'_v$. For every $\omega \in \Omega'_v$, our goal is to show that for any $(n,x) \in \Z\times{\mathbb R}$ and $(\nu_N)$ satisfying LLN with slope $v$, $\mu_{\nu_N, x}^{N,n}\pi_m^{-1}$ converges to $\mu_x^{-\infty,n}(v)\pi_m^{-1}$ in total variation for all $m<n$. Let $c > |v| + 1$. Denoting the conditioning of $\nu_N$ on $[-c|N|, c|N|]$ by $\tilde{\nu}_{N}$, we get \begin{align*} &\| \mu_{\nu_N, x}^{N,n}\pi_m^{-1} - \mu_x^{-\infty,n}(v)\pi_m^{-1} \|_{\text{TV}} \\ \le & \| \mu_{\nu_N, x}^{N,n}\pi_m^{-1} - \mu_{\tilde{\nu}_N, x}^{N,n}\pi_m^{-1} \|_{\text{TV}} + \| \mu_{\tilde{\nu}_N, x}^{N,n}\pi_m^{-1} - \mu_x^{-\infty,n}(v)\pi_m^{-1} \|_{\text{TV}} \\ \le & \| \nu_N - \tilde{\nu}_N \|_{\text{TV}} + \| \mu_{\tilde{\nu}_N, x}^{N,n}\pi_m^{-1} - \mu_x^{-\infty,n}(v)\pi_m^{-1} \|_{\text{TV}}. \end{align*} The first term goes to $0$ since $(\nu_N)$ satisfies LLN with slope~$v$. To see that the second term goes to $0$, we notice that $\bigl( \tilde{\nu}_N \bigr)$ satisfies LLN with slope $v$ and~(\ref{eq:finite-support-assumption}), so we can apply Theorem~\ref{thm:convergence-of-density-functions} to conclude that the densities of~$\mu_{\tilde{\nu}_N, x}^{N,n}\pi_m^{-1}$ converge to that of~$\mu_x^n(v)\pi_m^{-1}$, uniformly on any compact set, which implies convergence in total variation. This completes the proof. \end{proof} In what follows, we could use the spatial smoothness of $F$ to simplify some of the arguments, but we prefer to give a version of the proof that does not even assume continuity of $F$. Let us define \begin{equation}\label{eq:definition-of-gN} g^N_{n,m}(x,y) = \int_{-c|N|}^{c|N|} \frac{Z^{N,m}(z,y)}{Z^{N,n}(z,x)} \nu_{N}(dz), \end{equation} so that \begin{equation} \label{eq:relation-between-f-and-g} f_{n,m}^N(x,y) = Z^{m,n}_{y,x} g^N_{n,m}(x,y), \end{equation} then $g^N_{n,m}(x,y)$ is continuous in both $x$ and $y$ as the following lemma shows. \begin{lemma} \label{lem:continuous-of-g} If $\nu_N$ satisfies (\ref{eq:finite-support-assumption}), then $g^N_{n,m}(x,y)$ is continuous in $x$ and $y$. \end{lemma} \begin{proof} By Lemma \ref{lem:smoothness-of-partition-ratio}, $h(x,y,z) = Z^{N,m}(z,y)/Z^{N,n}(z,x)$ is uniformly continuous in $x,y$, and $z$ on any compact set, so the lemma follows from~\eqref{eq:definition-of-gN}. \end{proof} The next lemma shows that if $n<m$, then for any compact set $K \subset {\mathbb R} \times {\mathbb R}$, the family $\bigl( \ln g^N_{n,m}(\cdot,\cdot) \bigr)_{N<m}$ is precompact in $\mathcal{C}(K)$, the space of continuous functions on $K$. From this we will derive Theorems~\ref{thm:convergence-of-partition-function-ratios} and~\ref{thm:convergence-of-density-functions}. \begin{lemma} \label{lem:precompactness-of-gN} Let $\omega \in \Omega'$ and $K$ be a compact subset of ${\mathbb R}\times{\mathbb R}$. If measures $\nu_{N}$ satisfy (\ref{eq:finite-support-assumption}), then for any $m,n\in{\mathbb Z}$ satisfying $m<n$, the family $\bigl( \ln g^N_{n,m}(\cdot,\cdot) \bigr)_{N<m}$ is precompact in $\mathcal{C}(K)$. \end{lemma} Note that for a compact set $K$ and a family of positive functions $h^N$, the precompactness of $\bigl( \ln h^N \bigr)$ implies that of $\bigl( h^N \bigr)$ in $\mathcal{C}(K)$. \begin{proof}[Proof of Theorems \ref{thm:convergence-of-partition-function-ratios} and \ref{thm:convergence-of-density-functions}] We fix $m < n$ and let $\omega \in \Omega'_v$. Using Lemma \ref{lem:precompactness-of-gN} and the standard diagonal procedure, we can find a subsequence $\bigl( g^{N_k}_{n,m}(x,y) \bigr)$ converging to some $\tilde{g}(x,y)$, uniformly on any compact set. Since $\ln Z^{m,n}(x,y)$ is bounded on every compact set, by (\ref{eq:relation-between-f-and-g}), we see that $f^{N_k}_{n,m}(x,y)$ converges to $\tilde{f}(x,y) = Z^{m,n}(x,y)\tilde{g}(x,y)$ uniformly on compact sets. Let us now identify the limit of any subsequence $\bigl( f^{N_k}_{m,n}(x,y) \bigr)$ converging uniformly on all compact sets. On $\Omega_v$, if $(\nu_{N_k})$ satisfies LLN with slope $v$, then $\mu_{\nu_{N_k}, x}^{N_k,n} \pi_m^{-1}$ converge weakly to $\mu_x^{-\infty,n}(v)\pi_m^{-1}$. Hence $ \tilde{f}(x,\cdot)$ must equal $f_{v,n,m}(x,\cdot)$, the density of $\mu_x^{-\infty,n}(v) \pi_m^{-1}$. So, the only possible limit of any subsequence of $\bigl( f^N_{n,m}\bigr)$, is $f_{v,n,m}$, and we obtain that $f^N_{n,m}$ converges to $f_{v,n,m}$, uniformly on any compact set. By (\ref{eq:relation-between-f-and-g}), $g^N_{n,m}$ also converges to $\tilde{g}(x,y) =\bigl( Z^{m,n}_{y,x} \bigr)^{-1} f_{v,n,m}(x,y)$, uniformly on any compact set. Next we will show \eqref{eq:expression-of-polymer-measure-density} and Theorem \ref{thm:convergence-of-partition-function-ratios}. Let $(y_N)$ be such that $y_N/N\to v$. Then $\nu_N = \delta_{y_N}$ satisfy (\ref{eq:finite-support-assumption}), so \begin{equation*} g^N_{n,m}(x,y) = Z^{N,m}_{y_N,y}/ Z^{N,n}_{y_N,x} \to \bigl( Z^{m,n}_{y,x} \bigr)^{-1} f_{v,n,m}(x,y), \end{equation*} where the convergence is uniform on any compact set. We denote the limit by $G_v \bigl( (m,y), (n,x) \bigr)$. By Lemma \ref{lem:precompactness-of-gN}, we know that $\bigl( \ln g^N_{n,m} \bigr)$ is uniformly bounded, hence $G_v$ is strictly positive. This proves \eqref{eq:expression-of-polymer-measure-density} and Theorem \ref{thm:convergence-of-partition-function-ratios} for $n_1 < n_2$. For $n_1 \ge n_2$, we can simply use the following two identities: \begin{equation*} \lim_{N\to -\infty} \frac{Z^{N,n_1}_{y_N,x_1}}{Z^{N,n_2}_{y_N,x_2}} = \Big( \lim_{N\to -\infty} \frac{Z^{N,n_2}_{y_N,x_2}}{Z^{N,n_1}_{y_N,x_1}} \Big)^{-1} \end{equation*} and \begin{equation*} \lim_{N\to -\infty} \frac{Z^{N,n_1}_{y_N,x_1}}{Z^{N,n_2}_{y_N,x_2}} = \lim_{N\to -\infty} \frac{Z^{N,n_1}_{y_N,x_1}}{Z^{N,n_3}_{y_N,x_3}} \lim_{N\to -\infty} \frac{Z^{N,n_3}_{y_N,x_3}}{Z^{N,n_2}_{y_N,x_2}}. \end{equation*} \end{proof} To prove Lemma \ref{lem:precompactness-of-gN}, we need the following corollary of Theorem \ref{thm:all-space-time-point-compactness}: \begin{lemma} \label{lem:one-step-control} Let $\omega\in \Omega'$. For $(l,q) \in {\mathbb Z} \times {\mathbb Z}$ and $c > 0$, there is a constant $n_1=n_1(l,q,c)$ such that for any $ s > n_1$, any $x \in [q,q+1]$, any terminal measure~$\nu$, any $s'\in[1,s]_{{\mathbb Z}}$, and any $N\in {\mathbb Z}$ such that $N \le l - 2s$, we have \begin{equation} \label{eq:one-step-large-small-probability} \mu_{\nu, x}^{N,l} \pi_{l-s'}^{-1} \big( [-(2c+R+1)s, (2c+R+1)s]^c \big) \le 4\nu \big( [-c|N|,c|N|]^c \big) + 8e^{-\sqrt{s}}. \end{equation} \end{lemma} \begin{proof} We apply the backward version of Theorem~\ref{thm:all-space-time-point-compactness}, taking $v' = 0, u_0 = c, u_1 = 2c$ and letting $n_1(l,q,c) = n_0(\omega, l, q, [c],[c^{-1}])$, so \eqref{eq:one-step-large-small-probability} is directly implied by the backward version of~\eqref{eq:compactness-control-by-terminal-measure-all-space-time}. \end{proof} \begin{proof}[Proof of Lemma \ref{lem:precompactness-of-gN}] Let us fix a compact set $K = [p,p+1]\times [-k,k]$ and times $m < n$. We denote $r = 2c+R+1$ and, for $\varepsilon \in (0, 1/2)$, define \begin{equation*} s_1 = \max\Big\{ n-m,\, n_1 (n,p,c),\, \frac{k}{r},\, \ln^2 \frac{\varepsilon}{64}\Big\} \end{equation*} and \begin{equation*} s_2 = \max \Big\{ n_1(m,i,c):\ |i| \le rs_1 +1\Big\} \vee \ln^2\frac{\varepsilon}{64}. \end{equation*} We will need several truncated integrals: \begin{align*} &\bar{Z}^{N,m}_{z,y} = \int_{-rs_2}^{rs_2} Z^{N,m-1}_{z,w} Z^{m-1,m}_{w,y} dw =\int_{-rs_2}^{rs_2} Z^{N,m-1}_{z,w} \frac{1}{\sqrt{2\pi}}e^{-\frac{(w-y)^2}{2} - F_{m-1}(w)} dw, \\ & \bar{Z}^{m,n}_{y,x} = \begin{cases} Z^{m,n}_{y,x},& m=n-1, \\ \int_{-rs_1}^{rs_1} Z^{m,n-1}_{y,w} Z^{n-1,n}_{w,x} dw ,& m < n-1, \end{cases} \\ & \bar{Z}^{N,n}_{z,x} = \int_{-rs_1}^{rs_1} \bar{Z}^{N,m}_{z,y} \bar{Z}^{m,n}_{y,x} dy, \\ & \bar{g}^N(x,y) = \int_{-c|N|}^{c|N|} \frac{ \bar{Z}^{N,m}_{z,y}}{ \bar{Z}^{N,n}_{z,x}} \nu_N(dz), \end{align*} where, according to \eqref{eq:Z}, $Z^{n-1,n}_{w,x}=\frac{1}{\sqrt{2\pi}} e^{-\frac{(w-x)^2}{2} - F_{n-1}(w)}$ in the definition of $\bar Z^{m,n}_{y,x}$. We also let $h_{\varepsilon}^N = \ln \bar{g}^N$, $N<m$, and define $\tilde{K} = [p,p+1]\times[-rs_1,rs_1] \supset K$. If we prove that for every ${\varepsilon}>0$, \begin{equation}\label{eq:eps-approximation-of-g} |\ln g^N_{n,m}(x,y) - h_{\varepsilon}^N(x,y) | \le \varepsilon,\quad (x,y) \in \tilde{K}, \end{equation} and $\big( h_{\varepsilon}^N\big)$ is precompact in $\mathcal{C}(\tilde{K})$, then the lemma will follow since, given any ${\varepsilon}>0$, we will be able to use an ${\varepsilon}$-net for $(h_{\varepsilon}^N)$ to construct a $2{\varepsilon}$-net for $(\ln g_{n,m}^N)$. Let $N \le \min\{n-2s_1, m-2s_2\}$. If $|y| \le rs_1$ and $|z| \le cN$, then, from $\delta_z ([-c|N|, c|N|]^c) = 0$ and \eqref{eq:one-step-large-small-probability} with $(l,s',s,w,\nu) = (m,1,s_2, y, \delta_z)$, we obtain \begin{equation*} 1 - \frac{\bar{Z}_{z,y}^{N,m}}{Z_{z,y}^{N,m}} \le \mu_{z,y}^{N,m} \pi_{m-1}^{-1} ([-rs_2, rs_2]^c) \le 8e^{-\sqrt{s_2}} \le \varepsilon/8. \end{equation*} Then using the elementary inequality $|\ln(1+x)| \le 2|x|$ for $|x| \le 1/2$ we find \begin{equation} \label{eq:estimate-on-above} e^{-\varepsilon/4} \le \bar{Z}_{z,y}^{N,m} / Z_{z,y}^{N,m} \le 1. \end{equation} Let \begin{equation*} \tilde{Z}^{N,n}_{z,x} = \int_{-rs_1}^{rs_1}Z^{N,m}_{z,y} \bar{Z}^{m,n}_{y,x} dy. \end{equation*} Then (\ref{eq:estimate-on-above}) implies \begin{equation} \label{eq:ratio-Z-tilde-and-Z-bar} 1\le \tilde{Z}_{z,x}^{N,n} / \bar{Z}_{z,x}^{N,n} \le e^{\varepsilon/4}. \end{equation} Similarly, if $x \in [p,p+1]$ and $|z| \le cN$, by (\ref{eq:one-step-large-small-probability}) with $(l,s',s,w,\nu) = (n,1,s_1, x, \delta_z)$ and $(n,n-m, s_1,x,\delta_z)$, we obtain \begin{equation*} 1- \frac{ \tilde{Z}_{z,x}^{N,n} }{Z_{z,x}^{N,n}} \le \mu_{z,x}^{N,n} \pi_{n-1}^{-1} ([-rs_1, rs_1]^c) + \mu_{z,x}^{N,n} \pi_m^{-1}([-rs_1,rs_1]^c) \le 16e^{-\sqrt{s_1}} \le \varepsilon/4. \end{equation*} Therefore, \begin{equation} \label{eq:estimate-on-below} e^{-\varepsilon/2} \le \tilde{Z}_{z,x}^{N,n}/ Z_{z,x}^{N,n} \le e^{\varepsilon/2}. \end{equation} Combining (\ref{eq:estimate-on-above}), (\ref{eq:ratio-Z-tilde-and-Z-bar}) and (\ref{eq:estimate-on-below}) we obtain \begin{equation*} e^{-\varepsilon} \le \bar{g}^N(x,y) / g^N_{n,m}(x,y) \le e^{\varepsilon}, \end{equation*} and (\ref{eq:eps-approximation-of-g}) follows. For any $|w| \le rs_2$ and $y,y' \in [-rs_1, rs_1]$, we have \begin{equation*} \bigg| \frac{(y-w)^2}{2} - \frac{(y'-w)^2}{2} \bigg| \le r(s_1+s_2) |y-y'|. \end{equation*} Hence \begin{equation*} \big| \ln \bar{Z}^{N,m}_{z,y}- \ln \bar{Z}_{z,y'}^{N,m} \big| \le r(s_1+s_2) |y-y'|. \end{equation*} Similarly, for all $x,x' \in [p,p+1]$, we have \begin{equation*} \big| \ln \bar{Z}^{N,n}_{z,x} - \ln \bar{Z}^{N,n}_{z,x'} \big| \le (rs_1 + |p| + 1 )|x-x'|. \end{equation*} Combining these two inequalities we see that \begin{equation}\label{eq:Lip-continuous} | h_{\varepsilon}^N(x,y) - h^N_{\varepsilon}(x',y') | \le L (|x-x'|+|y-y'|) \end{equation} for $L = r(s_1+s_2) + |p| + 1$. So, $h_{\varepsilon}^N$ are uniformly Lipschitz continuous and hence equicontinuous on $\tilde{K}$. It remains to show that $h_{\varepsilon}^{N}$ are uniformly bounded. Let \begin{equation*} \bar{f}^N(x,y) = \bar{g}^N(x,y) \bar{Z}^{m,n}_{y,x} = \int_{-c|N|}^{c|N|} \frac{ \bar{Z}^{N,m}_{z,y} \bar{Z}^{m,n}_{y,x}}{ \bar{Z}^{N,n}_{z,x}} \nu_N(dz). \end{equation*} Then for fixed $x$, \begin{equation*} \int_{-rs_1}^{rs_1} \bar{f}^N(x,y') dy' = 1. \end{equation*} By (\ref{eq:Lip-continuous}), we have for $y,y' \in [-rs_1,rs_1]$, \begin{equation*} \bar{g}^N(x,y) e^{-L\cdot 2rs_1} \le \bar{g}^{N}(x,y') \le \bar{g}^N(x,y) e^{L\cdot 2rs_1}. \end{equation*} Let $M$ be the supremum of $|\ln \bar{Z}^{m,n}_{\,\cdot,\, \cdot}|$ on $\tilde{K}$. Then \begin{equation*} \bar{f}^N(x,y) e^{-L\cdot2rs_1-2M} \le \bar{f}^{N}(x,y') \le \bar{f}^N(x,y) e^{L\cdot 2rs_1+2M}. \end{equation*} Integrating this inequality over $y'\in [-rs_1, rs_1]$ gives us \[ 2rs_1\bar{f}^N(x,y) e^{-L\cdot 2rs_1-2M} \le 1 \le 2rs_1 \bar{f}^N(x,y) e^{L\cdot 2rs_1+2M}. \] and hence $\ln \bar{f}^N(x,y)$ are uniformly bounded on $\tilde{K}$. Therefore, $$h^N_{\varepsilon}(x,y) = \ln \bar{f}^N(x,y) - \ln \bar{Z}^{m,n}_{y,x} $$ are also uniformly bounded. \end{proof} \subsection{Existence of global solutions} \label{sec:existence-of-global-solution} In this section, for every $v\in{\mathbb R}$, we will prove the existence of global solutions on a full measure set $\tilde{\Omega} \cap \Omega_v'$. Here, $\Omega_v'$ has been introduced in the beginning of Section~\ref{sec:limits-of-partition-function-ratios} and $\tilde{\Omega}$ is introduced in the following lemma controlling the tail of $\mu_{Nv,x}^{N,n} \pi_{n-1}^{-1}$. \begin{lemma} \label{lem:small-first-step-almost-surely-all-range} There is a full measure set $\bar{\Omega}$ on which for every $c > 0$ and $(n,q) \in {\mathbb Z}\times {\mathbb Z}$, there are constants $a_1, a_2,L_0 > 0$ and $N_0$ depending on $c$, $n$ and $q$ such that \begin{equation} \label{eq:small-first-step-almost-surely-all-range} \mu_{\nu, x}^{N, n} \pi_{n-1}^{-1} \bigl( [-L,L]^c \bigr) \le 4\nu \bigl( [-c|N|,c|N|]^c \bigr) + a_1 e^{-a_2 \sqrt{L}} \end{equation} for any $N \le N_0$, $L \ge L_0$, $x \in [q,q+1]$ and any terminal measure $\nu$. \end{lemma} A proof of the lemma will be given at the end of this section. Let us fix $v\in{\mathbb R}$ and assume that $\omega \in \bar{\Omega} \cap \Omega_v'$ throughout this section. Let us define $u^N_v(n,x)$, its Hopf--Cole transform $V^N_v(n,x)$, and the constants $C^N_{v,m,n}$ by~\eqref{eq:Hopf-Cole-for-prelimit},~\eqref{eq:Normalizing-constants-product-cocycle}, and \eqref{eq:approx-solution-via-polymer-measures}. We can use the function $G_v$ introduced in Theorem \ref{thm:convergence-of-partition-function-ratios} to define \begin{equation*} V_v(n,x) = G_v \bigl( (n,x), (n,0) \bigr) ,\quad C_{v,m,n} = G_v \bigl( (m,0), (n,0) \bigr). \end{equation*} \begin{lemma} \label{lem:burgers-in-HC-satisfied} The functions $V_v(n,x)$ and constants $C_{v,m,n}$ satisfy \eqref{eq:evolution-for-V}. \end{lemma} \begin{proof} Fix $m<n$ and $x$. We want to show \begin{equation*} G_v \bigl( (n,x), (n,0) \bigr)= G_v \bigl( (m,0), (n,0) \bigr) \int Z^{m,n}(y,x) G_v \bigl( (m,y), (m,0) \bigr)dy, \end{equation*} which, by Theorem \ref{thm:convergence-of-partition-function-ratios}, is equivalent to \begin{equation*} 1 = \int Z^{m,n}_{y,x} G_v \bigl( (m,y), (n,x) \bigr)dy. \end{equation*} This identity is true because by Theorem \ref{thm:convergence-of-density-functions}, the integrand is the density of $\mu_x^{-\infty,n}(v) \pi_m^{-1}$. \end{proof} Let $f^N_{v,n,n-1}(x,y)$ be the density of $\mu_{Nv,x}^{N,n}\pi_{n-1}^{-1}$. Then~\eqref{eq:approx-solution-via-polymer-measures} rewrites as \begin{equation*} u^N_v(n,x) = \int_{{\mathbb R}} (x-y) f_{v,n,n-1}^N(x,y) dy. \end{equation*} Recalling that we expect the global solution to be given by (\ref{eq:global-solution-via-marginal-polymer-measures}), we use the limiting density $f_{v,n,n-1}(x,y)$ from Theorem \ref{thm:convergence-of-density-functions} to define \begin{equation*} u_v(n,x) = \int_{{\mathbb R}} (x-y) f_{v,n,n-1}(x,y) dy. \end{equation*} \begin{lemma} \label{lem:convergence-of-velocity} The functions $u_v^N(n,\cdot)$ converge to $u_v(n,\cdot)$ as $N \to -\infty$, uniformly on compacts sets, and the Hopf--Cole transform of $u_v(n,\cdot)$ is $V_v(n,\cdot)$. \end{lemma} \begin{proof} Let $q \in {\mathbb Z}$. Lemma \ref{lem:small-first-step-almost-surely-all-range} implies that for some constants $a_1,a_2, L_0$ and~$N_0$, \begin{equation*} \mu_{Nv,x}^{N,n} \pi_{n-1}^{-1}\bigl( [-L, L]^c \bigr) = \int_{ |y| > L} f^N_{v,n,n-1}(x,y) dy \le a_1 e^{-a_2 \sqrt{L}},\quad x\in[q,q+1], \end{equation*} for all $N \le N_0$ and $L \ge L_0$, if we take $c > |v|$. Moreover, by Theorem \ref{thm:convergence-of-density-functions}, $f_{v,n,n-1}^N(x,y)$ converges to $f_{v,n,n-1}(x,y)$ uniformly on compact sets. Therefore $u_v^N(n,\cdot)$ converges to $u_v(n,\cdot)$ uniformly on $[q,q+1]$. Since $u^N_v(n,\cdot)$ and $V^N_v(n,\cdot)$ converge to $u_v(n,\cdot)$ and $V^N(n,\cdot)$ on compact sets, taking the limit $N \to -\infty$ on both sides of \begin{equation*} V^N_v(n,x) = e^{-\int_0^x u^N_v(n,x')dx' }, \end{equation*} we see that $V_v(n,x)$ is the Hopf--Cole transform of $u_v(n,x)$. \end{proof} To show that $u_v(n,\cdot)\in \mathbb{H}'(v,v)$, we need the following lemma which we will prove in the end of this section. \begin{lemma} \label{lem:uniform-integrability-of-uN} Given $n\in{\mathbb Z}$ and a compact set $K\subset{\mathbb R}$, the family of random variables $\big\{ u^N_v(n,x):\ N<n,\ x\in K\big\}$ is uniformly integrable. \end{lemma} \begin{proof}[Proof of the existence part of Theorem \ref{thm:global_solutions}] By Lemmas \ref{lem:burgers-in-HC-satisfied} and \ref{lem:convergence-of-velocity}, $u_v(n,x)$ is a global solution. It remains to show that $u_v(n,\cdot) \in \mathbb{H}'(v,v)$. All the other properties are easy to check. Lemma \ref{lem:uniform-integrability-of-uN} implies that \begin{equation}\label{eq:L-one-limit} \lim_{N\to -\infty} \mathsf{E} u_v^N(n,x) = \mathsf{E} u_v(n,x). \end{equation} By Lemma \ref{lem:shear-for-Z_v}, for any $(m_1,x_1)$ and $(m_2,x_2)$ such that $m_1<m_2$, we have \begin{equation*} Z^{m_1,m_2}(x_1,x_2) \stackrel{d}{=} e^{-\frac{(x_1-x_2)^2}{2(m_2-m_1)}} Z^{0,m_2-m_1}(0,0). \end{equation*} Taking logarithm and then expectation, we obtain \begin{equation*} \mathsf{E} \ln Z^{m_1,m_2}(x_1,x_2) = -\frac{(x_1-x_2)^2}{2(m_2-m_1)} + \mathsf{E} \ln Z^{0,m_2-m_1}(0,0), \end{equation*} so \begin{equation} \label{eq:expectation-of-V} \mathsf{E} \ln V^N_v(n,x) = \mathsf{E} \ln Z^{N,n}(Nv,x) - \mathsf{E} \ln Z^{N,n}(Nv,0) = -\frac{x(x-2Nv)}{2(n-N)}. \end{equation} For any $N$, by Hopf--Cole transform we have \begin{equation*} \int_0^{x} u_v^N(n,x')dx' = - \ln V^N_v(n,x). \end{equation*} Taking expectation of both sides, using the Fubini theorem and (\ref{eq:expectation-of-V}), we obtain \begin{equation*} \int_0^x \mathsf{E} u_v^N(n,x') dx' = \frac{x(x-2Nv)}{2(n-N)}. \end{equation*} Taking the limit $N \to -\infty$ and using \eqref{eq:L-one-limit}, we obtain \begin{equation*} \int_0^x \mathsf{E} u_v(n,x')dx' = vx. \end{equation*} By stationary of $u_v(n,\cdot)$, the left hand side is $x \cdot \mathsf{E} u_v(n,0)$. Therefore, $\mathsf{E} u_v(n,0)=v$ and hence by ergodic theorem $u_v(n,\cdot) \in \mathbb{H}'(v,v)$. \end{proof} Now we turn to the proofs of Lemma \ref{lem:small-first-step-almost-surely-all-range} and Lemma \ref{lem:uniform-integrability-of-uN}. The following lemma generalizes Lemma \ref{lem:large-deviation-for-Z-bar}: \begin{lemma} \label{lem:small-probability-for-first-large-step-outside-linear-box} There are constants $d_1>0$ and $R_0$ such that for all $r \ge R_0$, $(m,p), (n,q) \in {\mathbb Z} \times {\mathbb Z}$ ($n-m\ge 2$), with probability at least $1- e^{-d_1 r|m-n|}$, \begin{equation*} \mu_{y,x}^{m,n} \Big\{ \gamma: \max_{m\le i \le n} | \gamma_{i}-[(m,y), (n,x)]_i| \ge r|n-m| \Big\} \le e^{-r|m-n|} \end{equation*} for all $x \in [q,q+1], y \in [p,p+1]$. \end{lemma} \begin{proof} Due to shear invariance, without loss of generality we can assume $m=p=q=0$. Lemma \ref{lem:main-monotonicity} implies $\mu_{0,0}^{0,n} \preceq \mu_{y,x}^{0,n} \preceq \mu_{1,1}^{0,n}$ for any $x, y \in [0,1]$. Therefore, it suffices to show \begin{equation}\label{eq:small-prob-outside-linear-box-2} \mathsf{P} \biggl\{\mu_{0,0}^{0,n} \Big\{ \gamma: \max_{0\le i \le n } | \gamma_{i}| \ge rn/2 \Big\} > \frac{1}{2}e^{-rn} \biggr\} \le e^{-k_1 r n} \end{equation} for some constant $k_1$ and sufficiently large $r$. Let $A(r) = \{ \gamma: \max\limits_{0 \le i \le n} |\gamma_i| \ge \frac{1}{2}rn \} $. Repeating the computation in \eqref{eq:nominator-not-too-big}, we see that for sufficiently large $r$ and some constant $k_2$, \begin{equation*} \mathsf{P} \left\{ Z^{0,n}_{0,0} \bigl( A(r) \bigr) > \frac{1}{2}\rho_0^n e^{-2rn} \right\} < \frac{2n\lambda^n e^{- (r/2)^2n}}{\pi r \sqrt{n-1} \rho_0^ne^{-2rn}} \le e^{-k_2 r n}. \end{equation*} By Lemma \ref{lem:partition-function-not-too-small}, for sufficiently large $r$ and a constant $k_3$, we have \begin{equation*} \mathsf{P} \left\{ Z^{0,n}_{0,0} \le \rho_0^ne^{-rn} \right\} \le \beta^n e^{-rnr_0} \le e^{-k_3 rn}. \end{equation*} Now (\ref{eq:small-prob-outside-linear-box-2}) follows from these two estimates. \end{proof} The next statement is a direct consequence of Lemma~\ref{lem:small-probability-for-first-large-step-outside-linear-box}. \begin{lemma} \label{lem:small-probability-for-first-large-step-outside-linear-box-all-r} There are constants $d_1,d_2,R_0>0$ such that for all $(m,p), (n,q) \in {\mathbb Z} \times {\mathbb Z}$ ($n-m\ge 2$), with probability at least $1- d_2e^{-d_1 R_0|m-n|}$, \begin{equation*} \mu_{y,x}^{m,n} \Big\{ \gamma: \max_{m\le i \le n} | \gamma_{i}-[(m,y), (n,x)]_i| \ge r|n-m| \Big\} \le e^{-r|m-n|} \end{equation*} for all $x \in [q,q+1], y \in [p,p+1]$ and $r \ge R_0$, $r \in {\mathbb N}$. \end{lemma} \begin{proof}[Proof of Lemma~\ref{lem:small-first-step-almost-surely-all-range}] It suffices to prove the statement for fixed $c$ and~$(n,q)$. Let $K = 2c+R+1$. Lemma \ref{lem:one-step-control} implies that with probability one, if $\frac{n-N}{2} \ge n_1=n_1(n,q,c)$, then for all $s$ satisfying $n_1 \le s \le \frac{n-N}{2}$ and all $x\in[q,q+1]$, \begin{equation*} \mu_{\nu,x}^{N,n} \pi_{n-1}^{-1} \bigl( [-Ks, Ks]^c \bigr) \le 4\nu \bigl( [-c |N|, c|N| ]^c \bigr) + 8e^{-\sqrt{s}}. \end{equation*} This implies that for some $k_1>0$ and all $L \in [Kn_1, K(n-N)/2]$, we have \begin{equation} \label{eq:less-than-linear} \mu_{\nu, x }^{N,n } \pi_{n-1}^{-1} \bigl( [-L,L]^c \bigr) \le 4 \nu \bigl( [-c|N|, c|N|]^c \bigr) + 8 e^{-k_1\sqrt{L}}. \end{equation} Noticing that for all $L \ge K(n-N)/2$, we have the trivial inequality \begin{equation*} \mu_{\nu,x}^{N,n}\pi_{n-1}^{-1}\big( [-L,L]^c \big) \le \mu_{\nu,x}^{N,n}\pi_{n-1}^{-1}\big( [-K(n-N)/2,K(n-N)/2]^c \big), \end{equation*} we can extend \eqref{eq:less-than-linear} to all $L \in [Kn_1, 2R_0(n-N)]$ by adjusting the constant $k_1$ appropriately. Here $R_0$ is taken from Lemmas~\ref{lem:small-probability-for-first-large-step-outside-linear-box} and~\ref{lem:small-probability-for-first-large-step-outside-linear-box-all-r}. The Borel--Cantelli lemma implies that with probability one, the statement of Lemma~\ref{lem:small-probability-for-first-large-step-outside-linear-box-all-r} holds true for all sufficiently negative $m=N$ and $p \le c|N| + 1$. In particular, for sufficiently negative $N$, we have \begin{equation*} \mu_{y,x}^{N,n} \pi_{n-1}^{-1} \Bigl( \Big[x+\frac{y-x}{n-N}-r|n-N|,\, x+\frac{y-x}{n-N}+r|n-N|\Big]^c \Bigr) \le e^{-r|n-N|} \end{equation*} for all $|y| \le c|N|$ and $r \ge R_0$. Applying this estimate to $y=\pm c|N|$ and using monotonicity, we obtain for $r \ge R_0$ and sufficiently negative~$N$: \begin{align*} \notag & \mu_{\nu,x}^{N,n} \pi_{n-1}^{-1} \bigl( [-2r|n-N|, 2r|n-N|]^c \bigr) \\ \notag \le&\mu_{\nu,x}^{N,n} \pi_{n-1}^{-1} \Bigl( \big[-(|q|+c+2+r|n-N|), |q|+c+2 + r|n-N|\big]^c \Bigr) \\ \le& \mu_{\nu,x}^{N,n} \pi_{n-1}^{-1} \Bigl( \Big[x+\frac{-c|N|-x}{n-N}-r|n-N|,\, x+\frac{c|N|-x}{n-N}+r|n-N|\Big]^c \Bigr) \\ \notag \le &\nu \bigl( [-c|N|, c|N|]^c \bigr)+ 2e^{-r|n-N|}. \end{align*} Therefore, for some constant $k_2 > 0$ and all $L \in [2R_0(n-N), +\infty)$, we have \begin{equation} \label{eq:more-than-linear} \mu_{\nu,x}^{N,n} \pi_{n-1}^{-1} \bigl( [-L, L]^c \bigr) \le \nu \bigl( [-c|N|, c|N|]^c \bigr) + 2e^{- k_2 L}. \end{equation} Combining the estimates (\ref{eq:less-than-linear}) and (\ref{eq:more-than-linear}), we see that \eqref{eq:small-first-step-almost-surely-all-range} holds for all $L\ge Kn_1$, which completes the proof of the lemma. \end{proof} To prove the uniform integrability of $u^N_v(n,x)$ in Lemma~\ref{lem:uniform-integrability-of-uN}, we need an additional lemma: \begin{lemma} \label{lem:small-probability-of-first-large-step} There is a constant $s_0$ such that for $|N|/2 \ge s \ge s_0$, \begin{equation*} \mathsf{P} \left\{ \mu_{0,0}^{N,0} \pi^{-1}_{-1} ([-(R+2)s, (R+2)s]^c) \le 4e^{-\sqrt{s}} \right\} > 1 - e^{-\sqrt{s}}. \end{equation*} \end{lemma} \begin{proof} In Lemma \ref{lem:sets-for-remote-control-to-holds}, we take $\delta = 1/9$, $\beta = 1/2$, $v=1$, $(m,q) = (0,0)$ and consider the set $\Omega_2^{(s)}(1/9, 1, 0,0)$. The lemma follows from the backward version of (\ref{eq:compactness-control-by-terminal-measure}) with $v_0=v'=0$, $v_1=1$ and $\nu = \delta_0$. \end{proof} \begin{proof}[Proof of Lemma \ref{lem:uniform-integrability-of-uN}] By Lemma~\ref{lem:monotonicity_of_x-ux}, \begin{equation*} u^N_v(n,x) - x = -\int_{{\mathbb R}} y \mu_{Nv, x}^{N,n} (dy) \end{equation*} is non-increasing in $x$. Therefore, it suffices to show the uniform integrability of $ \bigl( u^N_v(n,x) \bigr)_{N < n}$ for fixed $(n,x) \in \Z\times{\mathbb R}$. We also notice that $u^N_v(0,0) \stackrel{d}{=} u^N_0(0,0) + v$. So, without loss of generality, let us assume $(n,x) = (0,0)$ and $v=0$. Let us write $f^N_{0,0,-1}(0,y) = f^N(y)$ and $u^N_0(0,0) = u^N$. Lemma~\ref{lem:small-probability-of-first-large-step} implies that if $L=(R+2)s \in [(R+2)s_0, (R+2)|N|/2]$, then \begin{equation} \label{eq:small-probability-first-step-less-than-linear} \mathsf{P} \Big\{ \int_{|y| > L } f^N(y) dy \le 4e^{-k_1\sqrt{L}} \Big\} > 1 - e^{-k_2\sqrt{L}} \end{equation} for some constants $k_1$ and $k_2$. Using the inequality $$\int_{|y| > L} f^N(y) dy \le \int_{|y| > (R+2)|N|/2} f^N(y) dy$$ for $L \ge (R+2)|N|/2$ and adjusting the constants $k_1$, $k_2$ appropriately, we can extend~(\ref{eq:small-probability-first-step-less-than-linear}) to all $L \in [(R+2)s_0, R_0|N|]$. Here, $R_0$ is taken from Lemma~\ref{lem:small-probability-for-first-large-step-outside-linear-box}. Next, Lemma~\ref{lem:small-probability-for-first-large-step-outside-linear-box} implies that if $L =r|N| \ge R_0|N|$, then \begin{equation}\label{eq:small-probability-of-first-large-step-larger-than-linear} \mathsf{P} \Big\{ \int_{|y| > L} f^N(y) dy \le e^{- L} \Big\} > 1 - e^{-d_1L}. \end{equation} Combining the estimates~(\ref{eq:small-probability-first-step-less-than-linear}) and (\ref{eq:small-probability-of-first-large-step-larger-than-linear}), we can find constants $c_1,c_2,c_3,c_4$, independent of $N$, such that for $L \ge (R+2)s_0$, \begin{equation*} \mathsf{P} \Big\{ \int_{|y| > L} f^N(y) dy \le c_1 e^{-c_2\sqrt{L}} \Big\} > 1 - c_3 e^{-c_4 \sqrt{L}}. \end{equation*} This implies that $u^N = -\int_{{\mathbb R}}y f^N(y) dy$ are uniformly integrable. \end{proof} \subsection{Uniqueness of global solutions}\label{sec:uniqueness-of-global-solutions} The main goal of this section is to finish the proof of Theorem~\ref{thm:global_solutions} by establishing the uniqueness of global solutions. Let $w(x) \in \mathbb{H}'$ and $V(x) = e^{-\int_0^x w(x')dx'}$ be its Hopf--Cole transform. We can introduce the following point-to-line polymer measures: \begin{equation*} \bar{\mu}_{V,x}^{N,n} (A_N \times ... \times A_{n-1}) = \frac{ \int_{A_N} dx_N \cdots \int_{A_{n-1}} dx_{n-1} \delta_{x}(dx_n) \ V(x_N) \prod\limits_{i=N}^{n-1} Z^{i,i+1}_{x_i, x_{i+1}} }{\int_{{\mathbb R}} V(x_N) Z^{N,n}_{x_N,x}dx_N }. \end{equation*} The fact that $w \in \mathbb{H}'$ guarantees that all integrals are finite. \begin{lemma} \label{lem:condition-second-kind-polymer-measure-converge-to-LLN} Let $\big(w_N(\cdot) \big)$ be a stationary sequence of random functions in $\mathbb{H}'$ and $\bigl( V_N (\cdot) \bigr)$ be the corresponding Hopf--Cole transforms. Let $v \in {\mathbb R}$. Suppose that one of the conditions \eqref{eq:no_flux_from_infinity}, \eqref{eq:flux_from_the_left_wins}, \eqref{eq:flux_from_the_right_wins} is satisfied by $W(\cdot)=W_N(\cdot) = \int_0^{\cdot} w_N(y')dy'$ for all $N$ with probability~$1$. Then for almost every $\omega$ and all $n \in {\mathbb Z}$, the probability measures $\nu_{N,n,x} (N < n)$ defined by \begin{equation*} \nu_{N,n,x}(dy) = \bar{\mu}_{V_N, x}^{N,n}\pi_N^{-1}(dy) = \frac{ Z^{N,n}(y,x) V_N(y)}{\int_{{\mathbb R}} Z^{N,n}(y',x) V_N(y')dy'} dy \end{equation*} satisfy \begin{equation*} \liminf_{N\to-\infty} e^{h|N|}\sup_{x \in [-L,L]} \nu_{N,n,x}( [(v+\varepsilon)N, (v-\varepsilon)N]^c ) = 0, \end{equation*} for all $L\in {\mathbb N}$ and $\varepsilon > 0$, and some constant $h(\varepsilon) > 0$ depending on $\varepsilon$. \end{lemma} We give the proof of this lemma after we derive uniqueness from it. \begin{proof}[Proof of the uniqueness part of Theorem \ref{thm:global_solutions}] Let $v \in {\mathbb R}$ and let $u_{v}(n,\cdot)$ be a stationary global solution in $\mathbb{H}'(v,v)$. We will prove that for almost every~$\omega$, $u_{v,\omega}$ coincides with the global solution constructed in Section~\ref{sec:existence-of-global-solution}. Let $V_v(n,\cdot)$ be the Hopf--Cole transforms of $u_v$ and $C_{v,m,n}$ be the family of constants such that (\ref{eq:evolution-for-V}) holds true. Let $\bar{\mu}_x^{-\infty,n}$ be defined as in \eqref{eq:def-of-mu-bar}. Then we have~\eqref{eq:solution-through-pt-to-line-measure}. Since $u_v(n,x) \in \mathbb{H}'(v,v)$, the potential of $u_v(n,x)$ satisfies one of the conditions \eqref{eq:no_flux_from_infinity}, \eqref{eq:flux_from_the_left_wins}, \eqref{eq:flux_from_the_right_wins} depending on the value of $v$. Therefore, by Lemma~\ref{lem:condition-second-kind-polymer-measure-converge-to-LLN}, we have \begin{equation*} \liminf_{N \to -\infty} \bar{\mu}_x^{-\infty,n} \pi_N^{-1}( [(v+\varepsilon)N, (v-\varepsilon)N ]^c)= 0. \end{equation*} By Theorem~\ref{thm:all-space-time-point-compactness} we have that for $m$ large enough and $n-N \ge 2m$, \begin{multline*} \bar{\mu}_x^{-\infty,n} \pi_{n-m}^{-1}( [ (v+2\varepsilon)(n-m), (v-2\varepsilon)(n-m)]^c ) \\ \le 4 \bar{\mu}_x^{-\infty,n} \pi_N^{-1} ( [(v+\varepsilon)N, (v-\varepsilon)N ]^c) + 6 e^{-\sqrt{m}}. \end{multline*} Taking $\liminf$ as $N\to -\infty$, we obtain \begin{equation*} \bar{\mu}_x^{-\infty,n} \pi_{n-m}^{-1}( [ (v+2\varepsilon)(n-m), (v-2\varepsilon)(n-m)]^c ) \le 6e^{-\sqrt{m}}. \end{equation*} So $\bar{\mu}_x^{-\infty,n}$ satisfies SLLN with slope $v$ and is supported on $S_x^{-\infty,n}$. Therefore, by Lemma \ref{lem:uniqueness-of-polymer-measures}, we have $\bar{\mu}_x^{-\infty,n} = \mu_x^{-\infty,n}(v)$. This shows that $u_v(n,\cdot)$ is exactly what we have constructed in Section~\ref{sec:existence-of-global-solution}, and the proof of uniqueness is complete. \end{proof} To prove Lemma \ref{lem:condition-second-kind-polymer-measure-converge-to-LLN} we first need several auxiliary statements. \begin{lemma} \label{lem:property-on-stationary-rvs} Let $(X_n)_{n\in{\mathbb N}}$ be a stationary sequence of random variables such that $\mathsf{P} (X_n < \infty) = 1$. Then there is a random number $k=k(\omega)$ such that \begin{equation*} \mathsf{P} \left\{ \omega: X_n(\omega) \le k(\omega) \quad \text{\rm for infinitely many $n$} \right\} = 1. \end{equation*} \end{lemma} \begin{proof} Let $A_k = \{ \omega: X_n(\omega) \le k \quad \text{for finitely many $n$} \}$. Clearly $A_{k+1} \subset A_{k}$ for all $k\in{\mathbb N}$. Let $A_{\infty} = \bigcap\limits_{k=1}^{\infty} A_{k}$. We want to prove that $\mathsf{P} (A_{\infty}) = 0$. By the ergodic theorem, on $A_{\infty}$, for any $k$, we have \begin{equation*} 0 = \lim_{n \to \infty} \frac{1}{n}\sum_{i=0}^{n-1} {\mathbf{1}}_{X_i \le k} = \mathsf{E} ({\mathbf{1}}_{X_0 \le k} | \mathcal{I}), \end{equation*} where $\mathcal{I}$ is the invariant $\sigma$-algebra for the stationary sequence $\big(X_n\big)$. Therefore \begin{equation*} 0 = \mathsf{E} \big({\mathbf{1}}_{A_{\infty}} \mathsf{E} ({\mathbf{1}}_{X_0 \le k}| \mathcal{I}) \big) = \mathsf{E} {\mathbf{1}}_{A_{\infty}}{\mathbf{1}}_{X_0 \le k}. \end{equation*} Since $\mathsf{P}(X_0 < \infty)=1$, by the Bounded Convergence Theorem we have \begin{equation*} 0 = \lim_{k\to \infty} \mathsf{E} {\mathbf{1}}_{A_{\infty}}{\mathbf{1}}_{X_0 \le k} = \mathsf{P} (A_{\infty}) \end{equation*} as desired. \end{proof} \begin{lemma} \label{lem:truncation-and-concentration-almost-surely-true} There is a full measure set $\Omega''$ such that the following is true for every $\omega \in \Omega''$. For all $c > 4\sqrt{\ln (2\lambda/\rho_0)})$ and $(n,q) \in {\mathbb Z} \times {\mathbb Z}$, there is a constant $m_0 = m_0(n,q,c)$ such that for all $m > m_0$, we have \begin{equation} \label{eq:truncation-almost-surely} \frac{\int_{|y| \ge cm} Z^{n-m,n}(y,q)e^{c|y|/17} dy }{\int_q^{q+1} Z^{n-m,n}(y,q)dy} \le 2^{-m} \end{equation} and \begin{equation} \label{eq:concentration-almost-surely} \Big| \ln \int_I Z^{n-m,n}(y,x)H(y) dy - \ln \int_I e^{\alpha(m,x-y)} H(y) dy \Big| \le 2m^{3/4}, \end{equation} for all intervals $I \subset [-cm,cm]$, all $x \in [q,q+1]$ and all positive functions $H(\cdot)$. Here, $\alpha(\cdot, \cdot)$ has been defined in~\eqref{eq:space-time-shape}. \end{lemma} We will prove this lemma after we use it to derive Lemma~\ref{lem:condition-second-kind-polymer-measure-converge-to-LLN}. We also need a monotonicity statement about point-to-line polymer measures. \begin{lemma} \label{lem:monotonicity-for-second-kind-polymer-measure} Let $x<x'$ and $V(x)$ be a positive function that grows at most exponentially. Then for any $m,n$ with $m<n$, the polymer measure $\bar{\mu}_{V,x}^{m,n}$ is stochastically dominated by $\bar{\mu}_{V,x'}^{m,n}$. \end{lemma} \begin{proof} First, we have \begin{equation*} \mu_{V,y}^{m,k}(A_m \times \dotsb \times A_{k-1}) = \int_{A_{k-1}} \bar{\mu}_{V,y}^{m,k}\pi_{k-1}^{-1}(dx_{k-1}) \bar{\mu}_{V,x_{k-1}}^{m,k-1}(A_m \times \dotsb \times A_{k-2}). \end{equation*} Therefore, similarly to Lemma \ref{lem:dominance-1}, it suffices to show that $\bar{\mu}_{V, x}^{m,n}\pi_{n-1}^{-1} \preceq \bar{\mu}_{V, x'}^{m,n}\pi_{n-1}^{-1}$ and use an induction argument. Now we compute the marginals at time $n-1$ : \begin{align*} \bar{\mu}_{V,x}^{m,n}\{X_{n-1}\le r\} &= \frac{ \int_{{\mathbb R}} dy\int_{(-\infty,r]} d\eta V(y)Z_{y,\eta}^{m,n-1} e^{-F_{n-1}(\eta)} g(\eta-x) }{ \int_{{\mathbb R}} dy\int_{{\mathbb R}} d\eta V(y)Z_{y,\eta}^{m,n-1} e^{-F_{n-1}(\eta)} g(\eta-x) } . \end{align*} Let \begin{equation*} \nu(d\eta) = \int dy V(y) Z^{m,n-1}_{y,\eta} e^{-F_{n-1}(\eta)} d\eta. \end{equation*} Then by Lemma \ref{lem:abstract-monotonicity}, $\bar{\mu}_{V,x}^{m,n}\{X_{n-1}\le r\}$ is decreasing in $x$, so $\bar{\mu}_{V, x}^{m,n}\pi_{n-1}^{-1}$ is dominated by $\bar{\mu}_{V, x'}^{m,n}\pi_{n-1}^{-1}$. \end{proof} \begin{proof}[Proof of Lemma~\ref{lem:condition-second-kind-polymer-measure-converge-to-LLN}] We take $\Omega''$ from the statement of Lemma \ref{lem:truncation-and-concentration-almost-surely-true} and fix an arbitrary $\omega \in \Omega''$. Fix $n$ and $L \in {\mathbb N}$. Since Lemma \ref{lem:monotonicity-for-second-kind-polymer-measure} implies $\nu_{N,n,-L} \preceq \nu_{N,n,x} \preceq \nu_{N,n,L}$ for $x \in [-L, L]$ , it suffices to show that for every $\varepsilon > 0$, \begin{equation*} \liminf_{N\to -\infty} e^{h|N|}\max_{a = \pm L} \nu_{N,n,a} \big( [(v+\varepsilon)N, (v-\varepsilon)N ]^c \big) = 0, \end{equation*} or, equivalently, that for every $ \varepsilon \in (0,1)$ there is a random sequence $m_k = m_k(\omega,\varepsilon) \uparrow + \infty$ such that \begin{equation} \label{eq:what-to-prove} \begin{split} 0 &=\lim_{k\to \infty}e^{hm_k} \nu_{n-m_k,n,a}\bigl( [-(v+\varepsilon)m_k, -(v-\varepsilon)m_k]^c \bigr)\\ &= \lim_{k\to\infty} e^{hm_k}\frac{\int_{|y+vm_k| > \varepsilon m_k} Z^{n-m_k,n}(y,a) V_{n-m_k} (y) dy }{ \int_{{\mathbb R}} Z^{n-m_k,n}(y,a) V_{n-m_k}(y)dy} \end{split} \end{equation} for $a = \pm L$. The proof consists of two steps. The first step is to use \eqref{eq:no_flux_from_infinity}, \eqref{eq:flux_from_the_left_wins}, \eqref{eq:flux_from_the_right_wins} and Lemma \ref{lem:property-on-stationary-rvs} to find a random sequence $(m_k)$ with certain properties; the second is to combine those properties and estimates provided by Lemma \ref{lem:truncation-and-concentration-almost-surely-true} to derive~\eqref{eq:what-to-prove}. We can assume that $v\ge 0$, since the case $v<0$ is totally symmetric to the case $v>0$. Let us fix some $\delta > 0$ such that \begin{equation} \label{eq:choice-of-delta} \delta < \begin{cases} \varepsilon/4, & v =0, \\ (\varepsilon/4) \wedge (v/2) \wedge \frac{\varepsilon^2}{8v}, & v > 0. \end{cases} \end{equation} \textbf{Step 1 --- find $(m_k)$:} we claim that there is a random constant $R=R(\omega)$ and a random sequence $(m_k)$ such that for every $m = m_k$, \begin{equation} \label{eq:boundedness-of-W} |W_{n-m}(y)| \le R, \quad y \in [-L,L+1], \end{equation} \begin{equation} \label{eq:W-finite-slope} W_{n-m}(y) \ge -R(|y|+1) , \quad y \in {\mathbb R}, \end{equation} and \begin{subequations} \label{eq:W-asymtopic-slope-at-infinity} \begin{align} \nonumber v&= 0, \\ \label{eq:zero-v-infinity-slope} W_{n-m}(y) &\ge -\delta|y|, && |y| \ge R, \\ \nonumber \text{or,}& \\ \nonumber v &> 0, \\ \label{eq:positive-v-negative-infinite-slope} |W_{n-m}(y) - vy| &\le \delta|y|, && y < -R,\\ \label{eq:positive-v-plus-infinity-slope} W_{n-m}(y) &\ge (-v+2\delta)|y|, && y > R. \end{align} \end{subequations} To see this, for each $m$, we let $X_m$, $Y_m$ and $Z_m$ be the infimum of $R$ such that \eqref{eq:boundedness-of-W}, \eqref{eq:W-finite-slope} and \eqref{eq:W-asymtopic-slope-at-infinity} are satisfied. Due to stationary of $W_N(\cdot)$, $(X_m)$, $(Y_m)$ and $(Z_m)$ are all stationary sequences of random variables. Also, $X_m $ are a.s. finite because $W_N(\cdot)$ are locally finite; $Y_m $ are a.s. finite because $W_N(\cdot) \in \mathbb{H}$; $Z_m$ are a.s. finite due to~\eqref{eq:no_flux_from_infinity} or~\eqref{eq:flux_from_the_left_wins}, depending on $v$. Therefore, by Lemma \ref{lem:property-on-stationary-rvs}, there is a random number $R=R(\omega)$ such that $X_m\vee Y_m\vee Z_m \le R$ for infinitely many $m$ almost surely. This proves the claim. \textbf{Step 2 --- show (\ref{eq:what-to-prove})}. For simplicity we will write $m = m_k$ in what follows, so $m \to \infty$ actually means $m=m_k$, $k \to \infty$. Let us fix $c \ge 17( v+1) \vee 4 \sqrt{\ln(2\lambda/\rho_0)}$ and write \begin{align*} &\quad\nu_{n-m,n,a}([-(v+\varepsilon) m, -(v-\varepsilon) m]^c) \\ &= \frac{\int_{|y|<cm, |y+vm| > \varepsilon m} Z^{n-m,n}_{y,a}e^{-W_{n-m}(y) } dy }{ \int_{{\mathbb R}} Z^{n-m,n}_{y,a}e^{-W_{n-m}(y) }dy} + \frac{\int_{|y| \ge cm} Z^{n-m,n}_{y,a}e^{-W_{n-m}(y) } dy }{ \int_{{\mathbb R}} Z^{n-m,n}_{y,a} e^{-W_{n-m}(y) } dy} \\ &= A^m + B^{m}. \end{align*} We will show that both $A^m$ and $B^m$ decay exponentially. First we look at $B^m$. By \eqref{eq:W-asymtopic-slope-at-infinity}, if $m$ is sufficiently large, then $-W_{n-m}(y) \le (|v|+\delta) |y| \le c|y|/17$ for all $|y| \ge cm$. Due to \eqref{eq:boundedness-of-W} we have \begin{equation*} \int_{{\mathbb R}} Z^{n-m,n}(y,a)e^{-W_{n-m}(y) }dy \ge e^{-R}\int_a^{a+1} Z^{n-m,n}(y,a)dy. \end{equation*} Therefore, by Lemma \ref{lem:truncation-and-concentration-almost-surely-true} we have \begin{equation* B^m \le e^R\frac{\int_{|y| \ge cm} Z^{n-m,n}(y,a)e^{c|y|/17 } dy }{ \int_a^{a+1} Z^{n-m,n}(y,a) dy} \le \frac{e^R}{2^m} \end{equation*} for sufficiently large $m$. Next we look at $A^m$. Using Lemma~\ref{lem:truncation-and-concentration-almost-surely-true}, we obtain that for sufficiently large~$m$, \begin{align*} A^m & \le \exp\big( 4m^{3/4}\big)\cdot \frac{\int_{|y|<cm, |y+vm| > \varepsilon m} e^{-\frac{(y-a)^2}{2m}-W_{n-m}(y) } dy }{ \int_{-cm}^{cm} e^{-\frac{(y-a)^2}{2m}-W_{n-m}(y) }dy} \\ &\le \exp\big( 4m^{3/4} + L^2/m + 2Lc \big)\cdot \frac{\int_{|y+vm| > \varepsilon m} e^{-\frac{y^2}{2m}-W_{n-m}(y) } dy }{ \int_{-cm}^{cm} e^{-\frac{y^2}{2m}-W_{n-m}(y) }dy}. \end{align*} Let us denote the ratio of integrals in the last line by $\tilde{A}^m$. It suffices to show that~$\tilde{A}^m$ decays exponentially. We will consider the cases $v=0$ and $v>0$ separately. Suppose $v=0$. For sufficiently large $m$, we have $W_{n-m}(y) \ge -\delta |y|$ for all $|y| > \varepsilon m$ by \eqref{eq:zero-v-infinity-slope} and $W_{n-m}(y) \le R$ for $y \in [0,1]$ by \eqref{eq:boundedness-of-W}. Therefore, \begin{equation*} \tilde{A}^m \le \frac{\int_{|y|> \varepsilon m } e^{- \frac{y^2}{2m} + \delta |y| } dy}{ \int_0^1 e^{-\frac{y^2}{2m} - R} } \le e^{\frac{1}{2m} + R} \int_{|y| > \varepsilon m } e^{-\frac{y^2}{2m} + \delta |y| } dy \le e^{\frac{1}{2m} + R} \cdot \frac{4}{\varepsilon} e^{- \frac{\varepsilon^2}{4}m} \end{equation*} as desired. Here, in the last inequality, we used $\delta < \varepsilon /4$ to obtain \begin{equation} \label{eq:gaussian-integral-estimate} \int_{|y| > \varepsilon m} e^{-\frac{y^2}{2m} + \delta |y|} dy \le \int_{|y| > \varepsilon m} e^{-\frac{y^2}{4m}} dy \le \frac{4}{\varepsilon} e^{-\frac{\varepsilon^2 }{4}m}. \end{equation} Suppose $v > 0$. Let $\tilde{A}^m = (A_1+A_2+A_3)/A_4$, where \begin{align*} A_1 &= \int_{ |y+vm| > \varepsilon m, y \le -R} e^{-\frac{y^2}{2m}-W_{n-m}(y)} dy, & A_2&= \int_{-R}^Re^{-\frac{y^2}{2m}-W_{n-m}(y)}dy, \\ A_3 &= \int_{R}^{\infty} e^{-\frac{y^2}{2m}-W_{n-m}(y)} dy, & A_4&= \int_{-cm}^{cm} e^{-\frac{y^2}{2m}-W_{n-m}(y)}dy. \end{align*} For sufficiently large $m$, by~(\ref{eq:positive-v-negative-infinite-slope}), (\ref{eq:gaussian-integral-estimate}), (\ref{eq:W-finite-slope}) and~(\ref{eq:positive-v-plus-infinity-slope}), we have \begin{align*} A_1 &\le \int_{ |y+vm| > \varepsilon m} e^{-\frac{y^2}{2m} - (v+\delta) y} dy \\&\le e^{( \frac{v^2}{2} + v\delta ) m} \int_{|y'| > \varepsilon m} e^{-\frac{y'^2}{2m} + \delta |y'|} dy' \le \frac{4}{\varepsilon} \exp \bigl( (v^2/2 + v\delta- \varepsilon^2/4) m \bigr), \end{align*} \begin{equation*} A_2 \le \int_{-R}^R e^{-\frac{y^2}{2m} + R(|y| +1) } dy \le 2Re^{R^2 +R}, \end{equation*} \begin{equation*} A_3 \le \int_R^{\infty} e^{-\frac{y^2}{2m} + (v-2\delta) y} dy \le\int_{-\infty}^{\infty} e^{-\frac{y^2}{2m} + (v-2\delta) y} dy = \sqrt{2m\pi} \exp \Bigl( \frac{(v-2\delta)^2}{2}m \Bigr), \end{equation*} and \begin{equation*} A_4 \ge \int_{-vm}^{-vm+1} e^{-\frac{y^2}{2m} + (v-\delta)y} dy \ge \exp \bigl( (v^2/2 - v\delta) m \bigr). \end{equation*} Therefore, \begin{multline*} \tilde{A}^m \le \frac{4}{\varepsilon}\exp \bigl( -(\varepsilon^2/4 - 2v\delta)m \bigr) \\+ 2Re^{R^2+R} \exp \bigl( -(v^2/2-v\delta) m \bigr) + \sqrt{2m\pi} \exp \bigl( -(v\delta - 2\delta^2)m \bigr), \end{multline*} and the right-hand side decays exponentially due to (\ref{eq:choice-of-delta}). \end{proof} Now we turn to the proof of~Lemma \ref{lem:truncation-and-concentration-almost-surely-true}. We begin with~\eqref{eq:truncation-almost-surely}. \begin{proof}[Proof of~\eqref{eq:truncation-almost-surely}] Let us fix $(n,q)$ and~$c$. Due to the Borel--Cantelli lemma and the fact that for sufficiently large $m$, \begin{equation*} \int_{|y| \ge cm } Z^{n-m,m}(y,q) e^{c|y|/17} dy \le \int_{|y-q| \ge cm/2} Z^{n-m,n}(y,q) e^{c|y-q|/16} dy, \end{equation*} the inequality~\eqref{eq:truncation-almost-surely} will follow if we prove that for some constant $k>0$ and sufficiently large $m$, \begin{equation} \label{eq:small-probability} \mathsf{P} \bigg\{ \frac{a_m}{b_m} > 2^{-m } \bigg\} \le e^{-k m}, \end{equation} where \begin{equation*} a_m = \int_{|y-q| \ge cm/2} Z^{n-m,n}(y,q)e^{c|y-q|/16} dy, \quad b_m = \int_q^{q+1} Z^{n-m,n}(y,q) dy. \end{equation*} By Lemma \ref{lem:partition-function-not-too-small}, we have $\mathsf{P} \{b_m \le \rho_0^m\} \le e^{-k_1m}$ for some constant $k_{1}$. By Markov inequality, \begin{equation*} \begin{split} \mathsf{P} \big\{ a_m \ge (\rho_0/2)^m \big\} \le \Big(\frac{2}{\rho_0}\Big)^m \mathsf{E} a_m =& \Big(\frac{2\lambda}{\rho_0}\Big)^m \int_{|y-q| \ge cm/2} \frac{1}{\sqrt{2m\pi}} e^{-\frac{(y-q)^2}{2m} + c|y-q|/16}dy \\ \le& \Big(\frac{2\lambda}{\rho_0}\Big)^m \int_{|y-q| \ge cm/2} \frac{1}{\sqrt{2m\pi}} e^{-\frac{(y-q)^2}{4m}}dy \\ \le& \frac{8}{c \sqrt{2m\pi}} e^{-\big( c^2/16- \ln (2 \lambda/\rho_0) \big) m} dy \le e^{-k_2m} \end{split} \end{equation*} for a constant $k_2>0$ if $c > 4 \sqrt{\ln(2\lambda/\rho_0)}$ and $m$ is sufficiently large. Combining these two inequalities, we obtain (\ref{eq:small-probability}) and complete the proof of~\eqref{eq:truncation-almost-surely}. \end{proof} To prove~\eqref{eq:concentration-almost-surely}, the second part of Lemma~\ref{lem:truncation-and-concentration-almost-surely-true}, we need auxiliary statements. \begin{lemma} \label{lem:continuity-of-partition-function-wrt-end-point} There is a number $K$ such that, with probability one, for any $c > 0$ and $(n,q) \in {\mathbb Z} \times {\mathbb Z}$, and for sufficiently large $m$, \begin{equation} \label{eq:continuity-of-partition-functions} \big| \ln Z^{n-m,n}(y, x) - \ln Z^{n-m,n}([y],[x]) \big| \le K \sqrt{m},\quad [x]=q,\quad |[y]| \le cm. \end{equation} \end{lemma} \begin{proof} Without loss of generality we can assume $n = 0$. Let us define \[ C_{x,y}^m=\bigg\{\mu_{y,x}^{-m,0} \{ \gamma: |\gamma_{-m+1} - [y]| \vee |\gamma_{-1} - [x]| \ge (R'+1) \sqrt{m} \} \le 2^{-m+1}\bigg\}. \] Due to Lemma \ref{lem:small-probability-for-single-square-root-jump} and $\mu_{p, q}^{-m, 0} \preceq \mu_{y,x}^{-m,0} \preceq \mu_{p+1,q}^{-m,0}$ for $[x]=q$, $[y]=p$, we have \[ \mathsf{P} \Big({\textstyle \bigcup_{[x]=p,\ [y]=q}}\, (C_{x,y}^m)^c \Big) \le 2e^{-r_2m} \] for sufficiently large $m$ and all $|p| \le cm$. Then, the Borel--Cantelli lemma implies that, with probability one, there is $m_1$ such that for $m>m_1$, all $[x]=q$ and $|[y]| \le cm$, $C_{x,y}^m$ holds. Also, using~\ref{item:exponential-moment-for-maximum}, Markov's inequality, and the Borel--Cantelli lemma, we obtain that with probability one, there is $m_2$ such that for $m>m_2$, \begin{equation} \label{eq:not-large-potential-almost-surely} |F_{-m}(y) | \le \sqrt{m}, \quad |[y]| \le cm. \end{equation} Assume $m \ge 2\vee m_1\vee m_2$. We consider only~$x$ and~$y$ such that $[x] = q$ and $|[y]| \le cm$. Let us fix $p=[y]$ and define \begin{equation}\label{eq:definition-of-Z-tilda-minus-m-0-y-x} \tilde{Z}^{-m,0}_{y,x} = \int_{(z_1,z_2) \in E}\,dz_1dz_2 e^{-\frac{1}{2}(y-z_1)^2 - F_{-m}(y) - \frac{1}{2}(x-z_2)^2 - F_{-1}(z_2)} Z^{-m+1,-1}_{z_1,z_2} \end{equation} where $E = \{(z_1,z_2)\, : \, |z_1 - p|\vee | z_2 -q| < (R'+1)\sqrt{m}\}$. Since $m>m_1$, we have $ \bigl| 1 - \tilde{Z}^{-m,0}_{y,x}/Z^{-m,0}_{y,x} \bigr| \le 2^{-m+1}$ and hence \begin{equation} \label{eq:diff-of-Z-tilde-and-Z} \big| \ln Z^{-m,0}_{y,x} - \ln \tilde{Z}^{-m,0}_{y,x} \big| \le 2^{-m+2}, \end{equation} since $|\ln (1+x)| \le 2|x|$ for $|x| \le 1/2$. Using~(\ref{eq:not-large-potential-almost-surely}) and~(\ref{eq:definition-of-Z-tilda-minus-m-0-y-x}), we see that \begin{equation} \label{eq:continuity-of-Z-tilde} \begin{split} &\big| \ln \tilde{Z}^{-m,0}_{y,x} - \ln \tilde{Z}^{-m,0}_{p,q} \big|\\ \le &2 \sqrt{m} + \sup_{(z_1, z_2) \in E}\bigg( \biggl| \frac{(y-z_1)^2}{2} - \frac{(p-z_1)^2}{2} \biggr| + \biggl| \frac{(x-z_2)^2}{2} - \frac{(q-z_2)^2}{2} \biggr| \bigg)\\ \le & 2\sqrt{m} + 2 \bigl( (R'+1) \sqrt{m} + 1 \bigr). \end{split} \end{equation} The lemma follows from~(\ref{eq:diff-of-Z-tilde-and-Z}), (\ref{eq:continuity-of-Z-tilde}) and the triangle inequality. \end{proof} \begin{lemma} \label{lem:concentration-simultaneouly-for-one-point} With probability one, for all $c > 0$ and $(n,q) \in {\mathbb Z} \times {\mathbb Z}$, we have \begin{equation*} |\ln Z^{n-m,n}(y,x) - \alpha(m,x-y)| \le 2 m^{3/4},\quad [x]=q,\quad |[y]|\le cm, \end{equation*} for sufficiently large $m$. \end{lemma} \begin{proof} Let us fix $c$ and $(n,q)$. By Theorem \ref{thm:concentration-of-free-energy} and shear invariance, for sufficiently large $m$ and all~$p \in {\mathbb Z}$, we have \begin{equation*} \mathsf{P} \Big\{ \big| \ln Z^{n-m,n}(p,q) - \alpha(m,q-p) \big| \ge m^{3/4} \Big\} \le e^{-m^{1/3}}, \end{equation*} The Borel--Cantelli lemma implies that almost surely, \begin{equation} \label{eq:concentration-for-partition-function-almost-surely} \big| \ln Z^{n-m,n}(p,q) - \alpha(m,q-p) \big| \le m^{3/4}, \end{equation} for sufficiently large $m$ and all $|p| \le cm$. Also, if $[x]=q$ and $|[y]| \le cm$, we have \begin{equation} \label{eq:Lip-continuous-for-alpha} \begin{split} | \alpha(m,[x]-[y]) - \alpha(m,x - y) | \le &\biggl| \frac{([x]-[y])^2}{2m} - \frac{(x-y)^2}{2m} \biggr| \\ \le & \frac{2(|q|+1) + 2 (cm+1)}{m}. \end{split} \end{equation} The lemma follows from Lemma \ref{lem:continuity-of-partition-function-wrt-end-point}, (\ref{eq:concentration-for-partition-function-almost-surely}) and (\ref{eq:Lip-continuous-for-alpha}) \end{proof} Now, the inequality~\eqref{eq:concentration-almost-surely} follows from Lemma~\ref{lem:concentration-simultaneouly-for-one-point}, which completes the proof of Lemma~\ref{lem:truncation-and-concentration-almost-surely-true} and concludes the entire uniqueness part. \subsection{Basins of pullback attraction}\label{sec:pullback} The global solutions play the role of one-point pullback attractors. The goal of this section is to prove Theorem \ref{thm:pullback_attraction}. First we need a version of Lemma \ref{lem:condition-second-kind-polymer-measure-converge-to-LLN} where $w_N\equiv w$ are independent of $N$, which is the case in Theorem \ref{thm:pullback_attraction}. \begin{lemma} \label{lem:LLN-for-p2l-polymer-measure-non-random} Let $v \in {\mathbb R}$ and $w(\cdot) \in \mathbb{H}'$. If one of the conditions \eqref{eq:no_flux_from_infinity}, \eqref{eq:flux_from_the_left_wins}, \eqref{eq:flux_from_the_right_wins} is satisfied by $W(\cdot)= \int_0^{\cdot} w(y')dy'$, then for almost every $\omega$ and every $n \in {\mathbb Z}$, the probability measures $\nu_{N,n,x} (N < n)$ defined by \begin{equation*} \nu_{N,n,x}(dy) = \frac{ Z^{N,n}(y,x) e^{-W(y)}}{\int_{{\mathbb R}} Z^{N,n}(y',x) e^{-W(y')}dy'} dy \end{equation*} satisfy \begin{equation*} \lim_{N\to-\infty} e^{h|N|}\sup_{x \in [-L,L]} \nu_{N,n,x}( [(v+\varepsilon)N, (v-\varepsilon)N]^c ) = 0, \end{equation*} for all $L\in {\mathbb N}$ and $\varepsilon > 0$, and some constant $h(\varepsilon) > 0$ depending on $\varepsilon$. \end{lemma} \begin{proof} The proof is similar to that of Lemma~\ref{lem:condition-second-kind-polymer-measure-converge-to-LLN}. Because $w_N(\cdot) \equiv w(\cdot)$ are independent of $N$, there is no need to choose a subsequence $(m_k)$ to satisfy \eqref{eq:W-asymtopic-slope-at-infinity}, \eqref{eq:boundedness-of-W}, and \eqref{eq:W-finite-slope} as we did in the first step of proving Lemma~\ref{lem:condition-second-kind-polymer-measure-converge-to-LLN}. Therefore, we obtain $\lim$ instead of $\liminf$ in the conclusion. \end{proof} \begin{proof}[Proof of Theorem \ref{thm:pullback_attraction}] We define $\hat\Omega=\bar{\Omega} \cap \Omega'' \cap \Omega_v'$ and let $\omega\in\hat\Omega$. We also define $V(x) = e^{-\int_0^{x } w(x') dx'}$ and consider the measures \begin{equation*} \nu_{N,n,x}(dy) = \bar{\mu}_{V, x}^{N,n}\pi_N^{-1}(dy) = \frac{V(y) Z^{N,n}(y,x)}{\int_{{\mathbb R}} V(y') Z^{N,n}(y',x)dy'} dy. \end{equation*} Then we have $\bar{\mu}_{V,x}^{N,n} = \mu_{\nu_{N,n,x},x}^{N,n}$ and \begin{equation*} \Psi^{N,n}_{\omega} w(x) = \int_{{\mathbb R}} (y-x) \bar{\mu}_{V, x}^{N,n}\pi^{-1}_{n-1} (dy). \end{equation*} Due to Lemma~\ref{lem:monotonicity_of_x-ux}, it suffices to prove pointwise convergence, i.e., to show that \begin{equation} \label{eq:what-to-prove-pullback} \lim_{N\to -\infty} \int_{{\mathbb R}} (y-x) \bar{\mu}_{V, x}^{N,n}\pi^{-1}_{n-1} (dy) = \int_{{\mathbb R}} (y-x) \mu_{x}^{-\infty,n}(v)\pi_{n-1}^{-1}(dy),\quad x\in{\mathbb R}. \end{equation} Using Lemmas~\ref{lem:LLN-for-p2l-polymer-measure-non-random} and~\ref{lem:small-first-step-almost-surely-all-range}, we obtain that for some constants $b_1$ and $b_{2}$, \begin{equation} \label{eq:growth-condition-for-g} \bar{\mu}_{V, x}^{N,n}\pi^{-1}_{n-1}([-L,L]^c) \le b_1 e^{-b_2 \sqrt{L}}. \end{equation} By Lemma \ref{lem:LLN-for-p2l-polymer-measure-non-random}, for fixed $(n,x) \in {\mathbb Z}\times{\mathbb R}$, $(\nu_{N,n,x})_{N<n}$ is a family of probability measures satisfying LLN with slope $v$. Hence by Lemma \ref{lem:uniqueness-of-polymer-measures}, $\mu_{\nu_{N,n,x},x}^{N,n}$ converges weakly to $\mu_x^n(v)$, so $\bar{\mu}_{V,x}^{N,n} \pi_{n-1}^{-1}$ converges weakly to $\mu_{x}^n(v)\pi_{n-1}^{-1}$. Now (\ref{eq:what-to-prove-pullback}) follows from this and~(\ref{eq:growth-condition-for-g}), and the proof is complete. \end{proof} \section{Overlap of polymer measures} \label{sec:overlap} In this section we prove Theorem~\ref{thm:convergence-total-variation}. In fact, we give a proof for the time-reversed version of the theorem because this is more convenient. We recall that \begin{equation*} \| \mu - \nu \|_{TV} = \sup_{A \in \mathcal{B}({\mathbb R})} |\mu(A) - \nu(A)|. \end{equation*} and that $\Omega_v'=\Omega'\cap \Omega_v$. The convergence of polymer measures in total variation distance is a consequence of the existence of ratios of partition functions and the LLN for polymer measures. For the rest of this section, we fix $v \in {\mathbb R}$ and always assume that $\omega\in\Omega'_v$. We will also fix $(n_1,x_1)$ and $(n_2,x_2)$, and write~$\mu_i^N = \mu_{x_i}^{n_i} \pi^{-1}_N$, $i=1,2$. \begin{lemma} \label{lem:TV-distance-controlled-by-ratio-of-density} Let $\mu$ and $\nu$ be two probability measures with densities $f$ and $g$ respectively, such that both $f$ and $g$ are positive on some Borel set $C$, and zero outside~$C$. Then \begin{equation*} \| \mu - \nu \|_{TV} \le 1- \inf_{x\in C} \frac{g(x)}{f(x)} \end{equation*} \end{lemma} \begin{proof} Let $A = \{ x \in C: f(x) \ge g(x)\}$ and $d = \inf_{x\in C} g(x)/f(x)$. Then \begin{equation*} \| \mu - \nu \|_{TV} = \int_A (f(x)-g(x)) \,dx \le \int_A (1-d)f(x)\, dx \le (1-d)\int_C f(x) \, dx = 1-d. \end{equation*} \end{proof} \begin{lemma} \label{lem:sublinear-growth-of-polymer-meausres} There are constants $\alpha_N, \beta_N$ depending on $\omega, x_i, n_i$ such that \begin{align*} \lim_{N\to -\infty} \frac{\alpha_N}{N}&=\lim_{N\to -\infty} \frac{\beta_N}{N}=v, \\ \lim_{N\to -\infty}\mu_1^N([\alpha_N, \beta_{N}]^c) &= \lim_{N\to -\infty}\mu_2^N([\alpha_N, \beta_{N}]^c) =0. \end{align*} \end{lemma} \begin{proof} Since the measures $\mu_i^{N}$ satisfy the LLN with slope $v$, there is a decreasing sequence of negative numbers $(N_k)$ such that \begin{equation*} \mu_i^N ([ (-v-2^{-k})|N|, (-v+2^{-k})|N|]^c) \le 2^{-k}, \quad N \le N_k,\ i=1,2. \end{equation*} For every $N$, let $k$ be such that $N_{k+1} \le N < N_k$. Then setting \begin{equation*} \alpha_N = (-v-2^{-k})|N|, \quad \beta_N = (-v+2^{-k})|N| \end{equation*} completes the proof. \end{proof} Let $f_i^N(x)$ be the density of $\mu_i^N$. We will need the following representation of~$f_i^N$. \begin{lemma} \label{lem:representation-through-solution-to-burgers-equation} Recall the function $V_v(N,x)$ which is the Hopf--Cole transform of the global solution $u_v(n,x)$. Then \begin{equation*} f_i^N(x) = \frac{ V_v(N,x) Z_{x,x_i}^{N,n_i}}{\int V_v(N,x') Z_{x',x_i}^{N,n_i} \, dx'}. \end{equation*} \end{lemma} \begin{proof} By (\ref{eq:expression-of-polymer-measure-density}) in Theorem \ref{thm:convergence-of-density-functions} we have \begin{equation*} f_i^N(x) = Z^{N,n_i}_{x,x_i} G_v \bigl( (N,x), (n_i,x_i) \bigr). \end{equation*} Thus for $x \neq y$, \begin{equation*} \frac{f_i^N(x)}{f_i^N(y)} = \frac{Z^{N,n_i}_{x,x_i}}{Z^{N,n_i}_{y,x_i}} \frac{G_v \bigl( (N,x), (n_i,x_i) \bigr)}{G_v \bigl( (N,y), (n_i,x_i) \bigr)} = \frac{Z^{N,n_i}_{x,x_i}}{Z^{N,n_i}_{y,x_i}} \frac{G_v \bigl( (N,x), (N,0) \bigr)}{ G_v \bigl( (N,y), (N,0) \bigr)} = \frac{Z^{N,n_i}_{x,x_i} V_v(N,x)}{Z^{N,n_i}_{y,x_i}V_v(N,y)}, \end{equation*} and our claim follows. \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:convergence-total-variation}] Let $D_i^N = \mu_i^N([\alpha_N,\beta_N])$, $i=1,2$, and let \begin{equation*} \tilde{\mu}_i^N (A) = (D_i^{N} )^{-1} \mu_i^N( A \cap [\alpha_N, \beta_N]). \end{equation*} Then the measures $\tilde{\mu}_i^N$, $i=1,2$, are supported on $[\alpha_N,\beta_N]$ with densities given by $\tilde{f}_i^N(x)=(D_i^N)^{-1} f_i^N(x)$. Also, \begin{equation} \label{eq:modified-measures} \| \tilde{\mu}_i^N - \mu_i^N \|_{TV} \le 1-D_i^{N}. \end{equation} Combining this with Lemma \ref{lem:TV-distance-controlled-by-ratio-of-density} we obtain \begin{equation*} \begin{split} \| \mu_1^N - \mu_2^N \|_{TV} &\le \| \mu_1^N - \tilde{\mu}_1^N \|_{TV} + \| \tilde{\mu}_1^N - \tilde{\mu}_2^N \|_{TV} + \| \tilde{\mu}_2^N - \mu_2^N \|_{TV} \\ &\le 3-D^N_1-D^N_2 - \inf_{x \in [\alpha_N, \beta_N]} \frac{\tilde{f}_2^N(x)}{\tilde{f}_1^N(x)}. \end{split} \end{equation*} Since $D^{N}_i \to 1$ as $N\to -\infty$, $i=1,2$, it suffices to show \begin{equation*} \lim_{N \to -\infty} \inf_{x \in [\alpha_N, \beta_N]} \frac{\tilde{f}_2^N(x)}{\tilde{f}_1^N(x)} = 1. \end{equation*} Using the representation of $f^N_i$ in Lemma \ref{lem:representation-through-solution-to-burgers-equation}, we see that \begin{equation*} \tilde{f}_i^N = \frac{ V_v(N,x) Z_{x,x_i}^{N,n_i}}{\int_{\alpha_N}^{\beta_{N}} V_v(N,x') Z_{x',x_i}^{N,n_i} \, dx'}. \end{equation*} and hence \begin{equation*} \frac{\tilde{f}_2^N(x)}{\tilde{f}_1^N(x)} = \frac{Z^{N,n_2}_{x,x_2}}{Z^{N,n_1}_{x,x_1}} \frac{\int_{\alpha_N}^{\beta_{N}} V_v(N,x') Z_{x',x_1}^{N,n_1} \, dx'}{\int_{\alpha_N}^{\beta_{N}} V_v(N,x') Z_{x',x_2}^{N,n_2} \,dx'} \ge \frac{m_N}{M_{N}}, \end{equation*} where \begin{equation*} m_N = \inf_{x \in [\alpha_N, \beta_N]} \frac{Z^{N,n_2}_{x,x_2}}{Z^{N,n_1}_{x,x_1}}, \quad M_n = \sup_{x \in [\alpha_N, \beta_N]} \frac{Z^{N,n_2}_{x,x_2}}{Z^{N,n_1}_{x,x_1}}. \end{equation*} Our goal is to show that $\lim_{N\to \infty} m_N/M_N = 1$. Since the partition function is continuous with respect to endpoints, both the supremum and infimum are achieved at some points $x = x^N_+$ and $x=x_-^{N}$. Since $\lim_{N\to -\infty} x^N_{\pm}/N=v$, Theorem \ref{thm:convergence-of-partition-function-ratios} implies \begin{equation*} \lim_{N\to -\infty} m_N = \lim_{N\to -\infty} M_N = \lim_{N\to -\infty} \frac{Z^{N,n_2}_{x^N_{\pm}, x_2}}{Z^{N,n_1}_{x^N_{\pm}, x_1}}= G_v \bigl( (n_2,x_2), (n_1,x_1) \bigr)\in(0,\infty), \end{equation*} so $\lim_{N\to -\infty} m_N/M_N =1$. This completes the proof. \end{proof} \section{Proofs of lemmas from Section~\ref{sec:setting}} \label{sec:aux} \begin{proof}[Proof of Lemma~\ref{lem:invariant_spaces}] Let us check that if $W\in{\mathbb H}$, then $\Phi^{n,n+1}_\omega W\in{\mathbb H}$ for all $n$ and $\omega$. Due to~\eqref{eq:forcing-averages-to-0}, there is a number $k=k(n,\omega)>0$ such that $F_n(x)+W(x)\ge -k(|x|+1)$ for all $x\in{\mathbb R}$. Since \begin{align*} \int_{\mathbb R} g_{2\varkappa}(y-x)e^{-\frac{F_n(x)}{2\varkappa}-\frac{W(x)}{2\varkappa}}dx\le \int_{\mathbb R} g_{2\varkappa}(y-x)e^{\frac{k|x|+1}{2\varkappa}}dx<\infty, \end{align*} $\Phi^{n,n+1}_\omega W(y)$ is well-defined for all $y\in{\mathbb R}$, and \begin{align*} \liminf_{y\to+\infty}\frac{\Phi^{n,n+1}_\omega W(y)}{y} &\ge -\liminf_{y\to+\infty}\frac{2\varkappa}{y}\ln \int_{\mathbb R} g_{2\varkappa}(y-x)e^{\frac{k|x|+1}{2\varkappa}}dx\\ &= -\liminf_{y\to+\infty}\frac{2\varkappa}{y}\ln \int_{\mathbb R} g_{2\varkappa}(y-x)e^{\frac{kx+1}{2\varkappa}}dx\\ &= -\liminf_{y\to+\infty}\frac{2\varkappa}{y}\ln (e^{\frac{ky}{2\varkappa}+\frac{k^2}{4\varkappa}+\frac{1}{2\varkappa}})=-k >-\infty. \end{align*} In the second line, we used that the contribution from the negative values of $x$ is asymptotically negligible due to the fast decay of the Gaussian kernel. For the last line, we used the Gaussian moment generating function. The behavior as $y\to-\infty$ is treated similarly. The local Lipschitz property follows from the $C^1$ property that can be obtained by differentiating the integrand in the definition of $\Phi$. Iterating this, we obtain parts~\ref{it:inv-of-HH} and \ref{it:cocycle} of the lemma. The proof of part~\ref{it:inv-of-HH-vv} is similar to that of part~\ref{it:inv-of-HH}. \end{proof} \begin{proof}[Proof of Lemma~\ref{lem:monotonicity_of_x-ux}] Let $V(\cdot)$ be the Hopf--Cole transform of $w(\cdot)$. Then \begin{equation*} x-\Psi^{n_0,n_1}w(x) = x - \int_{{\mathbb R}} (x-y) \bar{\mu}_{V, x}^{m,n}(dy) = \int_{{\mathbb R}} y \bar{\mu}_{V,x}^{m,n}(dy). \end{equation*} The conclusion then follows from Lemma~\ref{lem:monotonicity-for-second-kind-polymer-measure}. \end{proof} \bibliographystyle{alpha}
060f29d31de6372ba3662b0f197eb1c4ab2d81ca
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Galaxies are born, live, and die, within dark matter halos. We have convincing evidence that the evolutionary history of galaxies and halos is correlated to a strong degree: brighter, bigger galaxies have higher clustering, indicative of being in more massive dark matter halos (see, e.g., \citealt{norberg_etal:02, zehavi_etal:05, zehavi_etal:11} for analyses at $z\sim 0$ and \citealt{coil_etal:06, zheng_etal:07, wake_etal:11, leauthaud_etal:12_shmr} at higher redshifts). Thus the growth of galaxies is related to the growth of dark matter halos. But how correlated are these two quantities? The purpose of this paper is to quantify this correlation by constraining the scatter in stellar mass at fixed halo mass for galaxies in the Baryon Oscillation Spectroscopic Survey (BOSS; \citealt{dawson_etal:13}). We will use two-point clustering as our probe of this scatter, $\sigma_{{\rm log}M}$. For massive galaxies, clustering is an especially sensitive diagnostic of the scatter because it directly impacts their large-scale bias (e.g., \citealt{reddick_etal:13}); more scatter means a sample of galaxies will contain a more significant sample of low-mass halos that will lower the overall clustering amplitude. BOSS galaxies specifically represent an excellent sample of highly biased objecs, with previous small-scale measurements yielding clustering amplitudes roughly four times higher than that of dark matter (\citealt{white_etal:11, guo_etal:13, saito_etal:16}). However, unlike magnitude and color, galaxy stellar mass is not an observable. Different methods for deriving $M_{\ast}$ produce different results. The scatter we constrain through clustering is the quadrature sum of the intrinsic scatter of stellar mass at fixed halos, $\sigma_{\rm int}$ and and measurement error $\sigma_{\rm err}$. Different methods of deriving stellar mass will have different $\sigma_{\rm err}$ but have the same intrinsic scatter since they are all estimates of the same physical quantity. Thus, we can also use clustering to determine which method of obtaining stellar mass has the least variance. In this context, systematic offsets between codes or choices within codes, such as switching stellar initial mass functions, are immaterial. What we care about here is the rank-ordering of galaxies from most massive to least massive. Using clustering, we cannot probe systematic offsets between stellar mass estimates. In addition to comparing different methods for determining stellar mass, we compare stellar mass to other physical properties of the galaxy. Specifically, \cite{wake_etal:12} used clustering to claim that stellar velocity dispersion, $\sigma_{\rm vel}$, correlates better with halo mass than $M_{\ast}$. We will compare the clustering of galaxies ranked by 12 different variations of thee different stellar mass codes, all of which are available in the public SDSS data releases, as well as two estimates of $\sigma_{\rm vel}$ and multiple galaxy luminosities. For converting redshift to distance, as well as calculating $M_{\ast}$, we assume a flat $\Lambda$CDM\ cosmology with $\Omega_m=0.3$ and $h=0.7$. \section{Data} We use results from Data Release Ten of the Sloan Digital Sky Survey (\citealt{dr10}). The spectroscopic footprint covers 6895 deg$^2$ combined in the North Galactic Cap and South Galactic Cap regions, roughly 70\% of the full BOSS footprint. \subsection{The CMASS Sample} The CMASS target sample is the workhorse of the BOSS large scale structure analysis. Further details can be found in \cite{dawson_etal:13}. These color cuts are meant to isolate massive galaxies at $z\gtrsim 0.4$, but are somewhat more inclusive of blue galaxies than the traditional luminous red galaxy (LRG) sample from SDSS-II (\citealt{eisenstein_etal:01}). To limit the effects of the color cuts and flux limit of the survey, our fiducial results are restricted to the range $z=[0.45, 0.60]$, which surrounds the peak in the redshift distribution around $4\times 10^{-4}$ ($h^{-1}\,$Mpc)$^3$. Figure \ref{color_mass} shows the mass-vs-$g-i$ color distribution for CMASS galaxies. We will describe the mass estimate used in this figure (PCA) in \S \ref{s.mass_estimates}. We use $g-i$ because it more naturally separates blue and red galaxies at these masses and redshifts than other colors. \cite{masters_etal:11} demonstrates that a simple $g-i>2.35$ color cut best separates early type and late type CMASS galaxies. At these redshifts, $g-i=2.35$ is similar to $u-r=2.22$ at $z=0$, used by \cite{strateva_etal:01} to describe bimodality in the local universe. Unlike samples that probe galaxies near the knee in the stellar mass function (e.g., \citealt{blanton_etal:03cmd}), the CMASS sample is not bimodal in its colors. Although the color cuts are specifically designed to be more efficient at selecting passive objects than actively star-forming objects, the dominant factor in the lack of bimodality is simply that the CMASS selection is targeting the very massive end of the galaxy distribution, with a median stellar mass of $M_{\ast}\sim 10^{11.5}\,$M$_\odot$. At these masses, even the SDSS Main sample exhibits no bimodality, but rather has a tail to bluer colors comprised of the few star-forming objects at these mass scales. The CMASS selection is not devoid of any star-forming objects; \cite{chen_etal:12} find roughly $3\%$ of CMASS objects at the median stellar mass have formed 10\% of their stellar mass in the last Gyr. \subsection{The SPARSE Sample} The CMASS\_SPARSE sample (hereafter SPARSE for brevity) targets the same luminosity range but with a slightly expanded color range. The density of target galaxies is sensitive to the exact value of the color-magnitude intercept, thus the SPARSE sample is meant to expand the color range of the CMASS in order to test any biases in the selection and to probe to lower stellar masses. The median redshift of the SPARSE sample is notably lower than CMASS due to the wider color cuts. Although the color shift is only 0.28 magnitudes, there are nearly as many objects that pass the SPARSE selection (and are not included in the CMASS sample) as in the CMASS sample itself. The name SPARSE is derived from the fact that these galaxies are randomly subsampled by a factor of 5 (to 5 targets/deg$^2$) in order to minimize the number of fibers allocated to this sample (see \citealt{dawson_etal:13} for more details on the target class as well). Figure \ref{color_mass} shows the color-mass distribution of these targets in comparison to the CMASS to sample. After trimming the SPARSE catalog to objects within $z=[0.45,0.60]$, the stellar masses probed by the SPARSE selection shifts the median stellar mass lower by $\sim 0.2$ dex, and the mode of the $g-i$ color distribution also shifts blueward by $\sim 0.2$ magnitudes. But, as with CMASS, the SPARSE sample is dominated by passive red galaxies and does not exhibit any bimodality, although the blue tail extends lower in $g-i$ than the CMASS sample. The definition of the SPARSE sample has fluctuated somewhat over the course of the survey. In the first few months of observations, the color range was somewhat broader. We exclude these areas from consideration in the statistics, removing $\sim 140$ deg$^2$ of the overall footprint (delineated ``chunk2'' in the BOSS order of observations; see \citealt{dawson_etal:13}). \subsection{The WISE\_COMPLETE Sample} Relaxing the color cuts even further than the SPARSE sample would dramatically reduce the efficiency of finding $z\gtrsim 0.4$ galaxies. To efficiently locate galaxies in our desired redshift range, but outside the BOSS color cuts, ancillary data must be brought in. To this end, a series of 26 plates covering 59.8 deg$^2$ were dedicated to observing three sets of ancillary targets, one of which was the WISE\_COMPLETE target set. These targets incorporate data from the Wide-field Infrared Survey Explorer (WISE; \citealt{wise}). These targets have the same magnitude range as CMASS and SPARSE, but employ a single color cut, \begin{equation} \label{e.wise} r-W1>4.165, \end{equation} \noindent to select galaxies in our desired stellar mass and redshift range that are outside the standard BOSS color cuts\footnote{In the input catalog, we removed any galaxy that also passed the CMASS cuts. In post-processing, we removed any target that also passed the SPARSE selection. This latter cut encompassed 11\% of the input WISE\ catalog.}. Here, $W1$ is the WISE 3.4 micron band. Using WISE near-IR photometry is promising because the IR bands are on the far side of the peak of stellar emission, so an optical-IR color is sensitive to redshift. We will refer to the WISE\_COMPLETE sample as WISE\ for brevity. There are 7368 objects within the WISE\ target catalog that received fibers and recovered accurate redshifts, 2200 of which are within the redshift range of interest. The $g-i$ color distribution is broad and flat, with a median stellar mass of $M_{\ast}=10^{11.06} $M$_\odot$. \begin{figure} \epsscale{1.2} \plotone{color_mass.ps} \vspace{-0.5cm} \caption{ \label{color_mass} The $g-i$ color-mass distributions of the three different target classes utilized in this paper. The color contours represent the CMASS galaxy sample, with color representing density of points in each cell. The black curves represent the density contours of SPARSE targets. The WISE targets are sparse enough that they can be shown individually, represented by the filled circles. All results are shown in the redshift range of $z=[0.45, 0.60]$.} \end{figure} \begin{figure} \epsscale{1.2} \plotone{smf_compare.ps} \vspace{-0.5cm} \caption{ \label{smf_compare} Comparison of the space densities of each individual BOSS targets class. All measurements implement at $V_{\rm max}$ correction, which has minimal effect at $\logM_{\ast}\ge 11.4$ but can significantly enhance the measured abundance at low masses. Red circles represent the CMASS targets. Green squares represent the SPARSE sample. Yellow triangles represent the WISE\ data. Each of these measurements peaks at a lower stellar mass, demonstrating the effect of reaching further outside the typical LRG color selection.} \end{figure} \subsection{Physical properties of the BOSS galaxies} \label{s.mass_estimates} As discussed in the introduction, we are interested in finding the physical property that correlates the closest with halo mass via the clustering of the sample. We use stellar mass, galaxy luminosity, and galaxy velocity dispersion. Our main focus is stellar mass, but the question of whether stellar mass correlates any better with halo mass than does luminosity is still open. Table \ref{t.props} shows all 16 variations of galaxy properties we use to rank-order the CMASS galaxy sample. The first 12 are stellar masses that are included in the standard BOSS pipeline. There are three different codes for deriving stellar masses from BOSS data, with each code producing variants based upon parameter choices (i.e., stellar initial mass function, stellar population synthesis code, etc.). The Portsmouth masses (\citealt{maraston_etal:13}) use two template star formation histories, one to describe a passively evolving, old stellar population meant to mimic an LRG-type galaxy, and the other is an actively star-forming template. These two templates are combined to make one catalog: each galaxy is assigned a template based on its color, with a redshift-dependent break point of $g-i=3.25 +1.67(z-0.53)$. Galaxies redder than this color are assigned the passive template, which bluer galaxies are assigned the star-forming template. Thus, although we list 14 different stellar mass definitions in Table \ref{t.props}, there are only 12 different stellar mass catalogs, two of which are comprised of the Portsmouth masses. The Granada masses are an implementation of the Flexible Stellar Population Synthesis (FSPS) mass modeling described in \cite{conroy_etal:09} . The PCA masses (`principal component analysis') are described in \cite{chen_etal:12}. In contrast to the first two methods, the PCA approach uses the entire galaxy spectrum, decomposing it into principal components which are then fit to a library of varying galaxy templates to obtain the physical properties of the galaxies, including $M/L$ ratio, velocity dispersion, and star formation rate. We use two different estimates of the stellar velocity dispersion, one that is produced by the PCA analysis, the other produced by the Portsmouth analysis code (see details for the latter in \citealt{thomas_etal:13}).. Lastly, we use absolute $i$-band magnitude as our baselines rank-order by which to compare all the methods. We also include a sample that has been $k$-corrected to $z=0.52$ using the {\tt kcorrect} code of \cite{blanton_roweis:07}. \section{Measurements} \subsection{The Completeness and Abundance of the BOSS Galaxy Samples} Figure \ref{smf_compare} shows the stellar mass function, $\Phi(\log M_{\ast})$ for each of the three samples discussed above. All results use the PCA stellar masses. We estimate $\Phi(\log M_{\ast})$ using the $1/V_{\rm max}$ method, while noting that the $V_{\rm max}$ weighting really only effects the results at stellar masses below the peak of each distribution. The main source of incompleteness in these measurements is the color selection imposed on each sample. The CMASS sample dominates the statistics at high masses, but the combination of the SPARSE and WISE\ samples becomes equivalent to the CMASS abundance at $\logM_{\ast}\sim 11.4$. We use the abundance of SPARSE and WISE\ galaxies to quantify the completeness of the BOSS target selection. Figure \ref{completeness} shows the ratio of the CMASS SMF to that of the total SMF (i.e., CMASS+SPARSE+WISE) as a function of $\logM_{\ast}$. The CMASS sample is roughly 50\% complete in stellar mass at $\logM_{\ast}=11.4$. These results are in reasonable agreement with those of \cite{leauthaud_etal:16}, which uses the extra near-infrared imaging data available in Stripe 82 to estimate completeness. Although the combination of WISE and SPARSE data get us most of the way to a fully complete sample, the total sample used in this paper is still missing $\sim 10\%$ of galaxies at $\log M=11.4$. The stellar masses of \cite{leauthaud_etal:16} are different than the PCA masses used here due to the inclusion of infrared imaging data. From the comparison between the near-IR masses and the PCA \cite{bundy_etal:15}, there is a shift of 0.15 dex between the two mass definitions, which we use to shift the completeness \cite{leauthaud_etal:16} completeness curves onto the PCA mass scale. At $\logM_{\ast} = 11.4$, the combination of CMASS and SPARSE is roughly 75\% complete. Henceforth, we will limit all analyses to be above this mass scale. The limited size of the WISE\ sample prevents us from making robust clustering measurements for WISE\ galaxies, thus we make the assumption that the bias of WISE\ galaxies and SPARSE galaxies is the same when combining the clustering results of the samples. This does not significantly affect our results; the constraining power on $\sigma_{\logM_{\ast}}$ derive from the region of the SMF where CMASS dominates the statistics. Figure \ref{smf_dr10_fit} shows the completeness-corrected SMF of massive galaxies down to the limit of $\logM_{\ast}=11.4$. We fit these data using a fitting function of the form: \begin{equation} \Phi(M_{\ast}) = \Phi^\ast\left(\frac{M_{\ast}}{M_1}\right)^{(1+\alpha)} \left[1+\left(\frac{M_{\ast}}{M_1}\right)^\beta\right]^{(\gamma-\alpha)/\beta}\ln(10) \end{equation} \noindent with best-fit parameter values of $\Phi^\ast=2.43\times10^{-2}$, $\log M_1=11.47$, $\alpha=0.728$, $\beta=1.173$, and $\gamma=-6.01$. This fitting function will be used to map galaxy mass onto halo mass using the abundance matching method in \S \ref{s.slogm}. \begin{figure} \epsscale{1.2} \plotone{completeness.ps} \vspace{-0.5cm} \caption{ \label{completeness} Completeness of the BOSS target samples, relative to the total abundance of all three target classes. At $\logM_{\ast}=11.4$, indicated by the vertical dash line, the CMASS sample is only 50\% complete. Adding the SPARSE sample brings the completeness at this mass scale up to 75\%. The CMASS sample by itself is >95\% complete at $\logM_{\ast}>11.7$. At lower masses, the total abundance is itself incomplete. We thus make a conservative completeness limit of $\logM_{\ast}=11.4$ for the full BOSS sample, at which scale both SPARSE and WISE\ abundances have steep mass functions. } \end{figure} \begin{figure} \epsscale{1.2} \plotone{smf_dr10_fit.ps} \vspace{-0.5cm} \caption{ \label{smf_dr10_fit} {\it Upper Panel:} The total BOSS stellar mass function at $z=[0.45,0.60]$, down to our completeness limit of $\logM_{\ast}=11.4$. The best-fit parameters of equation (x) are shown with the solid curve. {\it Lower panel:} Residuals of the fit. The $y$-axis is $(\Phi_{BOSS}-\Phi_{fit})/\Phi_{fit}$. The $\chi^2$ for this fit is 7.0 for $(10-5)$ degrees of freedom.} \end{figure} \subsection{Measuring Clustering} Our statistical measure of the clustering of galaxies within BOSS is the projected two-point correlation function $$w_p(r_p)$$. This statistic integrates the two-dimensional redshift-space correlation function, $\xi(r_p,r_\pi)$, along the line-of-sight direction $r_\pi$. $$w_p(r_p)$$ is defined as \begin{equation} $w_p(r_p)$ = 2\times \int_0^{\pi_{\rm max}} \xi(r_p,r_\pi) dr_\pi, \end{equation} \noindent where $r_p$ is the projected separation between galaxy pairs and $\pi_{\rm max}$ is the maximal line-of-sight separation, which here we make 80 $h^{-1}\,$Mpc. To estimate $\xi(r_p,r_\pi)$ we use the Lansy-Szalay estimator (\citealt{landy_szalay:93}) with $10^7$ randoms. When $\pi_{\rm max}=\infty$, the projected correlation function is identical whether the argument in the integrand is the real- or redshift-space correlation function. Having a finite $\pi_{\rm max}$ introduces some deviation from a real-space-only calculation (see, e.g., \citealt{vdb_etal:13}). The projected clustering of galaxies is our preferred statistic for two reasons. First, peculiar motions are (mostly) removed in the line-of-sight integration. At large scales there is still a residual effect, but we will model this out when fitting for bias. Second, in the third paper in this series, we will present halo occupation fits of the clustering to determine the stellar-to-halo-mass ratio of BOSS galaxies. Removing the peculiar motions makes such modeling less cosmologically dependent and easier to implement analytically. To estimate the errors on the correlation function, we jackknife the DR10 footprint into 100 roughly equal-size angular regions, removing one subsample at a time and calculating the covariance matrix as \begin{equation} C_{ij} = \frac{N-1}{N}\sum_{n=1}^N \left(w_i - \bar{w}_i\right) \left(w_j - \bar{w}_j\right), \end{equation} \noindent where $N$ is the total number of jackknife subsamples (100), and $i$ and $j$ represent $r_p$ bins, and $\bar{w}$ represents the mean correlation function in each bin. To correct for fiber collisions, we use the angular-upweighting method as described in \cite{white_etal:11}. On a given plate, fibers can only be positioned within 62 arcsec of one another. At $z=0.5$, this translates to a projected separation of about 400 kpc/$h$. Roughly 40\% of the area within the survey is covered by multiple plates. These areas have nearly unit completeness, at least in terms of fiber assignment. Galaxy pairs within the collision radius are upweighted by the ratio of the angular correlation function of all CMASS targets to those for which fibers were assigned at the angle of the pair in question. Inside 62'', this ratio\footnote{Which, incidentally, is the ratio of total area of the survey to that covered by more than one plate.} is nearly a constant value of 2.57. In practice, we only use this method to correct for one $r_p$ bin inside the collision radius, and for no bins that effect our bias calculation. We measure the autocorrelation of the CMASS sample, but the SPARSE sample, by definition, has too low a density to make a robust measurement of the clustering. We thus cross-correlate the SPARSE and CMASS samples to get the relative bias between these two samples and then divide by the bias of the overall CMASS sample, which is $1.99\pm 0.04$ for the cosmology chosen. \subsection{Measuring Bias} \label{s.measuring_bias} We use the projected separation range $5< r_p <35$ $h^{-1}\,$Mpc\ to determine the bias relative to the clustering of matter. Although clustering is more linear at larger scales, errors on clustering rise monotonically with scale. Additionally, as $r_p$ increases, the contribution of peculiar velocities to $$w_p(r_p)$$ also monotonically increases. Although we model out the peculiar motions, the optimal approach is to extend the $r_p$ range out as far into the linear regime as is possible, while keeping the effect of a finite $\pi_{\rm max}$ to under 25\%. To recover linear bias from our clustering measurements, we need to model both peculiar velocities and non-linear clustering. Rather than try to subtract these effects out of our measurements, we add them into the model for the matter clustering. First, we model real-space galaxy clustering as \begin{equation} \xi_{\rm gal} (r) = b_{\rm gal}^2\xi_m(r)\zeta(r), \end{equation} \noindent where $\xi_m(r)$ is the non-linear matter correlation function from \cite{smith_etal:03}, $b_{\rm gal}$ is the bias parameter for the galaxy sample, and $\zeta(r)$ is the scale-dependent halo bias function from equation (B7) in \cite{tinker_etal:05}. This scale-dependence of halo bias, relative to the large-scale value, asymptotes to unity at $r\gtrsim 20$ $h^{-1}\,$Mpc, with a maximum deviation of $\sim 10\%$ at $r\sim 2$ $h^{-1}\,$Mpc\ (see Figure 12 in \citealt{white_etal:11} for a halo occupation fit of the scale-dependent bias of BOSS CMASS galaxies). At smaller scales, the scale-dependence of galaxy bias becomes dependent on the details of halo occupation and it is only partially correlated with bias at large scales. Second, we incorporate the effects of peculiar velocities using the linear approximation of \cite{kaiser:87}. Although peculiar velocities are poorly described by linear theory at $r\lesssim 20$ $h^{-1}\,$Mpc, at $\pi_{\rm max}$ of 80 $h^{-1}\,$Mpc\ eliminates the contribution of redshift space distortions at scales where linear theory breaks down. The configuration-space implementation of the Kaiser effect is described in detail in Appendix A of \cite{hawkins_etal:03}. In practice we fit for $b_{\rm gal}$ by $\chi^2$ minimization using the full covariance matrix. \subsection{Numerical Simulations} \label{s.nbody} Our theoretical models are created using the publicly available halo catalog from the MultiDark simulation (\citealt{nuza_etal:13}). The cosmology for this simulation is $\Omega_m=0.27$ and $\sigma_8=0.82$. The halo catalog incorporates halos and subhalos using the {\small ROCKSTAR} algorithm of \cite{rockstar}. We populate these halos with galaxies using the sub-halo abundance-matching model (a.k.a. SHAM; see \citealt{reddick_etal:13} and references therein). In its simplest form, abundance matching provides a unique mapping of halo mass onto galaxy mass (or galaxy luminosity) by assuming a one-to-one relation between the two without any scatter. There are multiple ways to incorporate scatter (\citealt{behroozi_etal:10, trujillo_gomez_etal:11}). We use the method in \cite{wetzel_white:10}, in which the stellar mass function is first deconvolve of lognormal scatter of width $\sigma_{{\rm log}M}$. This allows us to employ the simple abundance matching method using the monotonic relation between $M_{\ast}$ and $M_h$. Then the $\logM_{\ast}$ of each galaxy is shifted by a random Gaussian deviate of the same width. Once halos are populated with galaxies, the large-scale bias of the model is calculated by binning the mock galaxies by $\logM_{\ast}$ in the same manner as the data, and summing over each halo in the stellar mass bin, weighting each galaxy by the bias value $b_h(M_{\rm halo})$ given by the \cite{tinker_etal:10_bias} bias function. For sub-halos, we use the mass of the host halo (the halo that the subhalo is located within) to calculate the bias. More explicitly, the bias of a sample galaxies is given by summing over each galaxy, $i$, in the sample: \begin{equation} b_{\rm gal} = \frac{1}{N_{\rm gal}}\sum_i b_h(M_{\rm host}^{(i)}), \end{equation} \noindent where $N_{\rm gal}$ is the total number of galaxies in the sample. This approach allows us to calculate the large-scale bias rapidly, without explicitly measuring the clustering of the mock galaxies. This is also less noisy than measuring the clustering, an important feature in this method as the BOSS data is actually a larger volume than the MultiDark simulation. \begin{deluxetable*}{lllll} \tablecolumns{5} \tablewidth{30pc} \tablecaption{Quantities to rank-order the CMASS galaxies\label{t.props}} \tablehead{\colhead{Number} &\colhead{Property} &\colhead{code} & \colhead{IMF} &\colhead{comments}} \startdata 1 & Stellar mass & PCA & Kroupa & BC03 SPS \\ 2 & Stellar mass & PCA & Kroupa & M11 SPS \\ 3 & Stellar mass & Granada & Salpeter & wide formation times, no dust \\ 4 & Stellar mass & Granada & Salpeter & wide formation times, dust \\ 5 & Stellar mass & Granada & Salpeter & early formation times, no dust \\ 6 & Stellar mass & Granada & Salpeter & early formation times, dust \\ 7 & Stellar mass & Granada & Kroupa & wide formation times, no dust \\ 8 & Stellar mass & Granada & Kroupa & wide formation times, dust \\ 9 & Stellar mass & Granada & Kroupa & early formation times, no dust \\ 10 & Stellar mass & Granada & Kroupa & early formation times, dust \\ 11 & Stellar mass & Portsmouth & Kroupa & SF template \\ 12 & Stellar mass & Portsmouth & Salpeter & SF template \\ 13 & Stellar mass & Portsmouth & Kroupa & passive template \\ 14 & Stellar mass & Portsmouth & Salpeter & passive template \\ 15 & Luminosity & --- & --- & absolute, $i$-band \\ 16 & Luminosity & --- & --- & absolute, $i$-band, k-corrected \\ 17 & Velocity dispersion & PCA & --- & --- \\ 18 & Velocity dispersion & Portsmouth & --- & --- \\ \enddata \end{deluxetable*} \begin{figure*} \plotone{wp2.ps} \vspace{-8cm} \caption{ \label{wp2} {\it Top Left}: Projected clustering relative to clustering of BOSS CMASS galaxies ordered by $i$-band absolute magnitude. The shaded region indicates the typical error of the clustering measurement. The red symbols indicate the clustering when BOSS galaxies are rank-ordered by the PCA masses. The yellow triangles indicate clustering when ranked by $i$-band absolute magnitude k-corrected to $z=0.52$. The blue dotted curves are the clustering when ranked by Granada Granada masses. The green dashed curves indicate the clustering ranked by Portsmouth masses. The orange symbols indicate clustering when galaxies are ranked by velocity dispersion; orange circles are the PCA estimate of velocity dispersion, while orange squares are the Portsmouth dispersion estimate. All samples are the top 100,000 out of 350,000 galaxies in this redshift slice. {\it Top Right}: Projected clustering for a subset of the samples in the left panel (to avoid confusion). In this panel, clustering for the top 25,000 galaxies is measured for each rank-order. Red squares show the PCA masses with BC03 SPS models. The blue dotted curve is the fiducial Granada mass: early formation times, Kroupa IMF, dust modeling. The green dash line is the fiducial Portsmouth mass, which uses Kroupa IMF. The yellow triangles represent k-corrected $i$-magnitude. The differences between PCA and other ranking methods is large enough to be seen without taking ratios. } \end{figure*} \begin{figure*} \plotone{wp_3panel.ps} \vspace{-6.5cm} \caption{ \label{wp_3panel} {\it Left Panel}: Projected clustering of CMASS galaxies binned by stellar mass above the completeness limit of the CMASS selection algorithm. The solid black curve shows the nonlinear clustering of dark matter. {\it Middle Panel}: Clustering of CMASS galaxies at and below the completeness limit of the sample. {\it Right Panel}: Cross-correlation function of the SPARSE sample, binned by stellar mass, with the full CMASS sample. In each panel, the numbers on the right hand side indicate the center of the bin in $\logM_{\ast}$.} \end{figure*} \section{Results} \subsection{What galaxy property correlates most with halo mass?} Figure \ref{wp2} compares the clustering of all 18 properties listed in Tabel \ref{t.props} at fixed number density. For each quantity, we rank-order the sample form highest to lowest value. In this method, cutting the samples at a given rank means that each sample has the same number density. In the left panel, we show $$w_p(r_p)$$ relative to the $i$-band $$w_p(r_p)$$ for the top 100,000 galaxies for each property (about $\sim 1/3$ the total sample). In the right panel, we show the results for the top 25,000 galaxies in each property, although in the right panel we only show a subset of the properties considered. These two sample sizes correspond roughly to number densities of $\sim 1.2\times10^{-4}$ and $0.3\times 10^{-4}$ $($$h^{-1}\,$Mpc$)^{-3}$, respectively. The left panel demonstrates that the observed scatter between halo mass and luminosity is larger than all other galaxy properties considered, regardless of the stellar mass code employed. The cabal of photometrically-based stellar mass estimates yield a bias roughly 3\% higher than $i$-band magnitude. Because all the different variations of these two codes are roughly consistent with one another, we do not attempt to differentiate them in the plot. The clustering for these samples increases sharply at $r_p<1$ $h^{-1}\,$Mpc\ relative to $i$-band clustering. Clustering at these scales is reflective of the number of satellite galaxies per halo (often referred to as the `one-halo' term). At fixed luminosity, red galaxies are less massive than blue galaxies. Because satellite galaxies are more often red than field galaxies of the same $M_{\ast}$, at fixed number density there will be a higher fraction of satellites when ranked by $M_{\ast}$ rather than by luminosity. This argument is clearly applicable for a complete sample like the SDSS Main galaxy sample. For the CMASS target selection, which preferentially selects red galaxies, the impact of this argument is less clear. However, the increase in the clustering in one-halo term argues that the small fraction of blue galaxies in CMASS is coming into play when rank-ordering the same by luminosity and by stellar mass. The amplitude of clustering for the samples defined by velocity dispersion is consistent with the photometrically-derived stellar masses. For some photometric stellar mass codes, the clustering of the velocity dispersion sample is slightly higher than the photometric stellar mass clustering, but when considering the entire ensemble of photometric mass definitions it is not clear that there is any differentiation between the two properties when it comes to clustering amplitude. The spectroscopic PCA masses have a clustering amplitude $\sim 25\%$ higher than $i$-band clustering, significantly higher than both velocity dispersion and photometric stellar mass. Although we do not perform any halo occupation fits of the clustering, it is improbable to ascribe the boost in the large-scale amplitude entirely to an increase in satellite galaxies; as opposed to the photometric stellar masses, the relative $$w_p(r_p)$$ for the PCA masses decreases in the one-halo term. In the right panel, we restrict the lists to only the most highest 25,000 objects by rank. In this sample, the difference between the PCA clustering and the other samples is large enough---$\sim 60\%$---to be seen clearly on a logarithmic scale, and without taking ratios. In this panel, we show only a subset of clustering results in order to avoid crowding of the plot. We therefore conclude that $\sigma_{{\rm log}M}$, and by extension $\sigma_{\rm err}$, for the PCA masses is smaller than for photometrically-defined stellar masses, galaxy luminosity, and stellar velocity dispersion. For the remainder of this paper, all results will use the PCA masses with the BC03 SPS code. \subsection{Bias as a function of stellar mass} Figure \ref{wp_3panel} show the measured values of $$w_p(r_p)$$ for CMASS and the total CMASS sample crossed with SPARSE galaxies, binned by stellar mass. Using the technique described in \S \ref{s.measuring_bias}, we fit for $b_{\rm gal}$ for each bin in $M_{\ast}$ for CMASS and for the CMASS-SPARSE cross-correlation. Figure \ref{bias} shows the results for each sample. The SPARSE results have had the bias of the overall CMASS sample divided out. At $\logM_{\ast}\ge 11.4$, the CMASS bias rises monotonically with stellar mass. Not coincidentally, this is the mass range where the CMASS sample is most complete in stellar mass. In terms of space density, the CMASS sample peaks at $\logM_{\ast}=11.4$, and rapidly decreases at smaller masses, mainly due to the color cuts. At $\logM_{\ast}<11.4$, the CMASS bias rises back up again. This rise at low masses can be explained if the target selection cuts preferentially select satellite galaxies at lower masses. Previous studies of color-dependent clustering have shown this U-shaped behavior is red galaxy clustering (see, e.g., \citealt{swanson_etal:08_bias, ross_etal:11}), a trend that is driven by the increasing fraction of satellite galaxies in the red subsample as luminosity or stellar mass decreases. In contrast to those results, the minimum bias shown in Figure \ref{bias} is at a much higher mass relative to the knee in the stellar mass function, but it is possible that the effect is amplified by the BOSS CMASS color cuts. The clustering of the SPARSE sample, however, is nearly independent of stellar mass. The WISE\ sample is not large enough to afford spatial clustering analysis, so we make the approximation that the WISE\ clustering is the same as the SPARSE and weight the SPARSE clustering results accordingly. The shaded band represents the total bias of galaxies in the combined sample weighted by the relative number if the CMASS vs. the (WISE+SPARSE) samples. \subsection{Scatter of stellar mass at fixed halo mass} \label{s.slogm} Figure \ref{bias_combined} compares our combined bias results to predictions from the numerical simulation described in \S \ref{s.nbody}. As described previously, each curve represents a model that matches the same stellar mass function, but has different amounts of scatter between halo mass and stellar mass. Four different curves representing different $\sigma_{{\rm log}M}$ values are shown for comparison. A model with zero scatter would predict clustering too high relative to the results. A model with scatter at $\sigma_{{\rm log}M}=0.26$ is ruled out because the clustering amplitudes are too low. The best-fit value of $\sigma_{{\rm log}M}$ is 0.18, with a 68\% confidence interval of $[0.16,0.19]$. The SDSS pipeline reports a mean error of 0.16 dex for the PCA masses. This includes both systematic and random errors. To estimate $\sigma_{\rm err}$---the statistical errors alone---we use repeated spectra of CMASS galaxies that occur on regions of the footprint covered by multiple tiles. The rms difference in masses between the repeat spectra is 0.11 dex, implying that the intrinsic scatter of stellar mass at fixed $M_{\rm halo}$ is 0.16 dex. \subsection{The Stellar-to-Halo Mass Relation} Figure \ref{shmr_grid} shows the stellar-to-halo mass relation (SHMR) using the best-fit $\sigma_{{\rm log}M}$ value of 0.18. The solid curve represents the mean stellar mass at fixed halo mass, $\langle M_{\ast} | M_{\rm halo}\rangle$. Due to the steepness of the halo mass function, the reverse relation, $\langle M_{\rm halo} | M_{\ast}\rangle$, is quite different. The filled circles show the $\langle M_{\rm halo} | M_{\ast}\rangle$ for the bins analyzed in this paper. The errors show the inner 68\% of the distribution of $\logM_{\rm halo}$ in each bin, demonstrating the significant overlap in the halo distribution analyzed in each bin. The left-hand side of Figure \ref{shmr2} shows the sensitivity of the SHMR to the assumed value of $\sigma_{{\rm log}M}$. At $\sigma_{{\rm log}M}=0.13$, the SHMR is a steeply rising function of $M_{\rm halo}$. At $\sigma_{{\rm log}M}=0.20$, $M_{\ast}$ becomes nearly independent of halo mass at $\logM_{\rm halo}\gtrsim14$. The reverse relations, $\langle M_{\rm halo} | M_{\ast}\rangle$ shows the opposite trend; as $\sigma_{{\rm log}M}$ gets larger, the mean halo mass at fixed $M_{\ast}$ decreases, yielding the trend of the theoretical models in Figure \ref{bias_combined} and producing the tight constraints on $\sigma_{{\rm log}M}$. The right-hand side of Figure \ref{shmr2} compares our best-fit relation to a sample of other measurements. Those based purely on abundance matching---i.e., no other data other than the stellar mass function is fit, (\citealt{behroozi_etal:13} and \citealt{moster_etal:13} in the Figure)---lie significantly below our relation. This comparison is complicated by different stellar mass estimates and significantly smaller data samples at $z=1$, but it is clear the estimates that probe this relation directly using galaxy groups and clusters are in better agreement with the BOSS relation (\citealt{lin_mohr:04, hansen_etal:09, yang_etal:09_groups3, yang_etal:12}). Last, our results are in good agreement with the SHMR for BOSS galaxies obtained by \cite{rodriguez_torres_etal:16}. This relation is based on the Portsmouth stellar masses (line 13 in Table 1) and the abundance matching is based on peak circular velocity, with a scatter of $0.31\times V_{\rm peak}$ obtained by matching the redshift space monopole of the galaxy correlation function. \begin{figure} \epsscale{1.2} \plotone{bias.ps} \vspace{-0.5cm} \caption{ \label{bias} Bias as a function of stellar mass for the CMASS sample (circles), the SPARSE sample (squares) and the combined sample (gray shaded region). } \end{figure} \begin{figure} \epsscale{1.2} \plotone{bias_combined.ps} \vspace{-0.5cm} \caption{ \label{bias_combined} The combined CMASS+SPARSE bias values compared to models derived by abundance matching dark matter halos and subhalos in the MultiDark simulation to the stellar mass function of BOSS. Different curves indicate different values of scatter (in $\logM_{\ast}$) in stellar mass at fixed halo mass. } \end{figure} \begin{figure} \epsscale{1.2} \plotone{shmr_grid.ps} \vspace{-1cm} \caption{ \label{shmr_grid} The stellar-to-halo mass relation (SHMR) for BOSS CMASS galaxies using the PCA stellar mass estimates. The solid green curve shows the mean $M_{\ast}$ in bins of $M_{\rm halo}$, with the dashed curves indicating the 0.18 dex in scatter of the best-fit relation. The colored circles show the mean $M_{\rm halo}$ in the observed bins in $M_{\ast}$ used in this paper. Error bars indicate the inner 68\% of the distribution of $\logM_{\rm halo}$ in each bin. The gray shaded region indicates the number of halos, scaled as $\log N_h$, at each point in this 2D parameter space. } \end{figure} \begin{figure*} \plotone{shmr2.ps} \vspace{-7.5cm} \caption{ \label{shmr2} {\it Left-hand side:} The sensitivity of the SHMR to the assumed value of $\sigma_{{\rm log}M}$. The solid curves show the SHMR for each value of scatter indicated in the key. The dashed curves show the mean $M_{\rm halo}$ in bins of $M_{\ast}$. This figure explains why the bias in Figure \ref{bias_combined}, $b(M_{\ast})$, decreases as the scatter increases. {\it Right-hand side:} Comparison of the SHMR derived here to other measurements in the literature. There are few measurements at $z=0.5$, show we show values at $z=0$ and $z=1$ from the same works, with the expectation that the $z=0.5$ value should lie somewhere in between. Estimates of the SHMR from abundance matching, such as \cite{behroozi_etal:13} and \cite{moster_etal:13}, lie significantly below the relation here. Values derived from cluster samples appear to be in much better agreement with BOSS. Lastly, we are in good agreement with the results of \cite{rodriguez_torres_etal:16}, who also analyze the CMASS sample but using a different stellar mass estimate and abundance matching technique. } \end{figure*} \section{Discussion} In this paper we have demonstrated a novel method of discriminating between disparate methods of estimating the stellar masses of galaxies. We caution that the clustering results represent only half the story when it comes to appraising different methods because clustering is only sensitive to the rank-ordering of a set of galaxies, from most massive to least massive, which in turn provides information on the scatter induced in the estimation of $M_{\ast}$. However, the method has nothing to say about absolute offsets between different methods. Using this technique, that stellar masses derived from the PCA code of \cite{chen_etal:12} correlate stronger with halo mass than $i$-band magnitude, velocity dispersion, and other estimates of stellar mass provided in the BOSS pipeline. These other stellar masses are based on photometric data only, while the PCA masses are based on analysis of the spectral information. The PCA method has the limitation that the spectra only contain information about the galaxy from within the diameter of the fiber, which for BOSS is $2\arcsec$. This may lead to aperture bias for the derived quantities, but any bias goes in the direction of strengthening the correlation with halo mass, which indicates something fundamental about the quantities being estimated. But given the average radius of the typical BOSS galaxy of $1.2\arcsec$ (\citealt{masters_etal:10}), aperture bias might be non-negligible but it is unlikely to be the dominant source of the differences between the methods. \cite{bundy_etal:15} present a detailed comparison of the Porstmouth and PCA masses with masses obtained using extra infrared imaging available in the SDSS southern equatorial stripe, `Stripe 82' in the collaboration parlance. The dispersion between the Portsmouth masses\footnote{The Portsmouth mases used here, which are a combination of the two templates based on galaxy color, are referred to in \cite{bundy_etal:15} `Porstmouth Best'.} and the near-IR masses is 0.29 dex, while the dispersion between near-IR and PCA masses 0.20. The tighter correlation with near-IR masses supports the results here that the PCA masses have a smaller intrinsic scatter relative to other photometric-based methods. As discussed above, however, there is a bias between the near-IR and PCA of 0.15 dex as well, as well as an offset of the PCA masses with respect to the Porstmouth values (\citealt{chen_etal:12}). Addtionally, the lower amplitude of the PCA $$w_p(r_p)$$ at $r_p<1$ Mpc$/h$ is indicative of fewer satellite galaxies in the PCA sample relative to the other stellar mass samples (cf. Figure \ref{wp2}). This does not mean that satellites are `missing' from the PCA catalog, but rather that they are assigned lower masses than in other catalogs. Satellites should be redder than the overall population, implying that the PCA method finds higher stellar masses for bluer galaxies relative to other methods. \cite{tinker_etal:12_cosmos} found that X-ray groups with bluer central galaxies have higher clustering at fixed halo mass, at the same redshift and stellar mass as the BOSS sample. \cite{tinker_etal:12_cosmos} concluded that this was assembly bias in massive galaxies, with the caveat that the sample was statistically limited. It is possible that the higher clustering amplitude of the PCA masses is partly due to an assembly bias effect imparted by the relative ranking of bluer and redder galaxies within the catalog, and not entirely from a minimization of intrinsic scatter. The evidence for this is circumstantial at best, but is worth further investigation. It is noteworthy that all stellar mass estimates considered here correlate with halo mass better than absolute magnitude. The CMASS sample, although it is primarily intended to target luminous red galaxies, does have a non-negligible component of star formation galaxies with blue(ish) colors that have a lower clustering amplitude at fixed $M_i$ (\citealt{guo_etal:13}). We conclude that almost any estimate of stellar mass provides a more robust rank-ordering of a heterogeneous color sample of galaxies relative to luminosity alone. In our analysis, the photometric stellar mass indicators provide the same level of scatter as velocity dispersion. All of which yield larger scatter than the PCA stellar masses. \cite{wake_etal:12} used clustering to claim that velocity dispersion had a stronger correlation with halo mass than stellar mass, a result that was challenged by \cite{li_etal:13}, who claim that the Wake et.~al.~result is contaminated by satellite galaxies, and once these satellites are removed, stellar mass has the strongest correlation. Our results suggest that the choice of stellar mass estimator can play a large role in this comparison. Both Wake et.~al.~and Li et.~al.~employ photometric mass estimates which are likely to include extra measurement scatter, although we do not test their exact methods. We make no attempt to remove satellite galaxies from our analysis, but the satellite fraction of BOSS galaxies is low ($f_{\rm sat}=0.10\pm0.02$ from \citealt{white_etal:11}), and it is clear in Figure \ref{wp2} that the PCA clustering in the one-halo term ($r_p\lesssim 1 Mpc/h$) is lower relative to the photometrically-based codes. This implies that the satellite fraction of the PCA sample is smaller than the satellite fraction of the other samples, thus not artificially enhancing the clustering of the sample. Our theoretical modeling includes subhalos in the abundance matching procedure, and this method yields the proper satellite fraction of BOSS galaxies (\citealt{nuza_etal:13}). Our constraint on $\sigma_{{\rm log}M}$ of $0.18^{+0.01}_{-0.02}$ compares favorably with other estimates from the literature. \cite{zu_mandelbaum:16} use abundance and lensing data to obtain $\sigma_{{\rm log}M}=0.20\pm0.01$ at $M_{\rm halo}=10^{13}$ $h^{-1}\,$M$_\odot$\ (linearly interpolating between their results at $10^{12}$ and $10^{14}$). \cite{more_etal:11}, using satellite kinematics, find 68\% confidence regions of $[0.14,0.21]$ for red galaxies and $[0.07,0.26]$ for blue galaxies in the SDSS Main sample. \cite{reddick_etal:13}, using the SHAM approach with multiple free parameters constrained by both clustering and comparison to an SDSS group catalog, find $\sigma_{{\rm log}M}=0.20\pm0.03$. \cite{leauthaud_etal:12_shmr} apply the SHMR approach to multiple statistics in the COSMOS field to find $\sigma_{{\rm log}M}=0.25\pm0.02$ at similar redshifts to those probed here. Our constraint has both the smallest uncertainty and the lowest value itself. The previous measurements used photometrically-derived stellar masses, which must be contributing to the measurement scatter in each study. \cite{kravtsov_etal:14} measure the scatter in brightest cluster galaxy mass for a sample of X-ray clusters, findering $\sigma_{{\rm log}M}=0.17\pm 0.02$, which is excellent agreement with our results. The notable aspect of the comparison between \cite{kravtsov_etal:14} and our results is that they are largely distinct in the halo masses probed; Kravtsov et.~al.~analyze clusters at $M_{\rm halo} \gtrsim 10^{14}$ $h^{-1}\,$M$_\odot$, which is the upper limit of the halo masses probed by out stellar mass bins. This implies that $\sigma_{{\rm log}M}$ is independent of halo mass over the range $12.7\lesssim \logM_{\rm halo} \lesssim 15.2$. This has largely been assumed, mainly because existing data could be fit with a constant scatter, and constraints on $\sigma_{{\rm log}M}$ at low halo and galaxy masses are very weak because of the lack of variation of bias with halo mass at those scales. Additionally, all of these results imply little to no redshift evolution in $\sigma_{{\rm log}M}$ for massive galaxies. This is expected, given that the maority of the galaxy population at these masses is passively evolving, although strict passive evolution does not fit the evolutino of the clustering of massive galaxies over the same timespan (\cite{zhai_etal:16}). \cite{gu_etal:16} investigate the origin of scatter at fixed halo mass by following the hierarchical buildup of both dark and stellar mass in simulations using abundance matching as a function of cosmic time. They find that $\sigma_{{\rm log}M}$ from merging alone (e.g., `ex-situ' stellar mass growth) can account for 0.16 dex of scatter at cluster-scale halo masses. This is in excellent agreement with \cite{kravtsov_etal:14}, but the comparison to the BOSS results is more nuanced. Our SHMR indicates that the average $10^{14}$ $h^{-1}\,$M$_\odot$\ halo contains at $10^{11.5}$ $$M$_\odot$ galaxy, but once binned by stellar mass, the average halo at that mass scale is $\sim 10^{13}$ $h^{-1}\,$M$_\odot$. The \cite{gu_etal:16} simulations indicate that, at that halo mass, `in-situ' processes dominate. The abundance matching analyses indicate that the fraction of stellar mass from merging for $M_{\rm halo}\sim 10^{13}$ $h^{-1}\,$M$_\odot$\ is low, $\lesssim 10\%$ from \cite{moster_etal:13} and $\lesssim 25\%$ from \cite{behroozi_etal:13}. Thus CMASS galaxies, although they are among the most massive in the universe, still are a sensitive probe of the physics of galaxy formation. \acknowledgements IZ is supported by NSF grant AST-1612085. Funding for the Sloan Digital Sky Survey IV has been provided by the Alfred P. Sloan Foundation, the U.S. Department of Energy Office of Science, and the Participating Institutions. SDSS-IV acknowledges support and resources from the Center for High-Performance Computing at the University of Utah. The SDSS web site is www.sdss.org. SDSS-IV is managed by the Astrophysical Research Consortium for the Participating Institutions of the SDSS Collaboration including the Brazilian Participation Group, the Carnegie Institution for Science, Carnegie Mellon University, the Chilean Participation Group, the French Participation Group, Harvard-Smithsonian Center for Astrophysics, Instituto de Astrof\'isica de Canarias, The Johns Hopkins University, Kavli Institute for the Physics and Mathematics of the Universe (IPMU) / University of Tokyo, Lawrence Berkeley National Laboratory, Leibniz Institut f\"ur Astrophysik Potsdam (AIP), Max-Planck-Institut f\"ur Astronomie (MPIA Heidelberg), Max-Planck-Institut f\"ur Astrophysik (MPA Garching), Max-Planck-Institut f\"ur Extraterrestrische Physik (MPE), National Astronomical Observatory of China, New Mexico State University, New York University, University of Notre Dame, Observat\'ario Nacional / MCTI, The Ohio State University, Pennsylvania State University, Shanghai Astronomical Observatory, United Kingdom Participation Group, Universidad Nacional Aut\'onoma de M\'exico, University of Arizona, University of Colorado Boulder, University of Oxford, University of Portsmouth, University of Utah, University of Virginia, University of Washington, University of Wisconsin, Vanderbilt University, and Yale University.
462c2f4624231dc11f473b6d1161c0ea6be7eef4
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In tokamaks, non-axisymmetric magnetic field perturbations such as toroidal field ripple, error fields and perturbation fields from Edge Localized Mode (ELM) mitigation coils produce non-ambipolar radial transport at non-resonant flux surfaces occupying most of the plasma volume. The toroidal torque associated with this transport significantly changes the toroidal plasma rotation -- an effect known as neoclassical toroidal viscosity~\cite{zhu06-225002,shaing09-075015,Park2009-65002,shaing10-025022,kasilov14-092506,Shaing2015-125001} (NTV). At low collisionalities, resonant transport regimes~\cite{yushmanov82-1123,yushmanov90-117}, namely superbanana plateau~\cite{shaing09-035009,shaing15-905810203}, bounce and bounce-transit (drift-orbit) resonance regimes~\cite{shaing09-075015}, have been found to play an important role in modern tokamaks, in particular in ASDEX Upgrade\red{~\cite{martitsch16-074007}}. In these regimes, which emerge if perturbation field amplitudes are small enough, transport coefficients become independent of the collision frequency (form a plateau). The interaction of particles with the (quasi-static) electromagnetic field in these plateau-like regimes is a particular case of collisionless wave-particle interaction with time dependent fields and can be described within quasilinear theory. The most compact form of this theory in application to a tokamak geometry is obtained in canonical action-angle variables~\cite{Kaufman1972-1063,Hazeltine1981-1164,Mahajan1983-700,becoulet91-137,timofeev94-336,kominis08}. Here, this formalism is applied to ideal quasi-static electromagnetic perturbations, which can be described in terms of flux coordinates. As a starting point, the Hamiltonian description of the guiding center motion in those coordinates in general 3D magnetic configurations (see, e.g., Refs.~\onlinecite{White1982-575,White1984-2455,White1990-845}) is used. For the particular case of Boozer coordinates the perturbation theory is constructed with respect to non-axisymmetric perturbations of the magnetic field module, which is the only function of angles relevant for neoclassical transport. \noindent The purpose of this paper is twofold: The first aim is to describe the NTV in all quasilinear resonant regimes in a unified form using the standard Hamiltonian formalism and to develop a respective numerical code allowing for fast NTV evaluation in these regimes without any simplifications to the magnetic field geometry. The second aim is to benchmark this approach with the quasilinear version of the NEO-2 code\red{~\cite{kasilov14-092506,martitsch16-074007}} which treats the general case of plasma collisionality. Since particular resonant regimes described in literature basically agree with the Hamiltonian approach within their applicability domains, such a benchmarking means also the benchmarking of NEO-2 against those results. The structure of the paper is as follows. In section~\ref{sec:transeqs}, basic definitions are given and two different quasilinear expressions for the toroidal torque density are derived for the general case of small amplitude quasi-static electromagnetic perturbations. In section~\ref{sec:tokamak} the perturbation theory for ideal perturbations described by small corrugation of magnetic surfaces in flux coordinates is outlined, and expressions for the canonical action-angle variables are given. In section~\ref{sec:neotorque} expressions for non-axisymmetric transport coefficients are derived, and in section~\ref{sec:numerics} the numerical implementation of the Hamiltonian formalism in the code NEO-RT is presented and its results compared with the results of NEO-2 code for typical resonant transport regimes. The results are summarized in section~\ref{sec:conclosion}. \section{Transport equations and toroidal torque in Hamiltonian variables} \label{sec:transeqs} \noindent In Hamiltonian variables the kinetic equation can be compactly written in the form \be{kineqnonlin} \difp{f}{t}+\left\{f,H\right\}=\hat L_c f, \end{equation} where $\hat L_c$ is the collision operator and \be{pois} \left\{f,g\right\} \equiv \difp{f}{{\bf r}}\cdot\difp{g}{{\bf p}}-\difp{f}{{\bf p}}\cdot\difp{g}{{\bf r}} = \difp{f}{\theta^i}\difp{g}{J_i}-\difp{f}{J_i}\difp{g}{\theta^i} = \difp{}{\theta^i}\left(f\difp{g}{J_i}\right) - \difp{}{J_i}\left(f\difp{g}{\theta^i}\right) \end{equation} is the Poisson bracket which is invariant with respect to the canonical variable choice. Here, $({\bf r},{\bf p})$ are Cartesian coordinates and canonical momentum components, $(\boldsymbol{\theta},{\bf J})$ are canonical angles and actions specified later, summation over repeated indices is assumed, and bold face describes a whole set of three variables (e.g. $\boldsymbol{\theta}=(\theta^1,\theta^2,\theta^3)$). In the following derivations, straight field line flux coordinates ${\bf x}=(r,\vartheta,\varphi)$ are used with a specific definition of the flux surface label (effective radius) such that $\langle|\nabla r|\rangle=1$, where the neoclassical magnetic flux surface average is given by \be{fluxav} \langle a \rangle=\frac{1}{S}\int\limits_{-\pi}^\pi {\rm d} \vartheta\int\limits_{-\pi}^\pi {\rm d} \varphi\;\sqrt{g}\; a, \qquad S=\int\limits_{-\pi}^\pi {\rm d} \vartheta\int\limits_{-\pi}^\pi {\rm d} \varphi\;\sqrt{g}, \end{equation} and $\sqrt{g}$ is the metric determinant. Due to the above definition of $r$, quantity $S$ has the meaning of the flux surface area. \noindent Multiplying~\eq{kineqnonlin} by a factor $a\,\delta\left(r-r_c\right)$ where $a=a(\boldsymbol{\theta},{\bf J})=a({\bf r},{\bf p})$ is some function of particle position in the phase space and $r_c=r_c(\boldsymbol{\theta},{\bf J})=r({\bf r}_c(\boldsymbol{\theta},{\bf J}))$ is the particle effective radius expressed via phase space variables, integrating over the phase space and dividing the result by the flux surface area $S$ leads to a generalized conservation law \be{conslaw} \difp{A}{t}+\frac{1}{S}\difp{}{r}S\Gamma_A=s_A+s_{A}^{(c)}, \end{equation} where \bea{pphidens} A=A(t,r) &\equiv& \frac{1}{S} \int{\rm d}^3 \theta \int{\rm d}^3 J\; \delta\left(r-r_c\right) a f = \frac{1}{S}\int{\rm d}^3 r \delta\left(r-r_c\right) \int{\rm d}^3 p\; a f \nonumber \\ &=&\frac{1}{S}\int\limits_{-\pi}^\pi {\rm d} \vartheta \int\limits_{-\pi}^\pi {\rm d} \varphi \; \sqrt{g}\int{\rm d}^3 p\; a f =\left\langle \int{\rm d}^3 p\; a f \right\rangle, \end{eqnarray} where $\delta(\dots)$ is the Dirac delta function. Generalized magnetic surface averaged flux and source densities in~\eq{conslaw} are given, respectively, by \bea{fluxsource} \Gamma_A &\equiv& \frac{1}{S}\int{\rm d}^3 \theta \int{\rm d}^3 J\; \delta\left(r-r_c\right) \left\{r_c,H\right\} a f, \\ s_A &\equiv& \frac{1}{S}\int{\rm d}^3 \theta \int{\rm d}^3 J\; \delta\left(r-r_c\right) \left\{a,H\right\} f, \end{eqnarray} where the second representation of the Poisson bracket~\eq{pois} has been used for these expressions, and the collisional source density is \be{nocolint} s_{A}^{(c)}=\left\langle\int{\rm d}^3 p\; a \hat L_c f\right\rangle. \end{equation} For $a=1$ the continuity equation is obtained with no sources, $s_n=s_n^{(c)}=0$ and surface averaged particle flux density $\Gamma_A=\Gamma$ given by \be{partflux} \Gamma = \frac{1}{S} \int{\rm d}^3 \theta \int{\rm d}^3 J\; \delta\left(r-r_c\right) \left\{r_c,H\right\} f. \end{equation} For $a=p_\varphi$ with \be{pphi} p_\varphi={\bf p}\cdot\difp{{\bf r}}{\varphi}=m_\alpha v_\varphi + \frac{e_\alpha}{c}A_\varphi(r) \end{equation} being the canonical angular momentum, the equation for the canonical angular momentum density is obtained with the source term $s_a=s_{p_\varphi}=T_\varphi^{\rm NA}$ being the toroidal torque density acting on the given species from the electromagnetic field, \be{torque} T_\varphi^{\rm NA} = -\frac{1}{S}\int{\rm d}^3 \theta \int{\rm d}^3 J\; \delta\left(r-r_c\right) \difp{H}{\varphi} f. \end{equation} In Eq.~\eq{pphi}, $v_\varphi={\bf v}\cdot \partial{\bf r}/\partial\varphi$ and $A_\varphi=-\psi_{\rm pol}$ are covariant toroidal velocity and vector potential components, respectively and $\psi_{\rm pol}$ is the normalized poloidal flux. In addition, speed of light $c$, and charge $e_\alpha$ and mass $m_\alpha$ of species $\alpha$ appear in the expression. \noindent One can see that torque density is determined only by the non-axisymmetric part of the distribution function while the particle flux density contains also the axisymmetric contribution. This property of the torque is rather helpful in the nonlinear transport theory which, however, is not the topic of the present paper. A conservation law of the kinematic toroidal momentum, $a=m_\alpha v_\varphi$, is obtained by subtraction of the continuity equation multiplied by $e_\alpha A_\varphi /c$. The source term in this equation is \be{kinmomsource} s_{m_\alpha v_\varphi}=T_\varphi^{\rm NA}+s_{p_\varphi}^{(c)}+\frac{e_\alpha}{c}\sqrt{g}B^\vartheta \Gamma, \end{equation} where $B^\vartheta$ is the poloidal contravariant magnetic field component. Assuming a static momentum balance and estimating $\Gamma_{m_\alpha v_\varphi} \sim m_\alpha v_\varphi \Gamma$, which means that contribution of the radial momentum transport term to this balance is negligible because it scales to the last term in~\eq{kinmomsource} as $q\rho_L R/r^2 \ll 1$, this balance is reduced to $s_{m_\alpha v_\varphi}=0$. Here $q$, $\rho_L$, $R$ and $r$ are safety factor, Larmor radius, major and minor radius, respectively. The result is a flux-force relation~\cite{hirshman78-917,shaing83-3315}, which links particle flux to the torques (a static density equilibrium without particle sources where $\Gamma=0$ demonstrates the fact that $T_\varphi^{\rm NA}$ is indeed a torque density because it balances collisional momentum source density $s_{p_\varphi}^{(c)}$ alone). The presence of the collisional force moment $s_{p_\varphi}^{(c)}$ in the flux force relation indicates that the calculation of torque and radial flux needs a certain caution when using a Krook collision model, which is usually the case in quasilinear ``collisionless'' plateau transport regimes described here. Due to momentum conservation by collisions, collisional torque $s_{p_\varphi}^{(c)}$ provides no contribution to the total torque that is of main interest here, which is not ensured by the simple Krook model. This is the case, in particular, for the ion component in the simple plasma where momentum is largely conserved within this component. Thus, when computing particle flux density in this case, one should keep in mind that direct computation of $\Gamma$ from the quasilinear equation provides a different result as compared to such computation through $T_\varphi^{\rm NA}$ via the flux force relation with no collisional torque $s_{p_\varphi}^{(c)}$, \be{eq:FluxForce} T_{\varphi}^{\rm NA}=-\frac{e_\alpha}{c}\sqrt{g}B^\vartheta \Gamma= -\frac{e_{\alpha}}{c}\frac{{\rm d} \psi_{{\rm pol}}}{{\rm d} r} \Gamma. \end{equation} Since $T_\varphi^{\rm NA}$ is not affected by details of the collision model, this more appropriate definition of $\Gamma$ is assumed below unless otherwise mentioned. \red{ It should be noted that in the standard neoclassical theory~\cite{hinton76-239} momentum conservation terms are usually treated first before any approximations on the collision operator are made thus avoiding the errors of the kind discussed above. } \noindent Further steps are standard for quasilinear theory in action-angle variables~\cite{Kaufman1972-1063}. One presents the Hamiltonian and the distribution function as a sum of the unperturbed part depending on actions only and a perturbation with zero average over canonical angles, $H(\boldsymbol{\theta},{\bf J})=H_0({\bf J})+\delta H(\boldsymbol{\theta},{\bf J})$ and $f(\boldsymbol{\theta},{\bf J})=f_0({\bf J})+\delta f(\boldsymbol{\theta},{\bf J})$, respectively and expands the perturbations into a Fourier series over canonical angles, \be{expser} \delta H(\boldsymbol{\theta},{\bf J})=\sum_{\bf m} H_{\bf m}({\bf J}){\rm e}^{im_k\theta^k}, \qquad \delta f(\boldsymbol{\theta},{\bf J})=\sum_{\bf m} f_{\bf m}({\bf J}){\rm e}^{im_k\theta^k}, \end{equation} where sums exclude ${\bf m}=(0,0,0)$ term. By using a Krook collision term with infinitesimal collisionality, $\hat L_c f = -\nu \delta f\rightarrow 0$, the amplitudes of the perturbed distribution function from the linear order equation follow as \be{linorder} \left\{\delta f,H_0\right\}+\left\{f_0,\delta H\right\}+\nu \delta f = \sum_{\bf m}\left( \left(im_k\Omega^k+\nu\right)f_{\bf m}-iH_{\bf m} m_k\difp{f_0}{J_k} \right) {\rm e}^{im_k\theta^k}=0. \end{equation} Here, $\Omega^{k}=\partial H/\partial J_{k}$ are canonical frequencies, and the time derivative has been omitted as small compared to all canonical frequencies in case of quasi-static perturbations of interest here. A quasilinear equation is obtained by retaining only secular, angle-independent terms in the second order equation, \be{eq:quasilinear} \difp{f_0}{t}+\overline{\left\{\delta f,\delta H\right\}} = \difp{f_0}{t}-\sum_{\bf m} m_k\difp{Q_{\bf m}}{J_k}=0, \end{equation} where the over-line stands for the average over the angles, and \be{eq:Qm} Q_{{\bf m}}=Q_{{\bf m}}({\bf J})=\frac{\pi}{2}|H_{{\bf m}}|^{2}\delta(m_{j}\Omega^{j})m_{k}\difp{f_{0}}{J_{k}} \end{equation} contains a resonance condition in the argument of a delta function that follows from the limit $\nu\rightarrow 0$. The knowledge of $f_{\bf m}$ is already sufficient for the evaluation of torque densities from Eq.~\eq{torque} where the derivative over $\varphi$ is equivalent to a derivative over the canonical toroidal phase $\theta^3$, \be{torque_ql} T^{\rm NA}_{\varphi}=-\frac{m_3}{S}\int d^{3}\theta\int d^{3}J\delta(r-r_c)\sum_{\bf m} Q_{\mathbf{m}}, \end{equation} and of the particle flux from~\eq{partflux} \be{partlux_ql} \Gamma_{\rm F}=-\frac{1}{S}\int d^{3}\theta\int d^{3}J\delta(r-r_c)\sum_{\mathbf{m}}m_{k} \difp{r_{c}}{J_k}Q_{\mathbf{m}}, \end{equation} which is distinguished here from~\eq{eq:FluxForce} by subscript $F$. Alternatively the same expressions are obtained computing the conservation laws using the quasilinear equation~\eq{eq:quasilinear} as a starting point~\cite{Hazeltine1981-1164}. If the collision model does not conserve the parallel momentum such as e.g. the Krook model, direct calculations of the torque in terms of viscosity~\cite{shaing09-075015} and calculation of the torque through particle flux~\cite{Park2009-65002} using the force-flux relation~\eq{eq:FluxForce} may lead to different results. This difference, however, is negligible in resonant transport regimes where details of the collision model are not important, and collisionality can be treated as infinitesimal. \section{Tokamak with ideal non-axisymmetric quasi-static perturbations} \label{sec:tokamak} \subsection{Canonical Hamiltonian variables for perturbed equilibria} \noindent Often in quasilinear theory in action-angle variables, both, the unperturbed and perturbed Hamiltonian correspond to physically possible motion with separation of the unperturbed electromagnetic field and its perturbation in real space. However, there is no mathematical need to do so. In particular, if the perturbed equilibrium is ideal such that it can be described in flux coordinates, it is more convenient to restrict the perturbations only to those quantities in the Hamiltonian which violate the axial symmetry. In case of Boozer coordinates and also in many cases described in Hamada coordinates the only important quantity is the magnetic field module which is generally adopted for the construction of perturbation theory for NTV models~\cite{shaing09-075015,Park2009-65002,shaing10-025022,kasilov14-092506,shaing15-905810203}. Thus the guiding center Lagrangian~\cite{Littlejohn1983-111} is transformed here to flux coordinates ${\bf x}=(r,\vartheta,\varphi)$ as a starting point, \be{gclagr} L=m_{\alpha}v_{\parallel}h_{r}\dot{r} +(m_{\alpha}v_{\parallel}h_{\vartheta}+\frac{e_\alpha}{c}A_{\vartheta})\dot{\vartheta} +(m_{\alpha}v_{\parallel}h_{\varphi}+\frac{e_{\alpha}}{c}A_{\varphi})\dot{\varphi} +J_{\perp}\dot{\phi}-H, \end{equation} where lower subscripts denote covariant components (in particular, $A_\vartheta=A_\vartheta(r)=\psi_{\rm tor}$ is the covariant poloidal component of the vector potential, which is equal to the normalized toroidal flux and $A_r=0$), $\mathbf{h}=\mathbf{B}/B$ is the unit vector along the magnetic field, $v_{\parallel}$ is the parallel velocity, $J_{\perp}=m_{\alpha}v_{\perp}^{2}/(2\omega_{c})$ is the perpendicular adiabatic invariant with $v_\perp$ and $\omega_{c}$ being the perpendicular velocity and cyclotron frequency, respectively, $\phi$ is the gyrophase and the Hamiltonian is given explicitly below in Eq.~\eq{expandham}. The canonical form of the Lagrangian is obtained by transforming the toroidal angle $\varphi$ to \be{modvarphi} \varphi_{H} =\varphi-\frac{c m_{\alpha}v_{\parallel}h_{r}}{e_\alpha A_{\varphi}^{\prime}}, \end{equation} where the prime stands for a radial derivative. Omitting a total time derivative, the Lagrangian transforms to \be{} L=p_\vartheta \dot{\vartheta} +p_\varphi\dot{\varphi}_{H} +J_{\perp}\dot{\phi}-H+ \frac{c m_{\alpha}^2 v_{\parallel}h_{\varphi}}{e_\alpha} \frac{d}{dt}\left(\frac{v_{\parallel}h_{r}}{A_{\varphi}^{\prime}}\right), \end{equation} where \be{momenta} p_\vartheta=m_{\alpha}v_{\parallel}h_{\vartheta}({\bf x})+\frac{e_{\alpha}}{c}A_{\vartheta}(r), \qquad p_\varphi=m_{\alpha}v_{\parallel}h_{\varphi}({\bf x})+\frac{e_{\alpha}}{c}A_{\varphi}(r) \end{equation} are canonical momenta in guiding center approximation, and the last term is of the next order in $\rho_{\parallel}=v_{\parallel}/\omega_{c}$ and should therefore be neglected. Transformation~\eq{modvarphi} affects only a small non-axisymmetric part of the field and is different from the one of Refs.~\cite{White1984-2455,White1990-845} where the poloidal angle $\vartheta$ is modified instead. Alternatively, for collisionless transport regimes of interest here, one can simply ignore the covariant magnetic field component $B_r$ because it does not contribute to the radial guiding center velocity, and its contribution to the rotation velocity vanishes on a time scale larger than bounce time. \noindent Since the momenta are the independent variables, Eq.~\eq{momenta} should be regarded as a definition of $r$ and $v_\parallel$. For the construction of perturbation theory in Boozer coordinates being the main choice here, the last quantity is redefined via the unperturbed parallel velocity $v_{0\parallel}$ as follows, \be{vparunpert} v_\parallel = v_{0\parallel}\frac{B({\bf x})}{B_0(r,\vartheta)} \end{equation} where subscript 0 corresponds to the axisymmetric part of the respective quantity. Due to such redefinition, $r$ and $v_{0\parallel}$ do not depend on the toroidal angle $\varphi$ because in Boozer coordinates this dependence vanishes in both expressions in~\eq{momenta} due to $h_{\vartheta,\varphi}=B_{\vartheta,\varphi}(r)/B({\bf x})$. For the comparison with the results obtained in Hamada coordinates for the superbanana-plateau regime, which is a resonant regime described by the bounce-averaged equation, the definition of the unperturbed parallel velocity is opposite to~\eq{vparunpert}, $v_\parallel = v_{0\parallel} B_0/B$. With this redefinition, angular covariant components of ${\bf h}$ in~\eq{vparunpert} are transformed within linear order in the perturbation field as follows, \be{transcovham} B h_k = B_0 h_{0k} +\difp{\delta \chi}{x^k}-h_{0k}h_0^j \difp{\delta \chi}{x^j}, \qquad k=2,3, \end{equation} where $\delta \chi$ is the non-axisymmetric perturbation of a function $\chi$ which enters the definition of co-variant magnetic field components in Hamada coordinates $B_k$ via their flux surface averages $\bar B_k=\bar B_k(r)$, with $B_k=\bar B_k+\partial \chi/\partial x^k$. Terms with $\delta \chi$, whose contribution in~\eq{transcovham} is orthogonal to the unperturbed magnetic field, can be simply ignored in bounce-averaged regimes because they do not contribute to bounce averaged velocity components. \noindent Thus, the Hamiltonian is expanded in Boozer coordinates up to a linear order in the perturbation field amplitude as follows, \be{expandham} H=\omega_c J_\perp+\frac{m_\alpha v_\parallel^2}{2}+e_\alpha\Phi= \frac{B}{B_0}\omega_{c0} J_\perp+\frac{B^2}{B_0^2}\frac{m_\alpha v_{0\parallel}^2}{2}+e_\alpha\Phi \approx H_0+\delta H, \end{equation} where $\Phi=\Phi(r)$ is the electrostatic potential, \be{H0deltaH} H_0=\omega_{c0} J_\perp+\frac{m_\alpha v_{0\parallel}^2}{2}+e_\alpha\Phi, \qquad \delta H=\left(\omega_{c0} J_\perp+m_\alpha v_{0\parallel}^2\right)\frac{\delta B}{B_0}. \end{equation} The Hamiltonian perturbation $\delta H$ in Hamada coordinates differs from~\eq{H0deltaH} by the opposite sign of the second term in the parentheses, $m_\alpha v_{0\parallel}^2$. This term is usually ignored in tokamaks with large aspect ratio $A$ because for trapped and barely trapped particles which are mainly contributing to NTV at small Mach numbers (at sub-sonic toroidal rotation velocities) it scales to the first term as $1/A$. \subsection{Action-angle variables in the axisymmetric tokamak} \label{ssec:canvar} \noindent Since this subsection deals only with unperturbed motion corresponding to $H=H_0$, the subscript 0 is dropped on all quantities here which are strictly axisymmetric. Here it is convenient to replace the toroidal momentum $p_{\varphi}$, which is now a conserved quantity, by another invariant of motion $r_\varphi$ which describes the banana tip radius for trapped particles~\cite{Hazeltine1981-1164} and is implicitly defined via \be{eq:pphi} \frac{e_{\alpha}}{c}A_{\varphi}(r_{\varphi})=p_{\varphi}. \end{equation} Expanding the vector potential components in~\eq{momenta} over $r-r_\varphi$ up to the linear order and using $A^\prime_\vartheta/A^\prime_\varphi=-{\rm d} \psi_{\rm tor}/{\rm d} \psi_{\rm pol}=-q$, the poloidal momentum is approximated by \be{} p_{\vartheta}=\frac{e_{\alpha}}{c}A_{\vartheta}+\frac{m_{\alpha}v_{\parallel}}{h^{\vartheta}}. \end{equation} In the above formula and in the remaining derivation, all quantities are evaluated at $r=r_{\varphi}$ if not noted otherwise. In this approximation it is possible to express derivatives with respect to $p_{\varphi}$ by radial derivatives. The poloidal action is defined for trapped $(\delta_{\rm t-p}=0)$ and passing $(\delta_{\rm t-p}=1)$ particles by \be{} J_{\vartheta}=\frac{1}{2\pi}\oint d\vartheta p_{\vartheta}=\frac{e_{\alpha}}{c}A_{\vartheta}\delta_{\rm t-p}+J_{\parallel}. \end{equation} The first term cancels when integrating back and forth between the turning points of a trapped orbit. The parallel adiabatic invariant may be written as a bounce average, \be{} J_{\parallel}=\frac{m_{\alpha}\tau_{b}}{2\pi}\left\langle v_{\parallel}^{2}\right\rangle _{b} , \end{equation} with bounce time $\tau_{b}$, orbit time $\tau$ and bounce averaging $\left\langle a(\vartheta)\right\rangle _{b}$ defined by \begin{align} \tau_{b} & =\oint\frac{dl}{v_{\parallel}}=\oint\frac{d\vartheta}{v_{\parallel}h^{\vartheta}},\label{eq:bouncetime}\\ \tau(\vartheta_{0},\vartheta_{{\rm orb}}) & =\int_{\vartheta_{0}}^{\vartheta_{{\rm orb}}}\frac{d\vartheta}{v_{\parallel}h^{\vartheta}},\label{eq:orbittime}\\ \left\langle a(\vartheta)\right\rangle _{b} & =\frac{1}{\tau_{b}}\oint\frac{d\vartheta}{v_{\parallel}h^{\vartheta}}a(\vartheta)=\frac{1}{\tau_{b}}\int_{0}^{\tau_{b}}d\tau\, a(\vartheta_{{\rm orb}}(\vartheta_{0},\tau)).\label{eq:bounceavg} \end{align} Here $a(\vartheta)$ is any function of the poloidal angle and integrals of motion $(J_{\perp},H_{0},s_{\varphi})$ and $\vartheta_{{\rm orb}}(\vartheta_{0},\tau)$ is the (periodic) solution of the unperturbed guiding center equations (the orbit) starting at the magnetic field minimum point $\vartheta_{0}$. Finally, we arrive at the expressions for the three canonical actions in a tokamak~\cite{Kaufman1972-1063}, \begin{align} J_{1} & =J_{\perp}=\frac{m_{\alpha}c}{e}\mu,\nonumber \\ J_{2} & =J_{\vartheta}=\frac{e_{\alpha}}{c}A_{\vartheta}\delta_{\rm t-p}+\frac{m_{\alpha}\tau_{b}}{2\pi}\left\langle v_{\parallel}^{2}\right\rangle _{b},\nonumber \\ J_{3} & =p_{\varphi}=m_{\alpha}v_{\parallel}h_{\varphi}+\frac{e_{\alpha}}{c}A_{\varphi}, \label{ptor} \end{align} where $\mu$ denotes the magnetic moment. Canonical frequencies $\Omega^{k}=\partial H/\partial J_{k}$ are \be{canfreq} \Omega^{1} =\left\langle \omega_{c}\right\rangle _{b}, \qquad \Omega^{2} =\omega_{b}, \qquad \Omega^{3} =q\omega_{b}\delta_{\rm t-p}+\left\langle v_{g}^{\varphi} \right\rangle_b, \end{equation} where the bounce frequency $\omega_{b}=2\pi/\tau_b$ is strictly positive for trapped particles, whereas for passing particles it can take both, positive and negative values. The bounce average of the toroidal precession frequency $v_{g}^{\varphi}$ due to the cross-field drift is separated in two parts, \be{vgphi} \left\langle v_{g}^{\varphi} \right\rangle_b\equiv \Omega_t = \left\langle \frac{v_{\parallel}}{\omega_{c}\sqrt{g}} \difp{}{r}\left(\frac{v_{\parallel}}{h^{\vartheta}}\right) \right\rangle_b =\left<\Omega_{tE}\right>_b+\left<\Omega_{tB}\right>_b. \end{equation} Here, bounce averages of electric drift frequency $\Omega_{tE}$ and magnetic drift frequency $\Omega_{tB}$ are \begin{align} \left<\Omega_{tE}\right>_b &= \Omega_{tE} =-\frac{c}{\psi_{{\rm pol}}^{\prime}}\difp{\Phi}{r},\nonumber \\ \left<\Omega_{tB}\right>_b & =\frac{v^{2}}{\psi_{{\rm pol}}^{\prime}} \left<-\frac{2-\eta B}{2\omega_c}\difp{B}{r} +\frac{1-\eta B}{\omega_c}h^{\vartheta}\left(\difp{B_{\vartheta}}{r}+q\difp{B_{\varphi}}{r}+ B_{\varphi}\frac{dq}{dr}\right)\right>_b,\label{eq:OmtE} \end{align} with equilibrium potential $\Phi$, and velocity space parameterized by velocity module $v$ and the parameter $\eta=v_{\perp}^{2}/(v^{2}B)=2 e_\alpha J_\perp/(c m_\alpha^2 v^2)$. Comparison of magnetic rotation frequency $\left<\Omega_{tB}\right>_b$ given by Eq.~\eq{eq:OmtE} with the expression obtained by bounce averaging of Eq.~(67) of Ref.~\onlinecite{kasilov14-092506} one can notice the absence in the latter expression of a term $B_{\varphi}q^{\prime}(r)$ describing the magnetic shear. This results from using the local neoclassical ansatz as a starting point in the linearized equation for the non-Maxwellian perturbation of the distribution function where the radial derivative of this perturbation is ignored. This local ansatz is the standard method in drift kinetic equation solvers in general 3D toroidal geometries~\cite{hirshman1986-2951,landreman14-042503} and is justified in most transport regimes, but not in resonant regimes, where magnetic drift plays a significant role. As shown in the example below, the shear term may lead to a significant modification of the superbanana resonance condition. This term is retained if linearization is applied after bounce-averaging the kinetic equation~\cite{shaing15-905810203}. \red{It should be noted that the guiding center Lagrangian~\eq{gclagr} used as a starting point here is valid for the general case of the magnetic field, which is not necessarily a force-free field. Therefore, the effects of finite plasma pressure on the toroidal rotation velocity~\cite{connor83-1702} are automatically taken into account in~\eq{eq:OmtE}. } \noindent The canonical angles in the leading order follow as \be{eq:angles} \theta^{1} =\phi-\Delta\phi(\theta^{2},{\bf J}), \qquad \theta^{2} =\Omega^{2}\tau, \qquad \theta^{3} =\varphi_{H}+q\theta^{2}\delta_{\rm t-p}-q\vartheta_{{\rm orb}}(\vartheta_{0},\tau), \end{equation} where $\Delta \phi$ is a periodic function of the canonical poloidal variable $\theta^2$. Since according to~\eq{eq:angles} $\phi$ and $\varphi_H$ differ from the respective canonical angles $\theta^{1}$ and $\theta^{3}$ by additional terms depending on $\theta^2$ only and $\vartheta$ depends only on $\theta^{2}$, the spectrum $a_{\mathbf{m}}$ in canonical angles of a function given by a single harmonic (l,n) of the original angles $\phi,\varphi$, \be{spectrum} a(\phi,\vartheta,\varphi)=a_{ln}(\vartheta)e^{i\left(l\phi+n\varphi\right)}=\sum_{\mathbf{m}}a_{\mathbf{m}}e^{im_{k}\theta^{k}}, \end{equation} contains non-zero contributions only from canonical modes with $m_{1}=l$ and $m_{3}=n$. In particular, for the gyroaverage $\left\langle a\right\rangle _{g}$ described by the harmonic $l=0$ of function $a$, one obtains to the leading order in $\rho_{\parallel}$ \be{eq:amn} a_{\bf m}=\left\langle a_{0n}(\vartheta)e^{inq\vartheta-i\left(m_{2}+nq\delta_{\rm t-p}\right)\omega_{b}\tau}\right\rangle _{b}, \end{equation} where ${\bf m}=(0,m_{2},n)$. \section{Neoclassical toroidal viscous torque and related radial transport} \label{sec:neotorque} \noindent For NTV applications, where the perturbed Hamiltonian~\eq{H0deltaH} is independent of gyrophase, only harmonics with first canonical mode number $m_{1}=0$ can contribute to fulfill the resonance condition inside the $\delta$ distribution of Eq. (\ref{eq:Qm}), and the latter is reduced to \be{eq:resonance} m_{j}\Omega^{j}=0\,\,\,\rightarrow\,\,\, (m_{2}+nq\delta_{\rm t-p})\omega_{b}+n\Omega_t=0. \end{equation} This equation includes all regimes of interest here: The superbanana-plateau resonance is described by the condition $m_{2}=0$ for trapped particles. For passing particles, $m_{2}=0$ corresponds to a transit resonance. This is the only resonance remaining in the infinite aspect ratio limit, where it reduces to the usual Cherenkov (TTMP) resonance. Finite mode numbers $m_{2}$ correspond to bounce and bounce-transit resonances for trapped and passing particles, respectively. Resonances where both, parallel motion and cross-field drift determine the resonance condition, i.e. all resonances except the superbanana-plateau resonance are mentioned below as ``drift-orbit'' resonances. \noindent Due to the properties of the spectrum~\eq{spectrum}, which follow from the axial symmetry of the unperturbed field, separate toroidal harmonics of perturbation Hamiltonian produce independent contributions to the torque and particle flux density. Therefore it is sufficient to assume the perturbation field $\delta B$ in~\eq{H0deltaH} in the form of a single toroidal harmonic, \be{} \delta B={\rm Re}(B_{n}(\vartheta)e^{in\varphi}). \end{equation} Making use of Eq. (\ref{eq:amn}), the associated modes of the Hamiltonian perturbation result in \be{eq:Hm} H_{\mathbf{m}}=\left\langle \left(m_{\alpha}v_{0\parallel}^{2}(\vartheta) +\frac{e_{\alpha}}{m_{\alpha}c}J_{\perp}B_{0}(\vartheta)\right)\frac{B_{n}(\vartheta)}{B_{0}(\vartheta)} e^{inq\vartheta-i\left(m_{2}+nq\delta_{\rm t-p}\right)\omega_{b}\tau}\right\rangle _{b}. \end{equation} For small enough perturbations, which are considered here, quasilinear effects are weak and thus $f_{0}$ is close to a drifting Maxwellian, \be{f0def} f_{0}=\frac{n_{\alpha}}{\left(2\pi m_{\alpha}T_{\alpha}\right)^{3/2}}e^{(e_{\alpha}\Phi-H_{0})/T_{\alpha}}, \end{equation} with parameters depending on $r_\varphi$ but not $r$. This Maxwellian differs from a local Maxwellian by linear terms in $\rho_\parallel$, which, as shown below, provide negligible contributions in resonant regimes with quasi-static perturbations. \red{ Let us check that within first order in $\rho_\parallel$ the toroidally drifting Maxwellian Eq.~\eq{f0def} is a solution to the axisymmetric kinetic equation for ions valid in all collisionality regimes in the absence of temperature gradients. For this purpose it is more convenient to replace the approximate expression for the canonical angular momentum~\eq{ptor} valid in first order in Larmor radius by the exact expression, \be{ptorex} p_\varphi=m_\alpha v_\varphi +\frac{e_\alpha}{c}A_\varphi, \end{equation} where $v_\varphi$ is the toroidal covariant component of the total particle velocity including the Larmor gyration and $A_\varphi$ is evaluated at the exact particle position $r_{\rm c}$ (not at the guiding center position denoted with $r$ here) related to $r_\varphi$ as follows, \be{exradtorphi} r_{\rm c}=r_\varphi+\frac{c m_\alpha v_\varphi}{e_\alpha \sqrt{g}B^\vartheta}. \end{equation} Then, the unperturbed distribution function~\eq{f0def} up to linear order in Larmor radius is \be{f0uptolin} f_{0}=\frac{n_{\alpha}}{\left(2\pi m_{\alpha}T_{\alpha}\right)^{3/2}} \exp\left(-\frac{m_\alpha}{2T_\alpha} \left({\bf v}_{\rm pol}^2+g_{\varphi\varphi} \left(v^\varphi-V_\alpha^\varphi\right)^2\right) \right), \end{equation} where $g_{\varphi\varphi}=R^2$ and all functions of radius are evaluated at $r_c$. Here, ${\bf v}_{\rm pol}$ and $v^\varphi$ are total poloidal and contra-variant component of the total toroidal particle velocity, respectively, and the contra-variant toroidal component of the ion flow velocity is explicitly given by \be{torionflow} V_\alpha^\varphi = \frac{c}{\sqrt{g}B^\vartheta} \left(E_r-\frac{T_\alpha}{e_\alpha n_\alpha}\difp{n_\alpha}{r_c}\right). \end{equation} In a simple plasma where the momentum is approximately conserved within a single ion component, the drifting Maxwellian~\eq{f0uptolin} annihilates the collision term. As a straightforward consequence of~\eq{f0uptolin}, the poloidal ion flow velocity is zero at all collisionalities if the temperature gradient is absent. Respectively, the toroidal flow velocity~\eq{torionflow} is the same as given by ideal MHD (see, e.g., Eq.~(6) of Ref.~\onlinecite{kasilov14-092506}). In the presence of temperature gradients and in a multi-species plasma Eq.~\eq{f0def} satisfies the kinetic equation only in zero order over Larmor radius. Additional anisotropic terms which appear in the first order solution are of same order as in~\eq{f0def}, nevertheless, they provide a negligible contribution for the following reason. } When substituting~\eq{f0def} in~\eq{eq:Qm} one can notice that only derivatives of the parameters over $r_\varphi$ provide non-zero contributions in presence of resonance condition, \be{substres} \delta(m_j\Omega^j) m_{k}\difp{f_{0}}{J_{k}}= - \delta(m_j\Omega^j) \frac{nc(A_{1}+A_{2}u^{2})}{e_\alpha} \frac{{\rm d} r}{{\rm d} \psi_{\rm pol}}f_0, \end{equation} where $u=v/v_{T}$ is the velocity module $v$ normalized by the thermal velocity $v_{T}=\sqrt{2T_\alpha/m_\alpha}$ and \be{thermforces} A_{1}=\frac{1}{n_{\alpha}} \difp{n_{\alpha}}{r}+\frac{e_{\alpha}}{T_{\alpha}}\difp{\Phi}{r}- \frac{3}{2T_{\alpha}}\difp{T_{\alpha}}{r}, \qquad A_{2}=\frac{1}{T_{\alpha}}\difp{T_{\alpha}}{r}, \end{equation} are the thermodynamic forces which are evaluated at $r=r_\varphi$. For any function $F$ of actions expressed in the form $F=F(H_0,J_\perp,p_\varphi)$, only the derivative over $p_\varphi$ remains in expressions such as~\eq{substres} because the derivative over $J_1$ enters with factor $m_1=0$ only, and the derivative over $H_0$ enters with factor $m_k\Omega^k$ which is zero due to the resonance condition~\eq{eq:resonance} (the energy is preserved for static perturbations). Therefore the contribution of the linear correction in $\rho_\parallel$ to the unperturbed distribution function which depends also on $J_\perp$ would contribute in~\eq{substres} only in the form of its derivative over $r_\varphi$ which is of higher order in $\rho_\parallel$ than such a derivative of the Maxwellian retained in~\eq{substres}. In the expression for the torque density~\eq{torque_ql} one can ignore finite Larmor radius effects together with finite orbit width effects in $r_c$ in the argument of the $\delta$-function by setting $r_c \approx r_\varphi$. Then an integration over $J_3=p_\varphi$ results in a replacement of $r_\varphi$ by $r$ in the subintegrand, and the integration over canonical angles is simply replaced by a factor $8\pi^3$. Changing the integration variables of the remaining integral over $J_1$ and $J_2$ to $v$ and $\eta$ and transforming the resulting $T_\varphi^{\rm NA}$ to a particle flux density using the flux-force relation~\eq{eq:FluxForce} results in \be{Gamma_ntv} \Gamma = \frac{2\pi^2 n m_{\alpha}^3 c}{e_\alpha S} \int_{0}^{\infty}dv\, v^{3}\int_{0}^{1/B_{{\rm min}}}d\eta\tau_{b}\sum_{{\rm m}_{2}}Q_{\mathbf{m}}. \end{equation} Substituting $Q_{\mathbf{m}}$ in~\eq{Gamma_ntv} explicitly and using the representation of $\Gamma$ in terms of thermodynamic forces~\eq{thermforces}, $\Gamma=-n_\alpha(D_{11}A_{1}+D_{12}A_{2})$, resonant transport coefficients follow as \be{eq:D1x} D_{1k}=\frac{\pi^{3/2}n^{2}c^{2}v_{T}}{e_{\alpha}^{2}S}\frac{{\rm d} r}{{\rm d} \psi_{\rm pol}} \int_{0}^{\infty}du\, u^{3}e^{-u^{2}} \sum_{m_{2}}\sum_{{\rm res}}\left(\tau_{b}|H_{\mathbf{m}}|^{2} \left|m_{2}\frac{\partial\omega_{b}}{\partial\eta}+n\difp{\Omega^{3}}{\eta}\right|^{-1}\right)_{\eta=\eta_{{\rm res}}}w_{k}, \end{equation} where $w_{1}=1$ for $D_{11}$ and $w_{2}=u^{2}$ for $D_{12}$, respectively. In this expression the $\delta$ term inside $Q_{\mathbf{m}}$ has been evaluated with respect to $\eta$, and $\eta_{\mathrm{res}}$ are (generally multiple) roots of Eq. (\ref{eq:resonance}). \noindent In the direct definition of the flux~\eq{partlux_ql} one can, again, replace $r_c$ by $r_\varphi$ in the argument of the $\delta$-function. Using the same arguments as in~\eq{substres} for ignoring the linear order term in $\rho_\parallel$ inside $f_0$, one can ignore the difference between $r_c$ and $r_\varphi$ in the derivative $m_k \partial r_c /\partial J_k$. Then $\Gamma_{\rm F}$ given by~\eq{partlux_ql} leads to a result identical to~\eq{Gamma_ntv}. \red{ \noindent The equivalence of $\Gamma_F$ and $\Gamma$ obtained here using a simple Krook collision model indicates that momentum conservation plays no role in resonant transport. While in case of superbanana plateau and bounce resonance regimes this can be concluded, in particular, from the fact that all resonant particles are trapped particles, which lose parallel momentum obtained from the perturbation field within a single bounce period due to magnetic mirroring, this explanation cannot be used for transit and bounce-transit resonances where passing particles are responsible. The general reason for the conclusion above is different and is actually the same as the reason to ignore the anisotropic correction in the unperturbed distribution function~\eq{f0def}. As already mentioned, the resonant interaction with a static perturbation field does not modify the total particle energy $H_0$ in contrast to the case of time dependent perturbations where the change of total energy scales with perturbation frequency $\omega$ due to the more general resonance condition $m_j\Omega^j=\omega$. In addition, since cyclotron resonances, $m_1\ne 0$, cannot be realized for bulk particles (particles with energies of the order of thermal energy), the perpendicular adiabatic invariant $J_\perp$ is also conserved. Consequently, the change of parallel velocity $v_\parallel=v_\parallel({\bf x},H_0,J_\perp)$ and of bounce frequency $\omega_b=\omega_b(r_\varphi,H_0,J_\perp)$, which represents the bounce averaged parallel momentum of passing particles, can only appear through the change by the resonant interaction of the particle position in space ${\bf x}$ (``radial'' variable $r_\varphi)$. In case of mild radial electric fields where the variation of the potential energy along the guiding center orbit with finite radial width is small (of the order of Larmor radius) compared to the thermal energy what corresponds to sub-sonic rotations (small toroidal Mach numbers), the contribution of the kinematic momentum change to the overall canonical momentum change is small of the same order too. Therefore, the momentum restoring term in the collision operator provides a correction proportional to the toroidal Mach number assumed to be small in the present paper. } \noindent As mentioned above, the Hamiltonian approach includes all quasilinear resonant transport regimes in a unified form where these regimes correspond to different resonances~\eq{eq:resonance}. In particular the expression for the contribution of the $m_2=0$ resonance for trapped particles corresponds to the superbanana-plateau regime and differs from such a result of Ref.~\onlinecite{shaing15-905810203} only in notation. The results for drift-orbit resonances $m_2\ne 0$ mostly agree with Ref.~\onlinecite{shaing09-075015} up to simplifications of the magnetic field geometry and the neglected magnetic drift in this reference. Differences appear only in resonant contribution of passing particles on irrational flux surfaces arising from the representation in Eqs.~(37) and~(38) of Ref.~\onlinecite{shaing09-075015} of an aperiodic function by a Fourier series. \section{Numerical implementation and results} \label{sec:numerics} \noindent In the scope of this work the coefficients (\ref{eq:D1x}) are computed numerically in the newly developed code NEO-RT for the general case of a perturbed tokamak magnetic field specified in Boozer coordinates. Bounce averages are performed via numerical time integration of zero order guiding center orbits as specified in~\eq{eq:amn}. An efficient numerical procedure for finding the roots in Eq. (\ref{eq:resonance}) is realized using the scalings \bea{} \omega_{b} & =u\bar{\omega}_{b}(\eta),\\ \left<\Omega_{tB}\right>_b & =u^{2}\bar{\Omega}_{tB}(\eta). \end{eqnarray} Normalized frequencies $\bar{\omega}_{b}$ and $\bar{\Omega}_{tB}$ (relatively smooth functions) are precomputed on an adaptive $\eta$-grid and interpolated via cubic splines in later calculations. \noindent For testing and benchmarking, a tokamak configuration with circular concentric flux surfaces and safety factor shown in Fig.~\ref{fig:sb-plateau} is used (the same as in Ref.~\onlinecite{kasilov14-092506}) and results are compared to calculations from the NEO-2 code. The perturbation field amplitude in Eq. (\ref{eq:Hm}) is taken in the form of Boozer harmonics \be{} B_{n}(\vartheta)=\varepsilon_{M}B_{0}(\vartheta)e^{im\vartheta}. \end{equation} Two kinds of perturbations are considered here: a large scale perturbation with $(m,n)=(0,3)$ referred below as ``RMP-like case'' because of the toroidal wavenumber typical for perturbations produced by ELM mitigation coils, and a short scale perturbation with $(m,n)=(0,18)$ typical for the toroidal field (TF) ripple. The remaining parameters are chosen to be representative for a realistic medium-sized tokamak configuration. In the plots, transport coefficients $D_{1k}$ are normalized by (formally infinitesimal) $\varepsilon_{M}^{2}$ times the mono-energetic plateau value \be{plateau} D_{p}=\frac{\pi qv_{T}^{3}}{16\, R\,\bar{\omega}_{c}^{2}}, \end{equation} where $R$ is the major radius, and the reference gyrofrequency $\bar{\omega}_{c}$ is given by the $(0,0)$ harmonic of $\omega_{c}$. Radial dependencies are represented by the flux surface aspect ratio $A=\left(\psi^a_{\rm tor}/\psi_{\rm tor}\right)^{1/2}R/a$ of the current flux surface where $a$ is the minor radius of the outermost flux surface and $\psi^a_{\rm tor}$ the toroidal magnetic flux at this surface. The radial electric field magnitude is given in terms of the toroidal Mach number $M_t \equiv R \Omega_{tE}/v_T$. In all plots there are at least 4 data points between subsequent markers. \begin{figure} \includegraphics{fig1a_supban_Mt0p036}\quad{}\includegraphics{fig1b_supban_Mtm0p036}\\ \includegraphics{fig1c_supban_res_Mt0p036}\quad{}\includegraphics{fig1d_supban_res_Mtm0p036} \protect\caption{Radial dependence of superbanana plateau $D_{11}$ in the RMP case for Mach number $M_t=0.036$ (left) and $-0.036$ (right). Comparison of Hamiltonian approach (NEO-RT) to analytical formula by Shaing~\cite{shaing09-035009} (solid line). Results with ($\diamond$) and without magnetic shear ($\square$) in the magnetic drift frequency (\ref{eq:OmtE}). A safety factor profile (dash-dotted) is shown on the second axis of the upper right plot. The lower plots show resonance lines ranging from deeply trapped ($\Delta\bar{\eta}=0$) to trapped passing boundary ($\Delta\bar{\eta}=1$) at flux surfaces of aspect ratio $A=5$ (solid) and $A=10$ (dashed). \label{fig:sb-plateau}} \end{figure} \noindent Fig.~\ref{fig:sb-plateau} shows the radial dependence of the transport coefficient $D_{11}$ in the superbanana plateau regime for the RMP-like perturbation for both positive and negative radial electric field. For this benchmarking case the relation between toroidal precession frequencies due to the ${\bf E}\times{\bf B}$ drift, $\Omega_{tE}$, and due to the magnetic drift $\Omega_{tB}$, has been fixed by setting the reference toroidal magnetic drift frequency $\Omega^{\rm ref}_{tB}\equiv cT_{\alpha}/(e_{\alpha} \psi^a_{\rm tor})$ (not the actual $\Omega_{tB}$) equal to $\Omega_{tE}$. Additional curves are shown for calculations where the magnetic shear term ($dq/dr$) in Eq.~\eq{eq:OmtE} has been neglected. The results are compared to the analytical formula for the large aspect ratio limit by Shaing~\cite{shaing09-035009}. Resonance lines in velocity space are plotted below the radial profiles for a flux surface relatively close to the axis ($A=10$) and one further outwards ($A=5$). Here $\Delta\bar{\eta}=(\eta-\eta_{\mathrm{tp}})/(\eta_{\mathrm{dt}}-\eta_{\mathrm{tp}})$ is the distance to the trapped passing boundary $\eta_{\mathrm{tp}}$ normalized to the trapped region between trapped-passing boundary $\eta_{\mathrm{tp}}$ and deeply trapped $\eta_{\mathrm{dt}}$. For flux surfaces with $A>10$ magnetic shear plays a small role due to the flat safety factor profile in the present field configuration: The diffusion coefficient $D_{11}$ is nearly identical to the result without shear and stays close to the analytical result for the large aspect ratio limit. For aspect ratio $A=10$, the agreement between NEO-2 calculations and large aspect ratio limit of Ref.~\onlinecite{shaing09-035009} has been demonstrated earlier in Ref.~\onlinecite{kasilov14-092506}. At larger radii, where the q profile becomes steep, a significant deviation between the cases with and without magnetic shear term is visible. This can be explained by the strong shift of the resonance lines due to the shear term in the rotation frequency $\Omega_{tB}$ that is visible in lower plots. For both signs of the electric field, the resonant $\eta_{\rm res}$ is closer to the trapped passing boundary when shear is included. \noindent In Figs.~\ref{fig:do_transport}-\ref{fig:do_integral} the radial electric field dependence of non-ambipolar transport induced by drift-orbit resonances with magnetic drift neglected ($\Omega_{tB}$ set to zero) is pictured. Here, several canonical modes $m_2$ contribute for both, trapped and passing particles. \begin{figure} \includegraphics{fig2a_n3_A10}\quad{}\includegraphics{fig2b_n3_A10_D11D12} \protect\caption{Drift-orbit resonances with neglected magnetic drift: Mach number dependence of $D_{11}$ (left) and the ratio $D_{12}/D_{11}$ (right) for an RMP-like perturbation at $A=10$. Comparison of Hamiltonian approach ($\diamond$), sum of Hamiltonian results and $\nu-\sqrt{\nu}$ regime by Shaing~\cite{shaing10-025022} ($\square$), and results from NEO-2 at collisionality $\nu^\star=3\cdot10^{-4}$ (solid line). \label{fig:do_transport}} \end{figure} \noindent In Fig.~\ref{fig:do_transport} the Mach number dependence of transport coefficient $D_{11}$ and the ratio $D_{12}/D_{11}$ is plotted for this regime for an RMP-like perturbation ($n=3$). NEO-2 calculations shown for the comparison have been performed at rather low collisionality (see the caption) characterized by the parameter $\nu^\ast=2\nu q R/v_T$ where $\nu$ is the collision frequency. In addition, also the curves with the sum of diffusion coefficients in the collisional $\nu-\sqrt{\nu}$ regime from the joint formula of Shaing~\cite{shaing10-025022} and resonant contributions from the Hamiltonian approach are shown. \begin{figure} \includegraphics{fig3a_int_n3_A10_p}\quad{}\includegraphics{fig3b_int_n3_A10} \protect\caption{Drift-orbit resonances, RMP at $A=10$ with $M_t=0.028$. Dependence of the subintegrands in Eq.~\eq{eq:D1x} on the normalized velocity $u$ for the dominant mode (solid line) of passing (left, $m_{2}=-3$) and trapped particles (right, $m_{2}=-1$) and resonance lines for these modes ($\square$,~right axis). Significant contributions are visible where the resonance is close to the trapped passing boundary $\eta_{\rm tp} = 4.6\cdot 10^{-5}$. \label{fig:do_integral}} \end{figure} For $M_t<0.02$, in contrast to the superbanana plateau regime, collisionless transport is small compared to collisional effects. Between $M_t=0.02$ and $0.04$ the sum of Hamiltonian and $\nu-\sqrt{\nu}$ results for $D_{11}$ is clearly below NEO-2 values. The reason for this are contributions near the trapped passing boundary, which are illustrated in Fig.~\ref{fig:do_integral} at $M_t=0.028$. There the integrand in Eq.~\eq{eq:D1x} for the mode $m_2$ with the strongest contribution is shown together with the resonance line in velocity space. For $M_t>0.04$ there is a close match between the results with slightly lower $D_{11}$ values from NEO-2 due to remaining collisionality effects. \red{ \noindent It should be noted that validity of the ``collisionless'' Hamiltonian model cannot be accessed with the help of a simple Krook model although this model is fully adequate for the present derivations. The details of the collision model are not important as long as the collisional width of the resonant line in velocity space is smaller than the distance from that line to the trapped-passing boundary where the topology of the orbits changes abruptly. This criterion is much more restrictive than the smallness of the collision frequency compared to the bounce frequency suggested by the Krook model. At small Mach numbers where the resonant line approaches the trapped-passing boundary rather closely (see Fig.~\ref{fig:do_integral}), the applicability of the ``collisionless'' approach is violated at much lower collisionalities than one could expect from the Krook model, and in that case a collisional boundary layer analysis including the resonant interaction is needed. As one can see from Fig.~\ref{fig:do_transport}, for such transitional Mach numbers where both, $\nu-\sqrt{\nu}$ regime and resonant regime are important, a simple summation of the separate contributions from these regimes obtained in asymptotical limits cannot reproduce the numerical result, similarly to the observation in Ref.~\onlinecite{sun10-145002}. With increasing Mach numbers, the resonant curve gets more separated from the trapped-passing boundary, and the collisionless analysis becomes sufficient, as it can be seen for higher Mach numbers in Fig.~\ref{fig:do_transport}. } \begin{figure} \includegraphics{fig4a_n18_A10}\quad{}\includegraphics{fig4b_n18_A10_D11D12} \protect\caption{Mach number dependence of transport coefficients of drift-orbit resonances for a toroidal field ripple at $A=10$. Comparison between Hamiltonian approach ($\diamond$), ripple plateau (dashed) and NEO-2 at collisionality $\nu^\star=10^{-3}$ (solid line). \label{fig:do_transport18}} \end{figure} \noindent Fig.~\ref{fig:do_transport18} shows the Mach number dependence of $D_{11}$ as well as $D_{12}/D_{11}$ for a toroidal field ripple ($n=18$) together with the analytical ripple plateau value~\cite{Boozer1980-2283} and results for finite collisionality from NEO-2. At low Mach numbers $M_t<0.01$ collisional effects are again dominant. A resonance peak of passing particles is visible for $D_{12}/D_{11}$ at $M_t=2.8\cdot10^{-3}$. In the intermediate region between $M_t=0.01$ and $0.05$ oscillations due to trapped particle resonances are shifted and reduced in the collisional case. For $M_t>0.05$ Hamiltonian results converge towards the ripple plateau. A small deviation of NEO-2 values for $D_{11}$, which is of the order of Mach number is caused by the low Mach number approximation used in NEO-2. \begin{figure} \includegraphics{fig5a_n3_A5_shear_D11}\quad{}\includegraphics{fig5b_n3_A5_noshear_D11} \protect\caption{Mach number dependence of $D_{11}$ for RMP at $A=5$ with shear term included (left) and neglected (right) in Eq.~\eq{eq:OmtE}. Total resonant transport ($\diamond$) and contributions by drift-orbit resonances with finite magnetic drift and excluding superbanana plateau ($\square$). Comparison to drift-orbit resonances with $\Omega_{tB}$ set to zero (solid line). \label{fig:n3_A5}} \end{figure} \noindent Finally, in Fig.~\ref{fig:n3_A5} the Mach number dependence of $D_{11}$ for the RMP case is plotted for both, positive and negative Mach numbers for finite toroidal precession frequency due to the magnetic drift $\Omega_{tB}$. To set the scaling with respect to $\Omega_{tE}$, the reference magnetic drift frequency defined above is fixed by $R\Omega^{\rm ref}_{tB}/v_T = 3.6\cdot10^{-2}$. In this case all resonance types contribute to transport coefficients. Due to the finite magnetic drift, the Mach number dependence is not symmetric anymore. If shear is neglected in Eq.~\eq{eq:OmtE}, the superbanana plateau is centered around slightly negative values of the electric field, and magnetic drift induces some deviation from the idealized case without magnetic drift. In the case with included shear superbanana plateau, contributions for positive Mach numbers vanish and a large deviation from the case without magnetic drift is visible also for drift-orbit resonances. \section{Conclusion} \label{sec:conclosion} \noindent In this article, a method for the calculation of the toroidal torque in low-collisional resonant transport regimes due to non-axisymmetric perturbations in tokamaks based on a quasilinear Hamiltonian approach has been presented. This approach leads to a unified description of all those regimes including superbanana plateau and drift-orbit resonances without simplifications of the device geometry. Magnetic drift effects including non-local magnetic shear contributions are consistently taken into account. An efficient numerical treatment is possible by pre-computation of frequencies appearing in the resonance condition. \noindent The analytical expressions for the transport coefficients obtained within the Hamiltonian formalism agree with the corresponding expressions obtained earlier for particular resonant regimes within the validity domains of those results. In particular, the agreement with formulas for the superbanana plateau regime, which have been updated recently for a general tokamak geometry in Ref.~\onlinecite{shaing15-905810203}, is exact. Minor inconsistencies in the treatment of passing particles have been found (see section~\ref{sec:neotorque}) in comparison to the analytical formulas for bounce-transit resonances of Ref.~\onlinecite{shaing09-075015}. \red{ In addition, it has been demonstrated that momentum conservation of the collision operator plays a minor role in resonant regimes in general as long as the toroidal rotation is sub-sonic. } \noindent Results from the newly developed code NEO-RT based on the presented Hamiltonian approach agree well with the results from the NEO-2 code at relatively high Mach numbers where finite collisionality effects are small ($M_t>0.04$ in the examples here). At these Mach numbers, both approaches also reproduce the analytical result for the ripple plateau regime~\cite{Boozer1980-2283} well. At intermediate Mach numbers $0.02<M_t<0.04$ which correspond to the transition between the $\nu-\sqrt{\nu}$ regime and resonant diffusion regime, the combined torque of $\nu-\sqrt{\nu}$ regime and resonant diffusion regime does not reach the numerical values calculated by NEO-2 even at very low collisionalities due to the contribution of the resonant phase space region very close to the trapped-passing boundary. Collisional boundary layer analysis is required in addition to obtain more accurate results in these regions. \noindent Within the Hamiltonian approach, which is non-local by its nature, i.e. it does not use truncated ``local'' orbits which stay on magnetic flux surfaces, an additional term describing the influence of magnetic shear that is absent in the standard local neoclassical ansatz naturally arises in the resonance condition. This term significantly increases the asymmetry of the superbanana plateau resonance with respect to the toroidal Mach numbers of ${\bf E}\times{\bf B}$ rotation and may even eliminate this resonance for a given Mach number sign (at positive Mach numbers in the examples here). This shear term has been included into analytical treatment recently~\cite{shaing15-905810203} but was absent in earlier approximate formulas~\cite{shaing09-035009,shaing10-025022}. This could be a possible reason for the discrepancy with the non-local $\delta f$ Monte Carlo approach observed in Ref.~\onlinecite{satake11-055001}. \red{ \noindent It should be noted that term ``non-local transport ansatz'' is used here with respect to the orbits employed in the computation of the perturbation of the distribution function, and it should not be confused with the nonlocal transport in the case where the orbit width is comparable to the radial scale of the parameter profiles and where the transport equations cannot be reduced to partial differential equations. In the sense used here, the shear term appears due to a radial displacement of the guiding center, what is a nonlocal effect. Namely, due to variation of the safety factor with radius, the toroidal connection length between the banana tips of the trapped particle is different at the outer and the inner sides of the flux surface containing these tips. Since particles with positive and negative parallel (and, respectively, toroidal) velocity signs are displaced from this surface in different directions, the sum of the toroidal displacements over the full banana orbit is not balanced to zero, what results in an overall toroidal drift proportional to the shear parameter. This effect cannot be described by the local ansatz in an arbitrary coordinate system but still can be retained within the local ansatz in the field aligned coordinates as the ones used in Ref.~\onlinecite{shaing15-905810203}. This ambiguity in the description of the magnetic drift within the local ansatz~\cite{hokan} results from the fact that setting to zero one of the velocity vector components which are not invariant under a coordinate transformation destroys the covariance of equations of motion during such transformations. } \noindent Magnetic shear can also have a strong influence on drift-orbit (bounce and bounce-transit) resonances. A comparison between the results in this regime with neglected magnetic drift and results including magnetic drift shows a strong discrepancy, especially if magnetic shear is considered. Therefore, for an accurate evaluation of NTV torque in low-collisional resonant transport regimes it is necessary to consider magnetic drift including magnetic shear in the resonance condition. This is especially important for modern tokamaks with poloidal divertors where magnetic shear is high at the plasma edge where the main part of the NTV torque is produced. \noindent It should be noted that benchmarking with NEO-2 performed in this work resulted in improvement of the analytical quasilinear approach~\cite{kasilov14-092506} used in NEO-2 as well as in the numerical treatment. In particular, the use of compactly supported basis functions~\cite{kernbichler-ppcf} for the discretization of the energy dependence of the distribution function instead of global Laguerre polynomials used in earlier NEO-2 versions allowed to obtain correct results also at high Mach numbers with rather low collisionality where the global basis resulted in artificial oscillations of the diffusion coefficients with Mach number. In addition, the standard local neoclassical approach used in Ref.~\onlinecite{kasilov14-092506} for the derivation of quasilinear equations has been generalized to a non-local approach where the effect of magnetic shear is treated appropriately. Details of the derivation will be published in a separate paper. NEO-2 results for NTV in an ASDEX Upgrade equilibrium from both, local and non-local approach are shown and compared in Ref.~\red{\onlinecite{martitsch16-074007}}. \begin{acknowledgments} This work has been carried out within the framework of the EUROfusion Consortium and has received funding from the Euratom research and training programme 2014-2018 under grant agreement No 633053. The views and opinions expressed herein do not necessarily reflect those of the European Commission. The authors gratefully acknowledge support from NAWI Graz \red{and funding from the OeAD under the grant agreement ``Wissenschaftlich-Technische Zusammenarbeit mit der Ukraine'' No UA 06/2015}. \end{acknowledgments}
e8d284e6dc47b06087d2903bcd9efe320abdf8e3
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The problem of sampling complex free energy landscapes that exhibit long lived metastable states separated by large barriers is of great current interest\cite{longscale1,longscale2,longscale4}. The vast literature on the subject is a clear evidence of its pressing relevance\cite{meta_review}. Roughly speaking, two classes of methods can be identified. In one, a set of collective variables (CVs) that depend on the microscopic coordinates of the system is chosen, and a bias that depends on these chosen collective variables is constructed so as to speed up sampling. Examples of this approach are Umbrella Sampling\cite{met1}, Local Elevation\cite{met2}, Metadynamics\cite{meta_pnas,wtm} and more recently Variationally Enhanced Sampling\cite{omar2}. However, identifying the appropriate CVs can at times require a lengthy, if instructive, process\cite{gervasioCV}. The other set of methods can be classified under the generic name of tempering. The precursor of this approach is called Parallel Tempering\cite{ptempering} (PT). In PT $M$ replicas of the same system at different temperatures are run in parallel. Periodically a Monte Carlo test is made and, if the test is successful, configurations are exchanged between replicas. The rationale for this approach is that at high temperature it is easier for the system to move from basin to basin and this information is carried down to the colder temperature by the Monte Carlo exchange process. This idea has been generalized and several tempering schemes have been proposed\cite{re1,re2,re3,wte,re4,bussi_cvt,re6,re7,re8,re9,re10} in which the Hamiltonians in each replicas are progressively modified\cite{ptempering} in order to favor the process of barrier crossing. Statistics is then collected in the unmodified replica. Since it is relevant for what follows, we highlight the recent proposal made by Voth and collaborators\cite{Voth_sampling} and by our group\cite{quantumruge,dbms} to use artificially induced quantum effects to favor sampling. In this approach quantum effects are described by the isomorphism that is generated by the use of the Path Integral representation of Quantum Statistical Mechanics\cite{feynmanhibbs}. In this popular isomorphism each particle is mapped onto a polymer of $P$ beads interacting between themselves through a harmonic potential while the interaction between beads belonging to different polymers is appropriately reduced\cite{fcenter}. The success of these attempts have stimulated us to reduce this approach to its bare essentials and somehow to generalize it. In order to reduce its computational cost the number of beads will be reduced to two, thus we shall abandon any pretense of describing each replicas as a representation, albeit approximate, of a quantum state. Having given ourselves such a freedom we shall also change the interaction between the beads. Thus our tempering scheme will be composed of $M$ replicas; in each replica every particle becomes a dimer, and the interaction that holds together the dimers is now anharmonic. By tightening the potential that holds the dimers together we can progressively come close to the Boltzmann distribution of the system. As in the Feynman's isomorphism, this is reached when the intra-dimer interaction becomes infinitely strong. Like in Ref. \onlinecite{dbms} a key to the success of this method is the use of Metadynamics to increase the dimer length fluctuations. Thus we shall refer to this method as Dimer Metadynamics (DM). \section{Method} As discussed in the Introduction, our method is inspired by previous attempts to use quantum effects to overcome sampling bottlenecks\cite{quantumruge,dbms}. However here we prefer to derive our formulas without making explicit reference to the Path Integral representation of Quantum Statistical Mechanics\cite{feynmanhibbs}. We want to sample a Boltzmann distribution whose partition function is written as: \begin{eqnarray}\label{z1} Z_0 = \int dR \: \mbox{e}^{-\beta V(R)} \end{eqnarray} where $N$ particles of coordinates $R=\vec{r}_i, \: i=1...N$ interact via the $V(R)$ potential at the inverse temperature $\beta$. Let us now introduce $3N$ new coordinates $X = \lbrace \vec{x}_i \rbrace_{i=1}^{N}$ and a function $F_\sigma (X)$ that depends parametrically on $\sigma$ and it is such that $ \int dX \mbox{e}^{-\beta F_\sigma (X)}=1$. We can now rewrite $Z_0$ as \begin{eqnarray}\label{zrew} Z_0 = \int dRdX \: \mbox{e}^{-\beta F_\sigma (X)}\mbox{e}^{-\beta V(R)} \end{eqnarray} and making the coordinate transformation $X = R_1 - R_2$ and $ R = (R_1 + R_2)/2$ we obtain \begin{eqnarray}\label{zrew2} Z_0 = \int dR_1dR_2 \: \mbox{e}^{-\beta F_\sigma (R_1 - R_2)}\mbox{e}^{-\beta V(\frac{R_1 + R_2}{2})} \:. \end{eqnarray} Of course the partition function in Eq. \eqref{zrew2} is fully equivalent to the Boltzmann distribution and the doubling of the coordinates is only apparent. We now choose a family of functions $F_\sigma$ such that \begin{eqnarray}\label{dfam} \lim_{\sigma\rightarrow 0} \mbox{e}^{-\beta F_\sigma (R_1 - R_2)} = \delta (R_1 - R_2) \end{eqnarray} where $\delta(R_1-R_2)$ is the Dirac's delta. We shall postpone the choice of $F_\sigma$ to later, in the meantime we note that if Eq. \eqref{dfam} holds for $\sigma \rightarrow 0$, we can make the approximation $V(\frac{R_1+R_2}{2}) \simeq \frac{V(R_1)}{2}+\frac{V(R_2)}{2}$ and rewrite the partition function as: \begin{eqnarray}\label{zrew3} Z_\sigma = \int dR_1dR_2 \: \mbox{e}^{-\frac{1}{2}V(R_1)}\mbox{e}^{-\beta F_\sigma(R_1-R_2)}\mbox{e}^{-\frac{1}{2}V(R_2)} \end{eqnarray} that is fully equivalent to the Boltzmann distribution (Eq. \eqref{z1}) in the limit $\sigma \rightarrow 0$, however, contrary to Eq. \eqref{zrew2} here the degrees of freedom are doubled in earnest. This partition function describes a system of dimers bound by $F_\sigma (R_1 - R_2)$ and interacting via the reduced potential $\frac{V(R)}{2}$. At large values of $\sigma$ the dimers are loosely coupled and the two images of the system $R_1$ and $R_2$ can explore rather different configurations. By reducing $\sigma$ the coupling becomes stronger and stronger until for $\sigma\rightarrow 0$ the Boltzmann limit is approached. We now devise a set of replicas of which the first one is $Z_0$ whose $\sigma$ we denote by $\sigma_0$ and the others are of the type in Eq. \eqref{zrew3} with progressively large values of $\sigma$, starting with $\sigma=\sigma_1$. For reasons that will become apparent later we take $\sigma_0 = \sigma_1$. The idea is then to set up a replica exchange scheme in which systems of different $\sigma_i, \: i=1,...,M$ are run in parallel and periodically Monte Carlo attempts at swapping the configurations are made\cite{Voth_sampling,dbms}. In the tempering scheme proposed here the Monte Carlo test between neighboring replicas reads: \begin{eqnarray} p_{i,i+1}= \min \left[ 1, \mbox{e}^{-\beta\Delta E } \right] \label{pswp} \end{eqnarray} with: \begin{flalign} & \Delta E = \left[F_{\sigma_i}(R_1^{i+1} - R_2^{i+1})+F_{\sigma_{i+1}}(R_1^{i} - R_2^{i})\right] - \left[F_{\sigma_i}(R_1^{i} - R_2^{i})+F_{\sigma_{i+1}}(R_1^{i+1} - R_2^{i+1})\right] \end{flalign} where the subscript in the many-body coordinates is the bead index of the dimer and the superscript identifies the replica index. The Monte Carlo test between $Z_0$ and $Z_{\sigma_1}$ is given by: \begin{flalign} p_{0,1}=\min\left[1,\mbox{e}^{-\beta\Delta V^{(0,1)}}\right] \label{pacccg} \end{flalign} with: \begin{flalign} \Delta V^{(0,1)} = \left[ \frac{V(R_1^0)+V(R_2^0)}{2} + V\left(\frac{R_1^1+R_2^1}{2}\right)\right] - \left[ \frac{V(R_1^1)+V(R_2^1)}{2} + V\left(\frac{R_1^0+R_2^0}{2}\right)\right] \end{flalign} and we use the fact that we have chosen $\sigma_0 = \sigma_1$. We now turn to the choice of $F_\sigma (R_1 - R_2)$. We use the form \begin{eqnarray} F_\sigma (R_1 - R_2 ) = \sum_{i=1}^N f_\sigma (\vec{r}_i^{\,1} - \vec{r}_i^{\,2}) \end{eqnarray} where $\vec{r}_i^{\,1}$ and $\vec{r}_i^{\,2}$ are the coordinates of atom $i$ that has been split into the two beads $1$ and $2$ (see Eq. \eqref{zrew2}). For our approach to work it is necessary that for $\sigma \rightarrow 0$ the behavior of $\mbox{e}^{-f_\sigma (r)}$ is $\delta$-function like. One such class of functions can be obtained by considering for $0 < q \le 1$ the following representations of the three dimensional delta functions \begin{eqnarray} \delta (r) = \lim_{\sigma\rightarrow 0} \frac{\mbox{e}^{-\left[\left(1+\frac{r^2}{2q\sigma^2}\right)^q -1\right]}}{Z_{\sigma}^q} \end{eqnarray} where the normalization constant $Z_\sigma^q$ is given by \begin{eqnarray} Z_\sigma^q = 4\pi \int dr\: r^2 \mbox{e}^{-\left[\left(1+\frac{r^2}{2q\sigma^2}\right)^q -1\right]} \end{eqnarray} Using Eq. \eqref{dfam} and neglecting the immaterial constant $\log Z_\sigma^q$ we get for $f_\sigma^q(r)$ \begin{eqnarray} f_\sigma^q (r) = \left(1+\frac{r^2}{2q\sigma^2}\right)^q -1 \end{eqnarray} The choice of $f_\sigma^q (r)$ determines the potential with which two beads interact. For $q=1$ one has \begin{eqnarray}\label{fsgauss} f_\sigma^{(1)} (\vec{r}\,) = \frac{r^2}{2\sigma^2} \end{eqnarray} as in the standard Path Integral isomorphism. Otherwise for $0 < q < 1$ all $f_\sigma^q (r)$ exhibit a quadratic behavior $\frac{r^2}{2\sigma^2}$ at small $r$ and a slower growth at larger distances. In particular, for $q=\frac{1}{2}$ $f_\sigma^q (r)$ grows linearly with $r$. The transition between small-$r$ and large-$r$ regimes is controlled by the parameter $\sigma$, that in the spirit of the present work has no physical meaning and is only a tempering parameter. As $q \rightarrow 0$ the asymptotic behavior is even slower and in the limit becomes logarithmic. However in this case $Z_\sigma^q \rightarrow \infty$ and the system becomes unstable. In this first application of the method we choose $q=\frac{1}{2}$. This value is possibly not optimal but it is better than $q=1$. In our experimentation we found that $q=0.4$ and $q=0.3$ are also viable options. However the advantages did not seem so great as to warrant abandoning the more aesthetically pleasing $q=\frac{1}{2}$ choice. In this respect we find amusing to note that also the quark-quark interaction has a linear asymptotic behavior. The final and essential ingredient of our approach is the use of Metadynamics. Following Ref. \onlinecite{quantumruge} we shall combine the replica exchange scheme described above with Well-Tempered Metadynamics. In particular we shall use as CV the elastic energy per particle stored in the dimer \begin{eqnarray}\label{scv} s=\frac{1}{N \beta}\sum_{i=1}^{N}\left[\left(1+\frac{\left(\vec{r}_i^{\,2}-\vec{r}_i^{\,1}\right)^2}{2q\sigma^2}\right)^q-1\right] \end{eqnarray} The role of Metadynamics is to enhance the fluctuations of $s$ since in a Well-Tempered Metadynamics that uses $\gamma$ as boosting parameter the probability distribution of the biased variable, $p_b(s)$ is related to that in the unbiased ensemble by the relation $p_b(s) \propto \left[p(s)\right]^{1/\gamma}$. In our case since $s$ is related to the elastic energy, configurations in which the dimer is highly stretched are more likely to be observed. \section{Results} \begin{figure}[h] \includegraphics*[width=5.6in]{images/ala2_img.eps} \caption{Free energy surfaces in eV of Alanine Dipeptide as function of the dihedral angles defined in the right panel. The temperature was 300 K, the six replicas had $\sigma_i=$0.02, 0.02, 0.06, 0.15, 0.6 and 3.0 \AA. The simulation was 100 ns long with a timestep of $dt=1$ fs and swaps between configurations were attempted every 6 ps. Each 2 ps a Metadynamics Gaussian was deposited with initial height $w_0 =300$ K and bias factor $\gamma = 7$, the width of the Gaussians depended on the replica index and were $\sigma_g=$ 9.8, 9.8, 5.1, 4.7, 4.3 and 3.9 meV}\label{ala2pic} \end{figure} Before discussing the two applications presented here we want briefly to substantiate the assertion made earlier that the choice $q=\frac{1}{2}$ is more efficient than $q=1$. For this reason we consider once more the case of Alanine Dipeptide, a classical simple example on which new sampling methods are often tested. In \ref{ala2pic} we show the results obtained with the DM method with $q=\frac{1}{2}$ where only 6 replicas where required while in contrast, for $q=1$ we had to use 10 replicas. We now tackle a simple, yet challenging two dimensional system composed of 7 atoms interacting via Lennard-Jones (LJ) potential. This cluster is known to have three metastable configurations that can be represented as local minima in the free energy expressed as a function of the second and third momentum of the coordination number\cite{tribello} as shown in \ref{figljfes}. \begin{figure}[h] \includegraphics*[width=5.6in]{images/fes_lj7.eps} \caption{Free energy surface of a two dimensional LJ cluster of 7 atoms as function of the second and third momentum of the coordination number in Lennard Jones units obtained with (a) DM and (b) standard Metadynamics at a temperature $T=0.1 \epsilon$. In (c) the result obtained with DM at $\sigma_1=0.01$ without using replica exchange.}\label{figljfes} \end{figure} Standard MD is not able to sample the metastable states of this simple cluster in practical times and enhanced sampling methods have been used to study it\cite{tribello}. In \ref{figljfes} the free energy surface of this LJ cluster at $T = 0.1 $ has been computed with DM using 5 replicas with $\sigma_i = $0.01, 0.01, 0.05, 0.2 and 0.6, where $\sigma_0 = \sigma_1$ as discussed before and LJ units are used. The timestep was $dt = 0.01$ and the simulation ran for $10^8$ steps. Swaps between replicas were tried every 500 steps and the Metadynamics Gaussians were deposited every 200 steps with initial height of $0.1$ and width depending on the replica index, $\sigma_g =$ 5.7, 3.6, 2.1 and 1.4; the bias factor was $\gamma = 2.71$. These results are compared to a $10^8$ steps long Metadynamics\cite{wtm} simulation with timestep $dt = 0.005$ and bias factor $\gamma = 10$, where every $500$ steps the second and third momentum of the coordination number are biased with Gaussians of initial height $w_0= 0.01$, width $\sigma = (0.02 ; 0.02)$. As shown in \ref{geom} \begin{figure}[h] \includegraphics*[width=5.6in]{images/geom.eps} \caption{Probability distribution of the length of the dimer for a two dimensional Lennard-Jones cluster of 7 atoms obtained from an unbiased simulation (blue line), with Metadynamics on the spring energy (red line) and the unbiased distribution recovered from the biased simulation (green line). }\label{geom} \end{figure} DM dramatically enhances the sampling of the long distances tail of the probability distribution of the dimer length. Analogously to the Path Integral case\cite{quantumruge,dbms}, this effect increases the delocalization of the particle and indeed even without replica exchange DM can locate all of the four minima in the free energy surface (\ref{figljfes}(c)). We consider now the more complex case of Alanine Tripeptide in vacuum as described by the Charmm22$^\star$ \cite{charmm22s} forcefield, a protein that can have different conformations separated by moderately high energy barriers. \begin{figure}[h] \includegraphics*[width=5.6in]{images/ala3_phi1d.eps} \caption{Free energy as function of the dihedral angles $\phi_1$ (left) and $\phi_2$ (right). Black circles are results from the $\sigma=0.01$~\AA~ DM replica, blue squares are obtained from the Boltzmann DM replica and dashed line from VES. }\label{ala1d} \end{figure} These results were obtained with DM using 4 replicas with interaction strengths $\sigma =$ 0.01, 0.05, 0.08, and 0.15 \AA~ plus one additional Boltzmann replica (Eq. \eqref{zrew2}) with $\sigma_0=\sigma_1$. The temperature was $T = 300$ K and the simulation ran for 100 ns with a timestep of $dt = 2$ fs. Each 2 ps a Gaussian was deposited with initial height $w_0 = 300$ K and bias factor $\gamma = 4.3$, the standard deviations of the Gaussians depended on the replica index and were $\sigma_g = $ 33.5, 22.3, 13.4, 6.7 and 4.5 meV. Swaps between replicas were attempted every 5 ps. \begin{figure}[h] \includegraphics*[width=5.6in]{images/fes_ala3_2d.eps} \caption{Free energy surface in eV of Alanine Tripeptide obtained with (a) DM for the $\sigma_0 = 0.01$~\AA~ replica and (b) VES. For each of the minima a representative sample of the sampled configurations is shown. }\label{ala2d} \end{figure} We calculate a reference free energy surface by using the Variationally Enhanced Sampling (VES) of Ref. \onlinecite{omar2} in the Well-Tempered variant of Ref. \onlinecite{Valsson-JCTC-2015}. We employed as CVs the three dihedral angles $\Phi_{1},\Phi_{2},\Phi_{3}$ and expanded the bias potential in a Fourier series of size 6 for each CV, resulting a total number of 2196 basis functions. To optimize the VES functional we used the method of Bach\cite{Bach-NIPS-2013}. The coefficients were updated every 1 ps and a fixed step size of 0.08 kJ/mol was used. In order to achieve a Well-Tempered distribution we use a bias factor of 10 and the target distribution was updated every 500 ps\cite{Valsson-JCTC-2015}. The simulation cost was equivalent to that of a 50 ns Molecular Dynamics run. The results of these two calculations are compared in \ref{ala1d} in which the free energy as function of the dihedral angle $\phi_1$ and of $\phi_2$ are shown along with the definition of the angles. The agreement between the two calculations is excellent and we note that strictly speaking use of the rigorously Boltzmannian replica $Z_0$ is not necessary. Also the correlations between $\phi_1$ and $\phi_2$ (\ref{ala2d}) are well represented as well as the location and population of the different conformers. We underline the fact that in DM as opposed to VES no CVs need to be introduced. \section{Conclusions} Taking inspiration from de Broglie swapping Metadynamics\cite{dbms} and from Ref. \onlinecite{Voth_sampling}, we have used artificial delocalization effects to enhance sampling of Boltzmann systems. The delocalization has been obtained by mapping each particle into a dimer in which atoms are bound by an anharmonic potential. The computational cost relative to previous simulation methods is reduced. In fact here we deal only with dimers and not with polymers as in Ref. \onlinecite{dbms} and also the more gentle behavior of $f_\sigma^{\frac{1}{2}}(r)$ at large distances favors conformational swaps reducing the number of replicas needed. Like previous Path-Integral-based methods\cite{dbms,quantumruge} and PT, DM does not require choosing a CVs. Furthermore it offers a natural way of enhancing sampling of only a part of the system. For instance the conformational landscape of a mobile loop in a large protein could be selectively targeted. \begin{acknowledgement} All calculations were performed on the Brutus HPC cluster at ETH Zurich and on the Piz Dora supercomputer at the Swiss National Supercomputing Center (CSCS) under project ID u1. We acknowledge the European Union Grant ERC-2014-Adg-670227 and Marvel 51NF40\_141828. We would also like to acknowledge Omar Valsson for his help with VES on Alanine Tripeptide. \end{acknowledgement}
0ae8c8caf9db59e50f65c4f1f5438d7ecb5ed101
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section*{Acknowledgments} H.A. thanks the Department of Physics, Astronomy and Geosciences and Fisher College of Science and Math, Towson University, for travel support.
f16137edc700ad405623b4ae7c891dfb90c05c9f
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section*{Funding Information} The authors acknowledge support from Yale University. \maketitle \newcommand{\beginsupplement}{% \setcounter{table}{0} \renewcommand{\thetable}{S\arabic{table}}% \setcounter{figure}{0} \renewcommand{\thefigure}{S\arabic{figure}}% } \beginsupplement \section{Supplement 1} \subsection{Analytical model} The forward Brillouin gain and resonance frequency of air in a hollow-core fiber can be predicted given the properties of air and the geometry of the fiber. In particular, the forward-SBS gain can be calculated if the optical and acoustic eigenmodes can be determined for the hollow-core fiber geometry. The Brillouin gain $G^{kl}$ is proportional to a force, $\mathbf{f}^{kl}$, for a pump electric field mode with polarization state $k$ (defined as either $s$ or $p$) and a Stokes mode with polarization state $l$. Unlike in solids, pressure waves do not produce birefringence in typical gases (e.g. nonchiral) and therefore $f^{kl}=0$ when $k\neq l$. Using $\parallel$ to represent $k=l$ and $\perp$ to represent $k\neq l$, and following Refs. \cite{qiu2013stimulated,RakichOE2,Renninger2016}, the Brillouin gain for gases can therefore be written as: \begin{equation} G^{\parallel}= \frac{2\omega \,Q}{\Omega^2v_g^2}\frac{|\langle \mathbf{f}^{\parallel},\mathbf{u} \rangle|^2}{\langle \mathbf{E},\epsilon \mathbf{E}\rangle^2\langle \mathbf{u},\rho \mathbf{u}\rangle}\quad \text{and} \quad G^{\perp}=0. \label{bgaineq} \end{equation} $\langle \mathbf{A},\mathbf{B}\rangle$ is a vector inner product, defined as $\langle \mathbf{A},\mathbf{B}\rangle\equiv\int\mathbf{A}\cdot \mathbf{B}^* \,dxdy$, $\epsilon$ is the transverse dielectric distribution of the fiber, $\rho$ is the transverse mass density distribution, $\omega$ is the angular frequency of the Stokes wave, $\Omega$ is the acoustic frequency, $Q$ is the quality factor, $v_g$ is the optical group velocity, $\mathbf{E}$ is the time varying electric field of the optical mode, and $\mathbf{u}$ is the acoustic displacement vector for the acoustic mode. We assume a simple rigid hollow cylinder with the diameter of the HC-PBF. The equation for the displacement field, $\mathbf{u}$, given velocity, $v$, is represented as \begin{equation} \frac{\partial^2\mathbf{u}}{\partial t^2}=v^2\nabla\nabla\cdot\mathbf{u}. \label{eqdisp} \end{equation}\ Because the optical mode is radially symmetric, only the radially symmetric axial-radial acoustic modes will have nonzero coupling. Assuming that the displacement field oscillates at frequency $\Omega$, the relevant equation for the displacement field is then \begin{equation} r^2\frac{\partial^2 u_r}{\partial r^2}+r\frac{\partial u_r}{\partial r}+ \left(r^2\frac{\Omega^2}{v^2}-1\right)u_r=0, \label{radeqdisp} \end{equation} where we have neglected the small contribution from $u_z$. This is the equation for the Bessel function of the first kind $J_1$. Therefore, the relevant acoustic modes, $\mathbf{u}_i$, are given by $\mathbf{u}_i=u_o J_1(\frac{\Omega_i}{v}r)\hat{r}$, where $u_o$ is a normalization constant. The frequencies ($\Omega_i$) are determined by requiring the displacement to be zero at the core of the fiber ($\mathbf{u}=0$ when $r=R$). This boundary condition gives $\Omega_i=X_{1i}\frac{v}{R}$, where $X_{1i}$ is the $i$th zero of $J_1$. The coupling integrals are then calculated with this acoustic profile and assuming a Gaussian electric field profile with a mode-field diameter equal to the core radius. Assuming a dilute gas, the Brillouin gain can then be written as: \begin{equation} G^{\parallel}_i=\frac{\pi Q \xi_i n_g^2(n^2-1)^2}{\lambda X_{1i}^2 v^2 n^4 c \rho R^2}\quad \text{and} \quad G^{\perp}_i=0, \label{ideagasG2} \end{equation} where $n$ is the refractive index, $n_g$ is the group index, and $\xi_i$ is a nondimensional coupling constant defined as $\frac{R^4\left(\int\frac{\partial \mathbf{E}\cdot\mathbf{E}}{\partial r}\cdot \mathbf{u_i}dA_R\right)^2}{\int \mathbf{u_i}\cdot\mathbf{u_i}dA_R \left(\int\mathbf{E}\cdot\mathbf{E}dA\right)^2}$, where $dA_R$ represents an area integration over the core and the first five values are given by $\xi=\{1.085,.017,.014,.00832,.00548\}$. The gain for the first five resonances is plotted in Fig. 2 of the main text along with the optical field and force distributions. Calculations assume an inner diameter of $5.65$ $\mu$m, $Q=8.75$ (from the experimentally measured linewidth), $\lambda=1550$ nm, $n=1.0003$ for air, an air density of $1.2754$ kg/m$^3$, and an air pressure-wave speed of 343 m/s. \subsection{Experimental procedure and analysis} Experimentally, we exploit the broadband nature of forward Brillouin scattering in order to sensitively measure the frequencies and coupling strengths of the interaction \cite{Tang1,Fellegara1997,shin2013,Renninger2016}. Phase matching for the forward-SBS process requires the phase velocity of sound to equal the group velocity of light. Because the group velocity can have little variation over a large frequency window in fiber, two different wavelength lasers can interact coherently with the same acoustic resonances. This aspect facilitates the two-color experimental apparatus used in this work (Fig. \ref{fig1sch}). The two pump-fields (red), $\omega_1$ and $\omega_2$, are synthesized from a monochromatic laser ($\lambda=1535$ nm) using an intensity modulator. The modulation frequency is swept through the Brillouin-active resonances producing a resonant excitation of phonons in the HC-PBF. A continuous-wave probe beam (blue), $\omega_3$, is simultaneously injected into the fiber ($\lambda=1546$ nm) with a 50:50 beam splitter, to permit detection of the excited phonons. Through this process, a new frequency ($\omega_4$) is detected at the output of the fiber. The pump (red) tones are filtered out before detection. Finally, a reference arm derived from the probe is frequency shifted with an acousto-optic modulator (AOM) in order to uniquely measure the Stokes and anti-Stokes sidebands \cite{shin2013}. The input frequencies ($\omega_1-\omega_3$) are aligned to a common polarization with half-wave plates and a common polarizer. The new output frequency, $\omega_4$, is then polarization analyzed through control of the reference arm polarization to separately measure intra-polarization scattering at the common input polarization and inter-polarization scattering at the orthogonal direction. \begin{figure}[htb] \centerline{ \includegraphics[width=8.0cm]{SIfig1.pdf}} \caption{Experimental apparatus: IM, intensity modulator; HWP, half-wave plate; BS, 50:50 beamsplitter; HC fiber, hollow-core fiber; AOM, acousto-optic modulator; pol, polarizer; PC, polarization controller; IF, interference filter; and SA, spectrum analyzer. $\Omega_i$ is the acoustic resonance frequency.} \label{fig1sch} \end{figure} In accordance with analytical predictions, intra-polarization scattering measurements reveal a single 4-MHz linewidth peak at 35 MHz (Fig. 4 of the main text). A peak is not observed through measurements of inter-polarization scattering, which is consistent with Brillouin scattering in a gas. In addition, multiple narrower peaks corresponding to elastic modes in the silica matrix (see Ref. \cite{Renninger2016}) are observed at various frequencies in both polarizations. The electronic Kerr nonlinearity contributes a nonzero background at all frequencies, through a four-wave mixing process. The nonlinear Brillouin gain, $G$, can be determined with the known powers, fiber length, and coupling factor. When coupled with the AOM shifted reference signal, the measured heterodyne signal is given by $P_{signal}=2\eta\sqrt{P_{4}P_{AOM}}=L G \eta\sqrt{P_{1}P_{2}P_{3}P_{AOM}}$, where $P_{1}$ and $P_{2}$ are the two pump powers (red in Fig. \ref{fig1sch}), $P_{3}$ is the probe power (blue in Fig. \ref{fig1sch}), and $P_{AOM}$ is the optical power in the AOM arm; here $\eta$ is a unitless scale-factor that accounts for the experimental losses produced by numerous fiber components in the optical path between the fiber segment and the detector. This coefficient was obtained through careful calibration of this apparatus using reference signals of known powers. Given $P_{1}=P_{2}=48$ mW, $P_{3}=12.8$ mW, $P_{AOM}=0.34$ mW, $L=1.61$ m and $\eta=6.7\%$, we find that the peak power of the 35-MHz peak is given by 15 nW. The four-wave mixing gain coefficient can be calculated from the frequency independent background signal (\cite{shin2013}) and is measured here as $\gamma=1.6\times10^{-4}$ (Wm)$^{-1}$. The forward-SBS gain after subtracting this contribution from Kerr four-wave mixing is then given by $G=9\times10^{-4}$ (Wm)$^{-1}$. The Brillouin polarization, linewidth, frequency, and gain agree with theoretical predictions. \subsection{Spontaneous scattering and noise} The stimulated Brillouin coefficient can be used to estimate spontaneous forward Brillouin scattering, a similar process in which a single pump tone scatters from thermally excited and phase-matched phonons \cite{Kharel2016}. The total spontaneously generated optical power produced by the Brillouin active phonon mode is given by \begin{equation} P_{s} \cong P_{as} \equiv \int S(\Omega)d\Omega \cong G P_{pump} L\frac{\pi}{2}\frac{c}{\lambda}\frac{ k_B T}{Q}. \label{stimtospon} \end{equation} Here, $P_{s}$($P_{as}$) represents the total spontaneously scattered Stokes (anti-Stokes) power, integrated over the power spectral density ($S(\Omega)$), $G$ is the stimulated gain coefficient, $P_{pump}$ is the pump power, $L$ is waveguide length, $k_B$ is the Boltzman constant, and $T$ is temperature. For example for forward-SBS in air in the core of the hollow-core fiber, we can use Eq. \ref{stimtospon} to estimate the spontaneous optical scattering rate as follows: considering $G=9\times10^{-4}$ mW$^{-1}$ and $Q_i=8.75$, with $P_{pump}=100$ mW, $L=10$ m, $T=298$ K, and $\lambda=1550$ nm, the total spontaneously generated power scattered by this phonon mode is given by $P_{s}=P_{as}=130$ pW.
50e3e8f7b9ec2336153d8eb2e10a605adb73d67d
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction and statement of the conjecture} Let $p$ be a prime number, and let $N$ be a positive integer co-prime to $p$. The goal of this article is to investigate $U_p$-slopes:\ the $p$-adic valuations of the eigenvalues of the $U_p$-operator acting on spaces of (overconvergent $p$-adic) cuspforms of level $\Gamma_0(Np)$. Ultimately, we formulate a conjecture which unifies currently disparate predictions for the behavior of slopes at weights ``in the center'' and ``towards the boundary'' of $p$-adic weight space. \subsection{Slopes of cuspforms} The study of slopes of cuspforms began with extensive computer calculations of Gouv\^ea and Mazur in the 1990s \cite{GouveaMazur-FamiliesEigenforms}. Theoretical advancements of Coleman \cite{Coleman-pAdicBanachSpaces} led to a general theory of overconvergent $p$-adic cuspforms and eventually, with Mazur, to the construction of so-called eigencurves \cite{ColemanMazur-Eigencurve}. To better understand the geometry of the newly constructed eigencurves, Buzzard and his co-authors returned to explicit investigations on slopes in a series of papers \cite{Buzzard-SlopeQuestions,BuzzardCalegari-2adicSlopes,BuzzardCalegari-GouveaMazur,BuzzardKilford-2adc}. In \cite{Buzzard-SlopeQuestions}, Buzzard produced a combinatorial algorithm (``Buzzard's algorithm'') that for fixed $p$ and $N$ takes as input $k$ and outputs $\dim S_k(\Gamma_0(N))$-many integers. He also defined the notion of a prime $p$ being $\Gamma_0(N)$-regular and conjectured that his algorithm was computing slopes in the regular cases.\footnote{Buzzard's algorithm only outputs integers, so Buzzard's conjecture implies that $U_p$-slopes are always integral in $\Gamma_0(N)$-regular cases.} \begin{definition}[{\cite[Definition 1.2]{Buzzard-SlopeQuestions}}]\label{definition:Gamma0(N)-regular} An odd prime $p$ is $\Gamma_0(N)$-regular if the Hecke operator $T_p$ acts on $S_k(\Gamma_0(N))$ with $p$-adic unit eigenvalues for $2 \leq k \leq p+1$. \end{definition} See Definition \ref{defn:2regular} for $p=2$, but we note now that $p=2$ is $\SL_2(\Z)$-regular. The first prime $p$ which is not $\SL_2(\Z)$-regular is $p=59$. Buzzard's algorithm is concerned with spaces of cuspforms without character, where the slopes vary in a fairly complicated way. By contrast, a theorem of Buzzard and Kilford \cite{BuzzardKilford-2adc} implies that if $j \geq 3$ and $\chi$ is a primitive Dirichlet character of conductor $2^j$ then the $U_2$-slopes in $S_k(\Gamma_1(2^j),\chi)$ are the neatly ordered numbers $ 2^{3-j} \cdot \left(1,2,3,\dotsc,k-2\right)$. See also analogous theorems of Roe \cite{Roe-Slopes}, Kilford \cite{Kilford-5Slopes} and Kilford--McMurdy \cite{KilfordMcMurday-7adicslopes}. In \cite{LiuXiaoWan-IntegralEigencurves}, Liu, Wan and Xiao gave a conjectural, but general, framework in which to view the Buzzard--Kilford calculation (see \cite{WanXiaoZhang-Slopes} also). Namely, those authors have conjectured that the slopes of the $U_p$-operator acting on spaces of overconvergent $p$-adic cuspforms at $p$-adic weights ``near the boundary of weight space'' are finite unions of arithmetic progressions whose initial terms are the slopes in explicit classical weight two spaces. They also verified their conjecture for overconvergent forms on definite quaternion algebras. The beautiful description of the slopes at the boundary of weight space is actually a consequence (see \cite{BergdallPollack-FredholmSlopes,{LiuXiaoWan-IntegralEigencurves}}) of a conjecture, widely attributed to Coleman, called ``the spectral halo'':\ after deleting a closed subdisc of $p$-adic weight space, the Coleman--Mazur eigencurve becomes an infinite disjoint union of finite flat covers over the remaining portion of weight space. Furthermore, families of eigenforms over outer annuli of weight space should be interpreted as $p$-adic families passing through overconvergent $p$-adic eigenforms in characteristic $p$ (see \cite{AndreattaIovitaPilloni-Halo, JohanssonNewton-Extended}). The existence of a spectral halo should not depend on regularity. In summary, for a space either of the form $S_k(\Gamma_0(Np))$ or $S_k(\Gamma_0(N)\intersect \Gamma_1(p^r),\chi)$, the slopes are conjectured to be determined by a finite computation in small weights together with an algorithm:\ Buzzard's algorithm in the first case and ``generate an arithmetic progression'' in the second. In this article, we present a unifying conjecture that predicts the slopes of overconvergent $p$-adic eigenforms over all of $p$-adic weight space simultaneously. The shape of our conjecture is the following:\ we write down a power series over $\Z_p$ in two variables, one of which is the weight variable. We then conjecture, in the $\Gamma_0(N)$-regular case, that the Newton polygon of the specialization of our series to any given weight has the same set of slopes as the $U_p$-operator acting on the corresponding space of overconvergent $p$-adic cuspforms. \subsection{Fredholm series} Our approach begins with overconvergent $p$-adic modular forms. Write $\cal W$ for the {\em even} $p$-adic weight space:\ the space of continuous characters $\kappa: \Z_p^\times \rightarrow \C_p^\times$ with $\kappa(-1)=1$. For each $\kappa \in \cal W$ we write $S_\kappa^{\dagger}(\Gamma_0(Np))$ for the space of weight $\kappa$ overconvergent $p$-adic cuspforms of level $\Gamma_0(Np)$. An integer $k$ gives rise to a $p$-adic weight $z \mapsto z^k$, and the finite-dimensional space $S_k(\Gamma_0(Np))$ sits as a $U_p$-stable subspace of $S_k^{\dagger}(\Gamma_0(Np))$. In \cite{Coleman-ClassicalandOverconvergent}, Coleman proved that the $U_p$-slopes in $S_k(\Gamma_0(Np))$ are almost exactly those $U_p$-slopes in $S_k^{\dagger}(\Gamma_0(Np))$ which are at most $k-1$. Thus, one could determine the classical slopes by attempting the seemingly more difficult task of determining the overconvergent slopes. Denote by \begin{equation*} P_\kappa(t) = \det\left(1 - t\restrict{U_p}{S_{\kappa}^{\dagger}(\Gamma_0(Np))}\right) = 1 + \sum_{i \geq 1} a_i(\kappa)t^i \in \Q_p[[t]] \end{equation*} the Fredholm series for the $U_p$-operator in weight $\kappa$. The series $P_\kappa$ is entire in the variable $t$ and the $U_p$-slopes in weight $\kappa$ are the slopes of the segments of the Newton polygon of $P_\kappa$. Coleman's proved (see \cite[Appendix I]{Coleman-pAdicBanachSpaces}) that $\kappa \mapsto a_i(\kappa)$ is defined by a power series with $\Z_p$-coefficients. To be precise, we write $\cal W = \bigunion_{\varepsilon} \cal W_{\varepsilon}$ where the (disjoint) union runs over even characters $\varepsilon:(\Z/2p\Z)^\times \rightarrow \C_p^\times$, and $\kappa \in \cal W$ is in $\cal W_{\varepsilon}$ if and only if the restriction of $\kappa$ to the torsion subgroup in $\Z_p^\times$ is given by $\varepsilon$. We fix a topological generator $\gamma$ for the procyclic group $1+2p\Z_p$. Each $\cal W_{\varepsilon}$ is then an open $p$-adic unit disc with coordinate $w_{\kappa} = \kappa(\gamma)-1$. The meaning of Coleman's second result can now be clarified:\ for each $\varepsilon$ there exists a two variable series \begin{equation*} P^{(\varepsilon)}(w,t) = 1 + \sum_{i=1}^\infty a_i^{(\varepsilon)}(w)t^i \in \Z_p[[w,t]] \end{equation*} such that if $\kappa \in \cal W_{\varepsilon}$ then $P_{\kappa}(t) = P^{(\varepsilon)}(w_{\kappa},t)$. In particular, the slopes of overconvergent $p$-adic cuspforms are encoded in the Newton polygons of the evaluations of the $P^{(\varepsilon)}$ at $p$-adic weights. \subsection{The ghost conjecture}\label{subsec:ghost-conjecture} Our approach to predicting slopes is to create a faithful, explicit, model $G^{(\varepsilon)}$ for each Fredholm series $P^{(\varepsilon)}$. We begin by writing $G^{(\varepsilon)}(w,t) = 1 + \sum g_i^{(\varepsilon)}(w)t^i$ for coefficients $g_i^{(\varepsilon)}(w)$ which we shortly determine. If decorations are not needed, we refer to $g(w)$ as one of these coefficients. Each coefficient will be non-zero and not divisible by $p$.\footnote{In \cite{BergdallPollack-FredholmSlopes}, the authors showed that if $N=1$ then the coefficients $a_i^{(\varepsilon)}(w)$ are not divisible by $p$. For $N >1$ this is not true, but we don't believe this divisibility plays a crucial role for predicting slopes.} In particular, $w_{\kappa} \mapsto v_p(g(w_{\kappa}))$ will depend only on the relative position of $w$ to the finitely many roots of $g(w)$ in the open disc $v_p(w) > 0$. To motivate our specification of the zeros of $g_i^{(\varepsilon)}(w)$, we make two observations: \begin{enumerate} \item If $g_i^{(\varepsilon)}(w_\kappa) = 0$ then the $i$-th and $(i+1)$-st slope of the Newton polygon of $G^{(\varepsilon)}(w_\kappa,t)$ are the same. \end{enumerate} So, one can ask:\ what are the slopes that appear with multiplicity in spaces of overconvergent $p$-adic cuspforms? The second observation is: \begin{enumerate} \setcounter{enumi}{1} \item If $k\geq 2$ is an even integer then the slope ${k-2\over 2}$ is often repeated in $S_k^{\dagger}(\Gamma_0(Np))$. \end{enumerate} In fact, any eigenform in $S_k(\Gamma_0(Np))$ which is new at $p$ has slope ${k-2\over 2}$. So, in order to model the slopes of $U_p$ it might be reasonable to insist that $g_i^{(\varepsilon)}(w)$ has a zero exactly at $w = w_{k}$ with $k \in \cal W_{\varepsilon}$ where the $i$-th and $(i+1)$-st slope of $U_p$ acting on $S_k(\Gamma_0(Np))$ are both ${k-2\over 2}$. This leads us to seek $g_i^{(\varepsilon)}$ such that: \begin{equation}\label{eqn:mi_positive} g_i^{(\varepsilon)}(w_{k}) = 0 \iff \dim S_k(\Gamma_0(N)) < i < \dim S_k(\Gamma_0(N)) + \dim S_k(\Gamma_0(Np))^{p-\new} \end{equation} for $k \in \cal W_{\varepsilon}$. Such a $g_i^{(\varepsilon)}$ exists because for fixed $i$, the right-hand side of \eqref{eqn:mi_positive} holds for at most finitely many $k$. We now need to specify the multiplicities of the zeros $w_k$.\footnote{The na\"ive idea of having all zeros of $g_i^{(\varepsilon)}$ be simple would not work because the ghost series defined below would not be an entire series (compare with Proposition \ref{proposition:ghost-entire}).} An integer $k \in \cal W_{\varepsilon}$ is a zero for $g_i^{(\varepsilon)}(w)$ for some range of consecutive integers $i = a,a+1,\dots,b$ for which the right-hand side of \eqref{eqn:mi_positive} holds. Roughly, we set the order of vanishing of $g_a^{(\varepsilon)}(w)$ and $g_b^{(\varepsilon)}(w)$ at $w = w_{k}$ to be 1; for $g_{a+1}^{(\varepsilon)}(w)$ and $g_{b-1}^{(\varepsilon)}(w)$ to be 2; and so on. More formally, define the sequence $s(\ell)$ by \begin{equation*} s_i(\ell) = \begin{cases} i & \text{if $1\leq i \leq \floor{\ell / 2}$}\\ \ell+1 - i & \text{if $\floor{\ell / 2} < i \leq \ell$}, \end{cases} \end{equation*} and $s(\ell)$ is the empty sequence if $\ell\leq 0$. For $d \geq 0$ we write $s(\ell,d)$ for the infinite sequence \begin{equation*} s(\ell,d) = (\underlabel{d\text{ times}}{0,\dotsc,0},s_1(\ell),s_2(\ell),\dotsc,s_\ell(\ell),0,\dotsc). \end{equation*} If $k$ is an integer then set $d_k := \dim S_k(\Gamma_0(N))$ and $d_k^{\new} := \dim S_k(\Gamma_0(Np))^{p-\new}$. We then define $m(k) = s(d_k^{\new}-1,d_k)$, and set \begin{equation*} g_i^{(\varepsilon)}(w) := \prod_{k \in \cal W_{\varepsilon}} (w-w_{k})^{m_i(k)} \in \Z_p[w] \subset \Z_p[[w]] \end{equation*} which we note is a finite product. \begin{definition} The $p$-adic ghost series of tame level $\Gamma_0(N)$ on the component $\cal W_{\varepsilon}$ is \begin{equation*} G^{(\varepsilon)}(w,t) := 1 + \sum_{i=1}^\infty g_i^{(\varepsilon)}(w)t^i \in \Z_p[[w,t]]. \end{equation*} \end{definition} The naming choice and the motivation for the multiplicities defined in the next paragraph are discussed in Appendix \ref{app:explicit-2adic}. We check in Proposition \ref{proposition:ghost-entire} that $G^{(\varepsilon)}$ is entire as a power series in the variable $t$ over $\Z_p[[w]]$. In particular, for each $p$-adic weight $\kappa$ we get an entire series $G_{\kappa} \in \C_p[[t]]$. In what follows, we write $\oper{NP}(-)$ for ``Newton polygon''. \begin{conjecture}[The ghost conjecture]\label{conj:intro-ghost} If $p$ is an odd $\Gamma_0(N)$-regular prime or $p=2$ and $N = 1$, then $\oper{NP}(G_\kappa) = \oper{NP}(P_\kappa)$ for each $\kappa \in \cal W$. \end{conjecture} We check below that the hypotheses on $p$ is necessary (Theorem \ref{thm:ghost-true-regular}). In Section \ref{subsec:modification}, we formulate a conjecture when $p=2$ is $\Gamma_0(N)$-regular with a modified ghost series. \pagebreak \subsection{Evidence for the ghost conjecture} \subsubsection{Buzzard's conjecture versus the ghost conjecture} Buzzard's algorithm exploits many known and conjectured properties of slopes, such as their internal symmetries in classical subspaces, their (conjectural) local constancy in large families, and their interaction with Coleman's $\theta$-operator, to recursively predict classical $U_p$-slopes. The ghost conjecture on the other hand, simply motivated by the properties of slopes of $p$-newforms, predicts all overconvergent $U_p$-slopes and one obtains classical slopes by keeping the first $d_k$-many. These two approaches are completely different and, yet, they appear to agree. We view such agreement as compelling evidence for both conjectures. If $G(t) \in 1 + t\C_p[[t]]$ is a power series and $d \geq 1$, then write $G^{\leq d}$ for the truncation of $G$ in degree $\leq d$. Write $\oper{BA}(k)$ for the output of Buzzard's algorithm on input $k$. \begin{fact}\label{fact:buzzard-agreement} If either \begin{enumerate} \item $N=1$ and $p\leq 4099$ and $2\leq k\leq 2050$, or \item $2 \leq N \leq 42$, $3 \leq p \leq 199$ and $2 \leq k \leq 400$, \end{enumerate} then the multiset of slopes of $\oper{NP}(G_k^{\leq d_k})$ is equal to $\oper{BA}(k)$. \end{fact} We note that Buzzard made an extensive numerical verification of his conjecture which included all weights $k\leq 2048$ for $p=2$ and $N=1$. The careful reader will note a striking omission in the statement of Fact \ref{fact:buzzard-agreement}:\ the agreement between the ghost slopes and the output of Buzzard's algorithm does not seem to be limited to $\Gamma_0(N)$-regular cases. Namely, neither the construction of the ghost series nor Buzzard's algorithm requires any {\em a priori} regularity hypotheses and the tests we ran to check Fact \ref{fact:buzzard-agreement} were not limited to regular cases. It seems possible that someone with enough patience could even prove, without any hypothesis on $p$ and $N$, that the output of Buzzard's algorithm agrees with the classical ghost slopes. Although neither conjecture is predicting $U_p$-slopes in the irregular case, the numbers they both output could be thought of as representing the $U_p$-slopes that ``would have occurred" if not for the existence of a non-ordinary form of low weight. \subsubsection{Comparisons with known theorems on slopes} There are a number of cases where the slopes of $\oper{NP}(P_{\kappa})$ have been determined. In such cases that we know of, we independently verify that the ghost series determines the same list of slopes. \begin{theorem}[Theorem \ref{theorem:agree-BK-p=2}, Corollary \ref{corollary:bc-k=0}, Theorem \ref{theorem:consisten-boundary}]\label{theorem:intro-actual-truth} $\oper{NP}(G_{\kappa}) = \oper{NP}(P_{\kappa})$ in the following cases: \begin{enumerate} \item $p=2$, $N=1$, $\kappa=0$, \item $p=2$, $N=1$, $v_2(w_{\kappa}) < 3$, \item $p=3$, $N=1$, $v_3(w_{\kappa}) < 1$, \item $p=5$, $N=1$, $\kappa$ of the form $z^k\chi$ with $\chi$ conductor $25$ and \item $p=7$, $N=1$, $\kappa \in \mathcal W_0 \cup \mathcal W_2$ of the form $z^k\chi$ with $\chi$ conductor $49$. \end{enumerate} \end{theorem} The determination of the $U_p$-slopes in these cases are due to, in order, Buzzard and Calegari \cite{BuzzardCalegari-2adicSlopes}, Buzzard and Kilford \cite{BuzzardKilford-2adc}, Roe \cite{Roe-Slopes}, Kilford \cite{Kilford-5Slopes} and Kilford and McMurdy \cite{KilfordMcMurday-7adicslopes}. We also check the ghost conjecture is consistent with a conjecture of Buzzard and Calegari in \cite{BuzzardCalegari-2adicSlopes} on $2$-adic, tame level one, slopes at negative integers (Theorem \ref{theorem:agree-with-buzzard-calegari}) and we derive formulas for the slopes of $\oper{NP}(G_0)$ when $p=3,5$ and $N=1$ which agree with formulas found in Loeffler's paper \cite{Loeffer-SpectralExpansions} (Proposition \ref{prop:loeffler-sequences}). \subsubsection{The ghost spectral halo}\label{subsec:ghost-spectral-intro} Coleman's spectral halo, mentioned above, is concerned with $p$-adic weights quite far away from the integers. Specifically, let us refer to the spectral halo as the conjecture: \begin{conjecture}[The spectral halo conjecture]\label{conj:spectral-halo} There exists a $v > 0$ such that ${1\over v_p(w_\kappa)}\oper{NP}(P_\kappa)$ is independent of $\kappa \in \cal{W}_{\varepsilon}$ if $0 < v_p(w_\kappa) < v$. \end{conjecture} On $\cal W_{\varepsilon}$, the constant value of ${1\over v_p(w_\kappa)}\oper{NP}(P_\kappa)$ is then beautifully realized as the $w$-adic Newton polygon $\oper{NP}(\bar P)$ where $\bar P$ is the mod $p$ reduction of the $P^{(\varepsilon)}$. The ghost series trivially satisfies this halo-like behavior. Indeed, the zeros of each coefficient $g(w)$ lie in the region $v_p(w_\kappa) \geq 1$ (or $v_2(w_\kappa) \geq 3$ if $p=2$). Thus, over the complement of those regions, we have $v_p(g(w_{\kappa})) = \lambda(g)v_p(w_{\kappa})$ where $\lambda(g) = \deg g$. This proves: \begin{theorem}[The ghost spectral halo] \label{thm:gsh} The function $\kappa \mapsto {1\over v_p(w_\kappa)}\oper{NP}(G_\kappa)$ is independent of $\kappa \in \cal W_\varepsilon$ if $0 < v_p(w_\kappa) < 1$ (and $0 < v_2(w_{\kappa}) < 3$ if $p=2$), and the constant value is equal to $\oper{NP}(\bar{G}^{(\varepsilon)})$. \end{theorem} Along with the spectral halo conjecture, one also predicts that the slopes of $\oper{NP}(\bar P)$ are a finite union of arithmetic progressions for $v_p(w_{\kappa})$ small (see \cite[Conjecture 1.2(3)]{LiuXiaoWan-IntegralEigencurves}). We prove this directly for the ghost series, up to finite error. Write $\mu_0(N)$ for the index of $\Gamma_0(N)$ inside $\SL_2(\Z)$. \begin{theorem}[{Corollary \ref{cor:progs}, Theorem \ref{theorem:agree-BK-p=2}}]\label{theorem:intro-aps-boundary} If $p$ is odd then the slopes of $\oper{NP}(\bar G)$ are a finite union of ${p(p-1)(p+1)\mu_0(N)\over 24}$-many arithmetic progressions with a common difference ${(p-1)^2\over 2}$, except for finitely many possible exceptional slopes. If $p=2$ and $N=1$ then $\oper{NP}(\bar G)$ has slopes $\set{1,2,3,\dotsc}$. \end{theorem} We note that Theorem \ref{thm:gsh} and \cite{BergdallPollack-FredholmSlopes} imply that if the ghost conjecture is true then the exceptional slopes do not appear. More specifically, if the ghost conjecture is true then Theorem \ref{thm:gsh} implies the spectral halo exists on $0 < v_p(w_{\kappa}) < 1$, and if that is true then \cite[Theorem 3.10]{BergdallPollack-FredholmSlopes} proves that the slopes in Theorem \ref{theorem:intro-aps-boundary} are a finite union of arithmetic progressions without exceptions. Moreover, as evidence for the ghost conjecture, one can independently verify that the number of progressions predicted by \cite[Theorem 3.10]{BergdallPollack-FredholmSlopes} is exactly the same number written in Theorem \ref{theorem:intro-aps-boundary}.\footnote{If Conjecture \ref{conj:spectral-halo} is true with $v = 1$ then \cite[Theorem 3.10]{BergdallPollack-FredholmSlopes} predicts the number of progressions with common difference ${(p-1)^2\over 2}$ is given by \begin{equation*} {(p-1)c_0(N)\over 2} + \sum_{j=0}^{{p-3\over 2}} \dim S_2(\Gamma_0(N)\cap \Gamma_1(p^2), \chi\omega^{-2j}) \end{equation*} where $\chi$ is an even primitive character modulo $p^2$, $\omega$ is the Teichm\"uller character and $c_0(N)$ is the number of cusps on $X_0(N)$. One can check that this is exactly ${p(p-1)(p+1)\mu_0(N)\over 24}$ (using \cite[Th\'eor\`eme 1]{CohenOesterle-Dimensions} for example). } In addition to the ghost spectral halo, we've also discovered interesting arithmetic properties of slopes over {\em other} regions of $p$-adic weight space. See Section \ref{subsec:intro-global-halo} below (specifically Theorem \ref{theorem:global-halo}, which is a vast generalization of Theorem \ref{theorem:intro-aps-boundary}). \subsection{Distribution of ghost slopes} In Theorem \ref{thm:asymptotic} below we prove an asymptotic formula for the $i$-th slope of $\oper{NP}(G_k)$ when $k\geq 2$ is an even integer. Here we highlight two corollaries related to conjectures of Buzzard--Gouv\^ea and Gouv\^ea on the distribution of classical slopes. We write $s_1(k) \leq s_2(k) \leq \dotsb$ for the slopes of $\oper{NP}(G_k)$. On \cite[Page 8]{Gouvea-WhereSlopesAre}, Gouv\^ea asks if $v_p(a_p) \leq {k-1\over p+1}$ with probability one as $k\goto \infty$ where $a_p$ ranges over eigenvalues for $T_p$ acting on $S_k(\Gamma_0(N))$. Buzzard asks in \cite[Question 4.9]{Buzzard-SlopeQuestions} if the bound is {\em always} true when $p$ is $\Gamma_0(N)$-regular.\footnote{Gouv\^ea also asks whether or not $v_p(a_p) \leq {k-1\over p+1}$ for all $k$ once it is true for $k \leq p+1$, which is a slightly stronger question (\cite[Page 9]{Gouvea-WhereSlopesAre}).} We prove that the ghost slopes satisfy an asymptotic version of the Buzzard--Gouv\^ea bound. \begin{theorem}[Corollary \ref{cor:highest}]\label{theorem:buzzard-gouvea-intro} For $k\geq 2$ even, \begin{equation*} s_{d_k}(k) = {k \over p+1} + O(\log k). \end{equation*} \end{theorem} We believe that in fact $s_{d_k}(k) \leq {k-1\over p+1}$ always, but we did not pursue this except if $p=2$. We will not include details here. In \cite{Gouvea-WhereSlopesAre}, Gouv\^ea also considered, for a fixed $k$, the set \begin{equation*} \mathbf x_k := \set{ {h\over k-1} \st h \text{ is a slope of $T_p$ acting on $S_k(\Gamma_0(N))$}} \subset [0,1]. \end{equation*} He then conjectured that the sets $\mathbf x_k$ become equidistributed on $[0,{1\over p+1}]$ as $k\goto \infty$. We establish an analogous property for the ghost slopes. Write $d_{k,p} = \dim S_k(\Gamma_0(Np))$. \begin{theorem}[{Corollary \ref{corollary:gouvea-distribution}}] \label{thm:gouvea-dist} As $k\goto \infty$, the sets \begin{equation*} \set{ {s_i(k)\over k-1} \st 1\leq i \leq d_{k,p}} \end{equation*} become equidistributed with respect to the unique probability measure on $[0,{1\over p+1}] \union \set{1\over 2} \union [{p\over p+1},1]$ whose mass is ${p-1\over p+1}$ at ${1\over 2}$ and is uniformly distributed otherwise. \end{theorem} The method for these investigations is to study asymptotics of the actual points underlying the construction of the Newton polygons for the ghost series. The extra flexibility of having a power series in hand allows one to establish results like Theorem \ref{thm:gouvea-dist} without the annoying combinatorics that would underlie proving an exact Buzzard--Gouv\^ea bound holds. \begin{remark}\label{remark:gouvea-mazur-remark} We also explored the relationship between the ghost series and the Gouv\^ea--Mazur conjecture \cite[Conjecture 1]{GouveaMazur-FamiliesEigenforms}. Namely, one might ask if one sees ``logarithmic-sized ghost families'' as suggested by Buzzard's conjecture. Indeed, we do. In the discussion of the ghost spectral halo, we observed that all the zeros of the ghost coefficients occur at integer weights. Moreover, the set of zeros of a given coefficient is a linear function of its index. For example, if $k\geq 2$ is an integer then the zeros of the coefficients $g_1(w),\dotsc,g_{d_k}(w)$ (over the component containing $k$) are completely contained in the list $2,4,\dotsc,k-2$. In particular, if $v_p(w_\kappa - w_k) \geq 1 + \ceil{\log_p(k)}$ then $v_p(g_i^{(\varepsilon)}(w_\kappa)) = v_p(g_i^{(\varepsilon)}(w_k))$ for $1 \leq i \leq d_k$ and so \begin{equation*} \kappa \mapsto \oper{NP}(G_{\kappa}^{\leq d_k}) \end{equation*} is constant on $v_2(w_{\kappa} - w_{k}) \geq 1 + \ceil{\log_p(k)}$. For example, $S_{62}(\SL_2 \Z)$ is four-dimensional with $T_2$-slopes $6,6,14,14$ and Figure \ref{fig:gen-slopes-2} illustrates these are the lowest four ghost slopes on $v_2(w_\kappa - w_{62}) \geq 7 = 1 + \ceil{\log_2(62)}$. \end{remark} \subsection{Halos and arithmetic progressions}\label{subsec:intro-global-halo} We turn now towards one consequence of the ghost conjecture. For $\kappa \in \cal W$, let us write $\alpha_{\kappa} := \sup_{w \in \Z_p} v_p(w_{\kappa} - w)$. Since the zeros of the ghost coefficients are all integers, it is easy to see that if $\kappa,\kappa'$ lie on the same component and $v_p(w_{\kappa'}-w_{\kappa}) > \alpha_{\kappa}$, then $\oper{NP}(G_{\kappa'}) = \oper{NP}(G_{\kappa})$. In particular, if $w_\kappa \nin \Z_p$, then $\alpha_\kappa$ is finite and there is a small disc around $w_\kappa$ on which the ghost slopes are all constant. The simplest example is to fix $r \geq 0$ an integer and $v$ a rational number $r < v < r+1$. Then $\kappa \mapsto \oper{NP}(G_{\kappa})$ is constant on the disc $v_p(w_\kappa) = v$, and the Newton polygons scale linearly with $v$, forming ``halos''. We've illustrated the halos in Figure \ref{fig:gen-slopes} below where we've plotted the first twenty slopes on $v_p(w_{\kappa}) = v$ for $v \nin \Z$ when $p=2$ and $N = 1$. (The omitted regions are indicated with an open circle.\footnote{We stress that the behavior of the slopes in the omitted regions may be very complicated, interweaving the disjoint branches that we've drawn.}) Note the picture over $v_2(w_\kappa) < 3$ illustrates the result of Buzzard--Kilford \cite{BuzzardKilford-2adc}. Over $3 < v < 4$ you see pairs of parallel lines which hints at extra structure in the set of slopes. \begin{figure}[htbp] \begin{center} \includegraphics[scale=.7]{p=2N=1-21sheets-center0.pdf} \caption{``Halos'' for $p=2$ and $N=1$.} \label{fig:gen-slopes} \end{center} \end{figure} The following theorem explains this regularity. If $r \geq 0$, write \begin{equation*} C_{p,N,r} = {p^{r+1}(p-1)(p+1)\mu_0(N)\over 24}. \end{equation*} \begin{theorem}[Theorem \ref{theorem:global-halo-progressions}, Remark \ref{rmk:p=2-aps-remark}]\label{theorem:global-halo} Let $p$ be odd and assume $w_\kappa \nin \Z_p$. Write $r = \floor{\alpha_{\kappa}}$. Then the slopes of $\oper{NP}(G_{\kappa})$ form a finite union of $C_{p,N,r}$-many arithmetic progressions with the same common difference \begin{equation*} {(p-1)^2\over 2}\left(\alpha_\kappa + \sum_{v=1}^r (p-1)p^{r-v} \cdot v\right) \end{equation*} except for finitely many possibly exceptional slopes. If $p = 2$ and $N=1$ then the same is true with $C_{2,1,r} = \max(2^{r-2},1)$ and common difference $\alpha_k + \sum_{v=3}^r 2^{r-v}\cdot v$. \end{theorem} The condition $\floor{\alpha_\kappa} = 0$ is equivalent to $0 < v_p(w_\kappa) < 1$ in which case $\alpha_{\kappa} = v_p(w_\kappa)$. Thus, Theorem \ref{theorem:global-halo} generalizes Theorem \ref{theorem:intro-aps-boundary}. Note that Theorem \ref{theorem:global-halo} applies to $p$-adic annuli $r < v_p(w_{\kappa} - w_{k_0}) < r+1$ for any integer $k_0$, and $\kappa\mapsto \oper{NP}(G_\kappa)$ is constant on each fixed radius $v_p(w_\kappa - w_{k_0}) = v \in (r,r+1)$. Thus the halo behavior is stable under re-centering the coordinate $w$ at any integral weight. We illustrate this in Figure \ref{fig:gen-slopes-2} below, showing the halos near the weight $w=w_{62}$ when $p=2$ and $N=1$. (The interested reader may want to compare Figure \ref{fig:gen-slopes-2} to the discussion in the last paragraph of \cite[Section 3]{Buzzard-SlopeQuestions}.) \begin{figure}[htbp] \begin{center} \includegraphics[scale=.7]{p=2N=1-15sheets-center62.pdf} \caption{``Halos'' centered at the weight $w = w_{62}$ when $p=2$ and $N =1$.} \label{fig:gen-slopes-2} \end{center} \end{figure} There are several interesting observations regarding Figure \ref{fig:gen-slopes-2}. First, if $v_2(w_\kappa - w_{62}) > 3$ then $v_2(w_\kappa) = 3$, so the picture in Figure \ref{fig:gen-slopes-2} is nearly completely contained within the omitted regions in Figure \ref{fig:gen-slopes} over $v_2(w_\kappa) = 3$. Second, we've drawn some lines in Figure \ref{fig:gen-slopes-2} thicker than others:\ the thickness of a line corresponds to the multiplicity of a slope. On $v_2(w_\kappa-w_{62}) > 6$, we see a double slope 6; on $v_2(w_\kappa - w_{62}) > 7$, we see two 14s; and so on. Compare with the example at the end of Remark \ref{remark:gouvea-mazur-remark}. Next, the thickest line is six slope 30 families:\ these should correspond under the ghost conjecture to six families of $p$-adic eigenforms converging to the six newforms of weight $62$ (which have slope ${62-2\over 2} = 30$). Finally, the lone family at the top of Figure \ref{fig:gen-slopes-2} is a slope $61$ family which should be thought of under the ghost conjecture as converging to the critical slope Eisenstein series of weight 62. If the ghost conjecture is true, there are halos for $U_p$-slopes, and the slopes of $\oper{NP}(P_{\kappa})$ satisfy Theorem \ref{theorem:global-halo}. Over the annulus $0 < v_p(w_{\kappa}) < 1$, one can observe this empirically by computing classical spaces of cuspforms of weight with character of large $p$-power conductor. However, everything is much more mysterious over a $p$-adic annulus $r < v_p(w_{\kappa}) < r+1$ once $r \geq 1$:\ there are no locally algebraic weights in that region and thus no classical spaces of cuspforms. \subsection{Irregular cases} The basic heuristic in the ghost series construction is that the zeros of the coefficients of Fredholm series give rise to repeated slopes and that newforms provide many repeated slopes. In a separate article (\cite{BergdallPollack-FractionalSlopes}) we show that non-integral, and thus repeated, slopes always appear when $p$ is an odd $\Gamma_0(N)$-irregular prime. One could hope that careful predictions of where these fractional slopes appear could lead to a modification of the ghost series which would work in any case. We examined carefully the case where $p=59$ and $N=1$ and came up with a way to modify infinitely many, relatively sparse, coefficients by adding a new zero. We tested our modified ghost series against the $U_{59}$-slopes for weights $k \leq 1156$ and they matched perfectly. However, computing actual slopes is computationally difficult and we feel we do not have enough data to support an actual conjecture.\footnote{And, we cannot compare to Buzzard's algorithm since it doesn't compute slopes in irregular cases.} The precise indices where the zeros are added and the precise zeros which are added are determined by the list of slopes in weight two spaces with character of conductor $p=59$ (some of these are fractional). It would be interesting to have a modification which works for general $p$ and $N$ (after computing this finite amount of data). Moreover, such a modification would hopefully be regular enough and sparse enough so that the results of Sections \ref{subsec:distribution} and \ref{sec:global} will go through in the general case. \subsection{Organization} Section \ref{section:explicit-analysis} is concerned with explicitly determining information about the ghost series, including proving that it is entire in the variable $t$. However, the reader may want to skip directly to Section \ref{section:comparison} where we give more precise information when $p=2$ and $N=1$. This section also contains the bulk of the numerical evidence for the ghost conjecture. Sections \ref{subsec:distribution} deals with asymptotics of ghost slopes. It relies heavily on Section \ref{section:explicit-analysis}. The same is true for Section \ref{sec:global}, where we describe the halos and discuss the arithmetic properties of ghost slopes. Section \ref{sec:modification} contains a modification of the ghost series when $p=2$. The main theme is dealing with fractional slopes that appear in certain spaces. \subsection{Conventions}\label{subsec:notations} We maintain all the notations presented in the introduction. We also make the following conventions. If $P(t) = 1 + \sum a_i t^i \in \C_p[[t]]$ is an entire series then we write $\oper{NP}(P)$ for its Newton polygon, which is the lower convex hull of the set of points $\set{(i,v_p(a_i)) \st a_i\neq 0}$. The {\em slopes} of $P$ are the slopes of $\oper{NP}(P)$. The {\em $\Delta$-slopes} of $P$ are the differences $v_p(a_i) - v_p(a_{i-1})$ for $i=1,2,\dotsc$ with $a_{i},a_{i-1}\neq 0$, i.e.\ the slopes of the line segment connecting consecutive points before taking the Newton polygon. When $P$ is the Fredholm series for $U_p$ we will use $U_p$-slopes and when $P$ is the ghost series we will say ghost slopes and ghost $\Delta$-slopes. If $f(x), g(x)$ and $h(x)$ are real-valued functions of a variable $x = (x_1,\dotsc,x_n) \in \R^n$ then we write \begin{equation*} f(x) = g(x) + O(h(x)) \end{equation*} to mean that there exists an $M \geq 0$ and a constant $A > 0$ such that $\abs{f(x) - g(x)} \leq A\abs{h(x)}$ whenever $\norm{x} \geq M$ (where $\norm{-}$ is the standard norm on $\R^n$). If $h_1(x_1),\dotsc,h_n(x_n)$ are $n$ functions on a single variable then we write $f(x) = g(x) + O(h_1(x_1),\dotsc,h_n(x_n))$ to mean the above with $h(x) := \sup_{i} h(x_i)$. \subsection*{Acknowledgements} We thank Kevin Buzzard and Liang Xiao for helpful discussions. The first author was supported by NSF grant DMS-1402005 and the second author was supported by NSF grant DMS-1303302. \section{Explicit analysis of the ghost series}\label{section:explicit-analysis} We fix a prime $p$, a tame level $N$, and an even Dirichlet character $\varepsilon$ of conductor $p$. We write $g_i = g_i^{(\varepsilon)}$ for the coefficients of the $p$-adic ghost series $G(w,t)=G^{(\varepsilon)}(w,t)$ of tame level $N$ over the component $\cal W_{\varepsilon}$. We have two goals in this section. First, we will prove that $G(w,t)$ is entire over $\Z_p[[w]]$ (see Proposition \ref{proposition:ghost-entire}). Second, we will show that if the ghost conjecture is true then either $p$ is an odd $\Gamma_0(N)$-regular prime or $p=2$ and $N=1$ (see Theorem \ref{thm:ghost-true-regular}). We begin by recalling that if $k \in \cal W_{\varepsilon}$ then \begin{equation}\label{eqn:zeros} g_i(w_{k}) = 0 \iff d_k < i < d_k + d_k^{\new}. \end{equation} Throughout this section, we also refer to the integer $k$ as the zero of $g_i$ when we truthfully mean the coordinate $w_{k}$. \begin{lemma} \label{lemma:arithprogs} If $N>1$ or $p>3$, then the zeros of $g_i$ are integers $k$ which form a finite arithmetic progression with common difference $p-1$, if $p$ is odd, and $2$ if $p$ is even. \end{lemma} \begin{proof} By \eqref{eqn:zeros} it suffices to show that \begin{equation*} d_k < i < d_k + d_k^{\new} \implies \text{one of } \begin{cases} d_{k+\varphi(2p)} < i < d_{k+\varphi(2p)} + d_{k+\varphi(2p)}^{\new}\\ i \leq d_{k+\varphi(2p)}. \end{cases} \end{equation*} But if $N > 1$ or $p > 3$ then $d_k \leq d_{k + \varphi(2p)}$ and $d_{k}^{\new} \leq d_{k+\varphi(2p)}^{\new}$ (Lemma \ref{lemma:dim-formulas-increasing}). Thus if the first possibility fails, it must be due to the lower inequality, which is what we wanted to show. \end{proof} \begin{remark} Lemma \ref{lemma:arithprogs} only misses $N=1$ and $p=2,3$. When $N=1$ and $p=2$, the zeros of $g_i$ are $$ 6i+8,6i+10,\dots,12i-4,12i-2,12i+2, $$ and thus are just missing the single term $12i$ in an arithmetic progression. Similarly, for $N=1$ and $p=3$, the zeros of $g_i$ are $$ 4i+6,4i+8,\dots,12i-4,12i-2,12i+2. $$ (See Proposition \ref{proposition:explicit-p=2-info} and Table \ref{table:least-zeros}.) \end{remark} For each coefficient $g$, write $\HZ(g)$ (resp.\ $\LZ(g)$) for the highest (resp.\ lowest) $k$ such that $w_{k}$ is a zero of $g$. The following proposition describes these highest and lowest zeros up to constants bounded independent of $i$. \begin{proposition}\label{prop:approximate-zero-location} As functions of $i$, $$ \HZ(g_i) = \frac{12i}{\mu_0(N)} + O(1) ~\text{~~~and~~~}~ \LZ(g_i) = \frac{12i}{\mu_0(N)p} +O(1) $$ \end{proposition} \begin{proof} By standard dimension formulas (see Appendix \ref{app:dimension-formulae}), we have that $d_k = \frac{k \mu_0(N)}{12} +O(1)$ and $d_k^{\new} = \frac{k \mu_0(N)(p-1)}{12} + O(1)$. Thus, the largest $k$ satisfying $d_k < i$ equals $\frac{12i}{\mu_0(N)} +O(1)$, and the smallest $k$ satisfying $i < d_k + d_k^{\new}$ is $\frac{12i}{\mu_0(N)p} +O(1)$. The proposition follows from the definition \eqref{eqn:zeros}. \end{proof} We also explicitly describe how the zeros of the coefficients and their multiplicities change as we increase indices. Write $\Delta_i(w) = g_i(w)/g_{i-1}(w)$. The definition of the multiplicity patterns $m_i(-)$ in Section \ref{subsec:ghost-conjecture} implies that $\Delta_i$ has only simple zeros and poles at some finite set of $w = w_{k}$. More specifically, if $k \in \cal W_{\varepsilon}$ then \begin{equation}\label{eqn:Deltaizeros} \Delta_i(w_{k}) = 0 \iff m_i(k) = m_{i-1}(k) +1 \iff d_k+1 \leq i \leq d_k + \bfloor{{d_k^{\new}\over 2}} \end{equation} and \begin{equation}\label{eqn:Deltaipoles} \Delta_i(w_{k}) = \infty \iff m_i(k) = m_{i-1}(k) - 1 \iff d_k + \bfloor{{d_k^{\new}-1\over 2}} +2 \leq i \leq d_k + d_k^{\new}. \end{equation} For notation, we will always write $\Delta_i = \Delta_i^+/\Delta_i^-$ in lowest common terms. Thus $\Delta_i^{\pm} \in \Z[w]$ and the zeros are of the form $w_{k}$ with $k \in \Z$. We write $\HZ(\Delta_i^{\pm})$ and $\LZ(\Delta_i^{\pm})$ for the highest and lowest zeros as with $g_i$ above. \begin{proposition} \label{prop:change_in_zeroes} \leavevmode \begin{enumerate} \item \label{change:b} The zeros of $\Delta_i^-$ form an arithmetic progression with common difference $p-1$ if $p$ is odd and $2$ if $p=2$. The same is true for the zeros of $\Delta_i^+$ unless $N=1$ and $p=2, 3$. \item \label{change:c} $\displaystyle \HZ(\Delta_i^+) = \frac{12i}{\mu_0(N)}+O(1)$ and $\displaystyle \LZ(\Delta_i^+) = \frac{24i}{\mu_0(N)(p+1)} + O(1)$. \item \label{change:d} $\displaystyle \HZ(\Delta_i^-) = \frac{24i}{\mu_0(N)(p+1)}+O(1)$ and $\displaystyle \LZ(\Delta_i^-) = \frac{12i}{\mu_0(N)p} + O(1)$. \end{enumerate} \end{proposition} \begin{proof} Part (\ref{change:b}) follows from \eqref{eqn:Deltaizeros} and \eqref{eqn:Deltaipoles} together with Lemma \ref{lemma:dim-formulas-increasing} (as in the proof of Lemma \ref{lemma:arithprogs}). Parts (\ref{change:c}) and (\ref{change:d}) follow similarly as in the proof of Proposition \ref{prop:approximate-zero-location}. \end{proof} \begin{remark} In Proposition \ref{proposition:explicit-p=2-info} and Table \ref{table:least-zeros}, we give formulas making the above $O(1)$-terms precise when $N=1$ and $p=2,3,5$, and $7$. The qualification for $p=2,3$ and $N=1$ will be inconsequential as we move forward (see the proofs of Lemma \ref{lemma:zeropole_count} and Proposition \ref{prop:ith_newtonslope}, for example). \end{remark} For a non-zero element $\Delta \in \Z_p[[w]]$, we write $\lambda(\Delta)$ for the number of zeros of $\Delta$ in the open disc $v_p(w) >0$. We extend this to the field of fractions in the obvious way. \begin{lemma} \label{lemma:zeropole_count} If $p > 2$ then \begin{enumerate} \item $\lambda(\Delta_i^+) = \displaystyle \frac{12i}{\mu_0(N)(p+1)} + O(1)$ and \item $\lambda(\Delta_i^-) = \displaystyle \frac{12i}{\mu_0(N)p(p+1)} + O(1)$. \end{enumerate} For $p=2$, the same formulas hold if we replace the 12 by a 6. \end{lemma} \begin{proof} For $p>2$, the number of zeros of $\Delta_i^+$ equals $$ \frac{\HZ(\Delta_i^+) - \LZ(\Delta_i^+)}{p-1}+1 = \frac{1}{p-1} \left( \frac{12i}{\mu_0(N)} - \frac{24i}{\mu_0(N)(p+1)} \right) + O(1) = \frac{12i}{\mu_0(N)(p+1)} + O(1) $$ by Proposition \ref{prop:change_in_zeroes}(\ref{change:b},\ref{change:c}).\footnote{The $O(1)$ term absorbs the qualification that Proposition \ref{prop:change_in_zeroes} isn't quite true if $p=2,3$ and $N=1$.} The zeros of $\Delta_i^-$ and $p=2$ is done similarly. \end{proof} \begin{remark}\label{remark:lambdaDeltai} The difference $\lambda(\Delta_i) = \lambda(\Delta_i^+) - \lambda(\Delta_i^-)$ equals the $i$-th $\Delta$-slope of the mod $p$ reduction $\overline{G(w,t)}\in \F_p[[w,t]]$ (with the $w$-adic valuation on $\F_p[[w]]$). By Lemma \ref{lemma:zeropole_count}, the $i$-th $\Delta$-slope equals $ \frac{12i(p-1)}{\mu_0(N)p(p+1)}$ up to a bounded constant. We return to this in Section \ref{sec:global}. \end{remark} Recall that if $R$ is a local ring with maximal ideal $\ideal m$ and $F(t) = \sum r_i t^i \in R[[t]]$ then $F$ is called entire if there exists a sequence of integers $c_i$ such that $r_i \in \ideal m^{c_i}$ and $c_i / i \goto \infty$. If $R = \Z_p[[w]]$ and $G(w,t)\in \Z_p[[w,t]]$ is entire over $\Z_p[[w]]$ then the specialized series $G(w',t) \in \C_p[[t]]$ is entire in the usual sense for all $w' \in \C_p$ with $v_p(w') > 0$ (see \cite[Section 1.3]{ColemanMazur-Eigencurve}). \begin{proposition}\label{proposition:ghost-entire} The ghost series $G^{(\varepsilon)}(w,t)$ is entire series over $\Z_p[[w]]$. In particular, if $\kappa \in \cal W$ then $G_\kappa(t)$ is an entire series. \end{proposition} \begin{proof} Every root $w_{k}$ of $g_i$ lies in $p\Z_p$, and so $g_i \in (p,w)^{\lambda(g_i)}$. We claim $\lambda(g_i)/i \goto \infty$ as $i \goto \infty$. To show the claim, it is enough to show \begin{equation}\label{eqn:need-lim-inf} \liminf_i \left(\lambda(g_{i}) - \lambda(g_{i-1})\right) = \infty. \end{equation} But $\lambda(g_{i})-\lambda(g_{i-1}) = \lambda(\Delta_i)$, so \eqref{eqn:need-lim-inf} follows from Lemma \ref{lemma:zeropole_count} and the remark following it. \end{proof} We now turn to showing that for the ghost conjecture to be true, either $p=2$ and $N=1$ or $p$ is an odd $\Gamma_0(N)$-regular prime. In addition to our running notation $d_k$ and $d_k^{\new}$, we now also write $d_k^{\ord}$ for the dimension of the $p$-ordinary subspace of $S_k(\Gamma_0(Np))$. Hida theory implies that $d_k^{\ord}$ depends only on the component $\cal W_{\varepsilon}$ containing $k$. If $\kappa \in \cal W$, write $d_{G_\kappa}^{\ord}$ for the multiplicity of the slope zero in $\oper{NP}(G_\kappa)$. We leave the following proof to the reader. (The supremums are finite by Proposition \ref{prop:approximate-zero-location}.) \begin{lemma}[Ghost Hida theory]\label{lemma:ghost-hida} The function $\kappa \mapsto d_{G_\kappa}^{\ord}$ is constant on connected components of $\cal W$. Specifically, if $\kappa \in \cal W_{\varepsilon}$ then \begin{multline*} d_{G_\kappa}^{\ord} = \sup \set{i \st g_i^{(\varepsilon)}(w) = 1} = \sup \set{i \st m_i(k) = 0 \text{ for all $k \in \cal W_{\varepsilon}$}}\\ \geq \min \set{d_k \st k\geq 2 \text{ and } k \in \mathcal W_{\varepsilon}}. \end{multline*} \end{lemma} \begin{lemma}\label{lemma:explicit-ghost-ordinary-dims} Let $p \neq 2$. \begin{enumerate} \item If $4 \leq k \leq p-1$ is an even integer then $d_{G_k}^{\ord} \geq d_k$. \item $d_{G_2}^{\ord} \geq d_2 + d_2^{\new} = d_{2 + (p-1)}$. \end{enumerate} \end{lemma} \begin{proof} First assume that $4 \leq k \leq p-1$ (so $p>3$). By Lemma \ref{lemma:dim-formulas-increasing}, $n \mapsto d_{k+n(p-1)}$ is weakly increasing with respect to $n\geq 0$, so Lemma \ref{lemma:ghost-hida} proves $d_k = \min_n d_{k+n(p-1)}\leq d_{G_k}^{\ord}$. For part (b), Lemma \ref{lemma:weight2-inequality} implies that $d_2 + d_2^{\new} = d_{2+(p-1)}$. If $p = 3$ and $N=1$ then $d_{2+(p-1)} = \dim S_4(\SL_2 \Z) = 0$, so (b) is trivial. If $p > 3$ or $N > 1$ then Lemma \ref{lemma:dim-formulas-increasing} applies and $d_2+d_2^{\new} = d_{2 + (p-1)} = \min_{n\geq 1} d_{2 + n(p-1)}$. So, the coefficient $g_i$ at index $i = d_{2+(p-1)}$ is trivial, showing $d_{G_2}^{\ord} \geq d_{2 + (p-1)}$ by Lemma \ref{lemma:ghost-hida}. \end{proof} \begin{remark} When $p$ is odd, Lemma \ref{lemma:explicit-ghost-ordinary-dims}(b) implies that one could remove $w_{2}$ as a root of any of the coefficients of the ghost series. We actually do that in Section \ref{sec:global} below (see Lemma \ref{lemma:remove-wt2-zero}). \end{remark} \begin{lemma}\label{lemma:p=2-ghost-bound} If $p=2$ then $d_4\leq d_{G_2}^{\ord}$. \end{lemma} \begin{proof} By Lemma \ref{lemma:ghost-hida} it suffices to show that $g_{d_4} = 1$. Since $d_4 \leq d_{2m}$ for all $m\geq 2$ (Lemma \ref{lemma:dim-formulas-increasing} if $N>1$ and trivial if $N=1$), the only possible zero for $g_{d_4}$ is $w = w_2$. But by Lemma \ref{lemma:appendix-2adic-wt-2}, $d_2+d_2^{\new} \leq d_4$ and so the last index where $w_2$ is possibly a zero is strictly less than $d_4$. \end{proof} \begin{theorem}\label{thm:ghost-true-regular} Suppose the ghost conjecture is true. \begin{enumerate} \item If $p$ is odd then $p$ is $\Gamma_0(N)$-regular. \item If $p = 2$ then $N = 1$. \end{enumerate} \end{theorem} \begin{proof} Let $p$ be odd and assume the ghost conjecture is true. To show that $p$ is $\Gamma_0(N)$-regular we need to show that $d_k^{\ord} = d_k$ for $k=4,\dotsc,p+1$, and we have $d_k^{\ord} \leq d_k$ in general. Since we are assuming the ghost conjecture we have $d_{G_k}^{\ord} = d_k^{\ord}$ and thus Lemma \ref{lemma:explicit-ghost-ordinary-dims} implies $d_k \leq d_{G_k}^{\ord} = d_k^{\ord} \leq d_k$. Thus we get equality throughout, proving (a). Now let $p=2$, and assume the ghost conjecture is true. First suppose that $N \neq 1,3,7$ and we will get a contradiction. If $N\neq 1,3,7$ then Lemma \ref{lemma:appendix-2adic-wt-2} implies that $d_4 > d_2 + d_2^{\new} \geq d_4^{\ord}$ (the final inequality by Hida theory). But if the ghost conjecture is true then Lemma \ref{lemma:p=2-ghost-bound} implies $d_4^{\ord} = d_{G_2}^{\ord} \geq d_4$, which is a contradiction. To finish the theorem, we show in Example \ref{example:issue-p=2} below that the ghost conjecture is false when $p=2$ and $N=3,7$. \end{proof} \begin{example}\label{example:issue-p=2} Let $N = 3$. Then the $2$-adic ghost series begins \begin{equation*} G(w,t) = 1 + t + (w-w_{8})t^2 + (w-w_{8})(w-w_{10})t^3 + \dotsb \end{equation*} so if the the ghost conjecture is true then there is at least one ordinary form appearing in $S_4(\Gamma_0(3))$. This is absurd since $S_4(\Gamma_0(3))$ is a zero-dimensional vector space. Similarly, if $N = 7$ then the $2$-adic ghost series begins \begin{equation*} G(w,t) = 1 + t + (w-w_{4})t^2 + t^3 + \dotsb \end{equation*} and so the ghost conjecture would imply that there exists a least three ordinary forms appearing in $S_4(\Gamma_0(7))$, which is only a one-dimensional space. \end{example} \begin{remark} In Example \ref{example:issue-p=2}, the number of ordinary forms predicted by the ghost series doesn't even match the correct dimension of a weight four space. When $p=2$ and $N = 23$ the ghost conjecture is false, but for more subtle reasons:\ here the ghost series begins \begin{equation*} G(w,t) = 1 + t + t^2 + t^3 + t^4 + t^5 + (w-w_{6})t^6 + \dotsb \end{equation*} and there are no more trivial terms up to $t^{20}$ at least. One could even prove $d_{G_4}^{\ord}=5$ and in this case $S_4(\Gamma_0(23))$ happens to be five-dimensional. But, the slopes of $U_2$ acting on $S_4(\Gamma_0(23))$ are $\set{0,0,0,1,1}$ and so there are actually only three ordinary forms. \end{remark} \section{Comparison with known or conjectured lists of slopes}\label{section:comparison} This section is devoted to proving the ghost conjecture is true in every case mentioned in Theorem \ref{theorem:intro-actual-truth} (where the $U_p$-slopes have been previously determined). We do this by determining the ghost slopes in each case. We also prove that the ghost conjecture implies a conjecture of Buzzard and Calegari on slopes of overconvergent $2$-adic cuspforms, and we derive formulas for the ghost slopes at the weight $\kappa=0$ for $p=3,5$ and $N=1$. We focus first on $p=2$. So, until after the proof of Theorem \ref{theorem:agree-with-buzzard-calegari} below, we write $G(w,t) = 1 + \sum g_i(w)t^i \in \Z_2[[w,t]]$ for the $2$-adic tame level 1 ghost series. The reader may freely check the first four terms are: \begin{multline}\label{eqn:2-adic-ghost-two-terms} G(w,t) = 1 + (w-w_{14})t + (w-w_{20})(w-w_{22})(w-w_{26})t^2 + \\ (w-w_{26})(w-w_{28})(w-w_{30})(w-w_{32})(w-w_{34})(w-w_{38})t^3 + \dotsb \end{multline} Recall we write $\Delta_i = g_i/g_{i-1}$ and in lowest terms $\Delta_i = \Delta_i^+/\Delta_i^-$. \begin{proposition}\label{proposition:explicit-p=2-info} Let $p=2$ and $N=1$. \begin{enumerate} \item \label{explicit:a} $g_i(w_{k}) = 0$ if and only if $k$ is an even integer among $\set{6i + 8 , \dotsc 12i - 2} \union \set{12i+2}$. \item \label{explicit:b} If $i\geq 1$ then: \begin{enumerate}[(i)] \item The zeros of $\Delta_i^+$ are $w_{k}$ where $k=8i+4, \dotsc, 12i - 2, 12i + 2$ is even. \item The zeros of $\Delta_i^-$ are $w_{k}$ where $k = 6i + 2, \dotsc, 8i-2$ is even. \end{enumerate} \end{enumerate} \end{proposition} \begin{proof} We check (\ref{explicit:a}), leaving the remainder to the reader. First note that $d_2 = d_2^{\new} = 0$ so $w_2$ does not occur as a zero. Further, if $k\geq 4$ is an even integer then $d_{k+12} = d_k + 1$ and $d_{k+12}^{\new} = d_k^{\new} + 1$ (as follows easily from Appendix \ref{app:dimension-formulae}). By \eqref{eqn:zeros}, part (a) follows from: \begin{claim} If $i\geq 1$ then for all even $k\geq 4$, \begin{align} d_k + 1 \leq i &\iff k \leq 12i + 2 \text{ and $k\neq 12i$, and}\label{eqn:formula-to-induct-1}\\ i \leq d_k + d_k^{\new} - 1 &\iff 6i + 8 \leq k.\label{eqn:formula-to-induct-2} \end{align} \end{claim} To prove \eqref{eqn:formula-to-induct-1} and \eqref{eqn:formula-to-induct-2}, we work inductively. Namely, if the inequalities on either side of \eqref{eqn:formula-to-induct-1} are true for $(i,k)$ then they are also true $(i+1,k+12)$ and the if inequalities on either side of \eqref{eqn:formula-to-induct-2} are true for $(i,k)$ then they are also true for $(i+2,k+12)$. By induction on $i$, it is enough to prove the claim for $i=1,2$, which is done by examination of \eqref{eqn:2-adic-ghost-two-terms}. \end{proof} \begin{theorem}\label{theorem:agree-BK-p=2} Let $p=2$ and $N = 1$. \begin{enumerate} \item \label{BK2:a} If $i\geq 0$ then $\lambda(g_i) = {i + 1 \choose 2}$. \item If $v_2(w_{\kappa}) < 3$ then the slopes of $\oper{NP}(G_\kappa)$ are $\set{ j\cdot v_2(w_{\kappa}) \st j \geq 1}$ and $\oper{NP}(G_\kappa) = \oper{NP}(P_\kappa)$. \end{enumerate} \end{theorem} \begin{proof} We first prove part (a). The case of $i = 0$ is trivial since $g_0(w) = 1$. If $i\geq 1$ then Proposition \ref{proposition:explicit-p=2-info}(b) implies that \begin{align*} \lambda(g_i) &= \lambda(g_{i-1}) + \sizeof \set{\text{even } 8i+4,\dotsc,12i-2,12i+2} - \sizeof \set{\text{even } 6i + 2, \dotsc, 8i-2}\\ &= \lambda(g_{i-1}) + (2i - 1) - (i-1) = \lambda(g_{i-1}) + i. \end{align*} Thus, $\lambda(g_i) = {i+1\choose 2}$ by induction. It follows from the ghost spectral halo (Theorem \ref{thm:gsh}) and part (\ref{BK2:a}) that if $v_2(w_\kappa) < 3$ then $\oper{NP}(G_{\kappa})$ is equal to the lower convex hull of the set of points $(i, {i+1\choose 2}v_2(w_\kappa))$, whose slopes are easily checked to be $v_2(w_\kappa), 2v_2(w_\kappa),\dotsc$. This is precisely the list of $U_p$-slopes on $v_2(w_\kappa)<3$ computed by Buzzard and Kilford in \cite[Theorem B]{BuzzardKilford-2adc}. \end{proof} Let's now compare the $2$-adic ghost series with actual Fredholm series at negative even integers (following Buzzard and Calegari \cite{BuzzardCalegari-2adicSlopes}). \begin{theorem}\label{theorem:agree-with-buzzard-calegari} Let $p=2$ and $N=1$. If $k\leq 0$ is an even integer and $i\geq 1$ then \begin{equation*} v_2(g_i(w_{k})) = v_2\left(\prod_{j=1}^i 2^{2j}{(-k+12j+2)!(-k+6j)!\over (-k+8j+2)!(-k+8j-2)!(-k + 12j)}\right) \end{equation*} In particular, the slopes of $\oper{NP}(G_k)$ agree with the slopes predicted by Buzzard and Calegari in \cite[Conjecture 2]{BuzzardCalegari-2adicSlopes}. \end{theorem} Since Buzzard and Calegari proved their conjecture for $k=0$ (\cite[Theorem 1]{BuzzardCalegari-2adicSlopes}) we deduce: \begin{corollary}\label{corollary:bc-k=0} If $p=2$ and $N=1$ then $\oper{NP}(G_0) = \oper{NP}(P_0)$. \end{corollary} \begin{proof}[Proof of Theorem \ref{theorem:agree-with-buzzard-calegari}] Note that if $k\leq 0$ then $g_i(w_{k}) \neq 0$ for all $i\geq 0$ and thus $\Delta_i(w_{k})$ is well-defined. By induction on $i\geq 1$, it suffices to show that if $k\leq 0$ is an even integer then \begin{equation}\label{eqn:formula-2adic-ghost-negintegers} v_2\left(\Delta_i(w_{k})\right) = v_2\left(2^{2i}{(-k+12i+2)!(-k+6i)!\over (-k+8i+2)!(-k+8i-2)!(-k + 12i)}\right). \end{equation} To this end, Proposition \ref{proposition:explicit-p=2-info}(\ref{explicit:b}) implies \begin{align*} v_2\left(\Delta_i(w_{k})\right) &= v_2\left({(w_{k} - w_{8i+4}) \dotsb (w_{k}-w_{12 i - 2})(w_{k} - w_{12 i + 2}) \over (w_{k} - w_{6i+2}) \dotsb (w_{k} - w_{8i-2})}\right). \end{align*} The $\dotsb$ indicate running over only even integers. Since $v_2(w_{k} - w_{k'}) = 2 + v_2(k - k')$, \begin{align*} v_2\left(\Delta_i(w_{k})\right) &= v_2\left(2^{2i} {(k-(8i+4))\dotsb (k-(12i-2))(k-(12i+2))\over (k-(6i+2))\dotsb (k-(8i-2))}\right) \\ &= v_2\left(2^{2i}{(-k+12i+2)!(-k+6i)!\over (-k+8i+2)!(-k+8i-2)!(-k + 12i)}\right).\end{align*} as desired. \end{proof} We now release our restriction to $p=2$ and $N=1$. Analogs of Proposition \ref{proposition:explicit-p=2-info} may be carried out for other values of $p$ and $N$. In Table \ref{table:least-zeros} below, we list the outcome for $p=3,5,7$ and tame level $N = 1$, on the weight component corresponding to $k\congruent 0 \bmod p-1$. With the details from Table \ref{table:least-zeros} available, it is easy to compute the ghost $w$-adic $\Delta$-slopes (see Table \ref{table:wadic-slopes} --- we've added a few more cases there as well). \begin{table}[htpp] \caption{Explicit determination of zeros of $\Delta_i^{\pm}$ for the $p$-adic tame level 1 ghost series $G(w,t) = 1 + \sum g_i(w)t^i$ on the component of weights $k\congruent 0 \bmod p-1$ for $p=3,5,7$.} \begin{center} \renewcommand{\arraystretch}{1.3} \begin{tabular}{|c|c|c|c|} \hline $p$ & $3$ & $5$ & $7$\\ \hline $\HZ(\Delta_i^+)$ & $12i+2$ & $12i-4$ & $12i-6$\\ \hline $\LZ(\Delta_i^+)$ & $6i+4$ & $4i+4$ & $6\floor{{i\over 2}}$\\ \hline $\HZ(\Delta_i^-)$ & $6i-2$ & $4i-4$ & $6\floor{{(i-1)\over 2}}$\\ \hline $\LZ(\Delta_i^-)$ & $ 4i+2$ & $4 \floor{\frac{3i}{5}} + 4$ & $ 6 \floor{\frac{2i}{7}}+6$\\ \hline \end{tabular} \end{center} \label{table:least-zeros} \end{table} \begin{table}[htpp] \caption{Differences of consecutive $\lambda$-invariants.} \begin{center} \renewcommand{\arraystretch}{1.3} \begin{tabular}{|c|c|c|c|c|c|c|} \hline $p$ & $3$ & $5$ & $5$ & $7$ & $7$ & $7$ \\ \hline Weight component & $0\bmod 2$ & $0\bmod 4$ & $2 \bmod 4$ & $0 \bmod 6$ & $2\bmod 6$ & $4\bmod 6$\\ \hline $\lambda(g_i)-\lambda(g_{i-1})$ & $2i$ & $\floor{{8i\over 5}}$ & $\floor{{(8i+4)\over 4}}$ & $\floor{{9i\over 7}}$ & $\floor{{(9i+6)\over 7}}$ & $\floor{{(9i+3)\over 7}}$\\ \hline \end{tabular} \end{center} \label{table:wadic-slopes} \end{table} \begin{theorem}\label{theorem:consisten-boundary} Suppose that $N=1$ and that $G(w,t) = 1 + \sum g_i(w)t^i$ is the ghost series on a component to be determined. Then, $\oper{NP}(G_\kappa) = \oper{NP}(P_\kappa)$ if: \begin{enumerate} \item If $p = 3$ and $v_p(w_\kappa) < 1$. \item If $p=5$ and $\kappa = z^k \chi$ where $\chi$ is a primitive modulo 25 and $\chi(-1) = (-1)^k$. \item If $p = 7$ and $\kappa = z^k\chi \in \cal W_0\union \cal W_2$ and $\chi$ is primitive modulo 49. \end{enumerate} \end{theorem} \begin{proof} The ghost $w$-adic $\Delta$-slopes in Table \ref{table:wadic-slopes} are always weakly increasing. So, for $p=3,5,7$ and $\kappa \in \cal W$ with $v_p(w_\kappa) < 1$, Theorem \ref{thm:gsh} implies that the slopes on $\oper{NP}(G_{\kappa})$ are given by $\set{(\lambda(g_i)-\lambda(g_{i-1}))\cdot v_p(w_\kappa) \st i = 1,2,\dotsc}$. The proof is then complete from Table \ref{table:wadic-slopes} once we verify these are the slopes of $\oper{NP}(P_{\kappa})$ in cases (a), (b) and (c). The case (a) is the main result of Roe's paper \cite{Roe-Slopes}. The case (b) is due to Kilford \cite{Kilford-5Slopes}. The case (c) was computed by Kilford and McMurdy in \cite{KilfordMcMurday-7adicslopes}.\footnote{In comparing our statement to \cite{Kilford-5Slopes,KilfordMcMurday-7adicslopes}, one is forced to unwind the various choices made in those papers regarding embeddings of cyclotomic fields into $\C_p$. Doing it carefully, one sees that \cite{KilfordMcMurday-7adicslopes} does not contain any result regarding the component of weights $k\congruent 4 \bmod 6$ when $p=7$.} \end{proof} One may also generalize Theorem \ref{theorem:agree-with-buzzard-calegari}. In Table \ref{table:bc-type-expressions} below, for $p=3$ and $p=5$, we give expressions that allows us to compute the Newton polygon of the ghost series at negative even integers as the Newton polygon of a series whose coefficients are rational functions involving simple factorials when $p=3$ and $p=5$. \begin{table}[htpp] \caption{Buzzard--Calegari-type expressions for $\oper{NP}(G_k)$ at negative integers $k \congruent 0 \bmod p-1$ when $p=3$ and $p=5$.} \begin{center} {\renewcommand{\arraystretch}{2} \begin{tabular}{|c|c|} \hline $p$ & $Q_j(k)$ such that $v_p(g_i(w_{k})) = v_2\left(\prod_{j=1}^i Q_j(k)\right)$\\ \hline $3$ & $\displaystyle {3^{2j}(-k/2 + 6j + 1)!(-k/2 + 2j)!\over (-k/2 + 3j+1)!(-k/2+3j-1)!(-k/2+6j)}$\\ $5$ & $\displaystyle {5^{\floor{8j/5}} (-k/4 + 3j-1)!(-k/4+\floor{3j/5} )!\over (-k/4+j)!(-k/4+j-1)!}$\\ \hline \end{tabular} } \end{center} \label{table:bc-type-expressions} \end{table} From Table \ref{table:bc-type-expressions} we can determine the slopes of the ghost series at $\kappa = 0$ for $p=3,5$. The expressions we derive agree with those conjectured in Loeffler's paper \cite[Conjecture 3.1]{Loeffer-SpectralExpansions}. \begin{proposition}\label{prop:loeffler-sequences} The Newton polygon $\oper{NP}(G_0)$ for $p=3,5$ has slopes \begin{equation*} \begin{cases} 2i + 2v_3\left({(2i)!\over i!}\right) & \text{if $p=3$;}\\ i + 2 v_5\left({(3i)!\over i!}\right) & \text{if $p=5$.} \end{cases} \end{equation*} \end{proposition} \begin{proof} The sequences given are increasing with respect to $i$. If we show they agree with the $\Delta$-slopes of $G_0$ then we will be done. The proof is similar in either case, so we just deal with the case $p=5$. By Table \ref{table:bc-type-expressions} we have \begin{equation}\label{eqn:p=5loeffler-1} v_2(\Delta_i(w_{0})) = v_5\left({5^{\floor{8i/5}} (3i-1)!(\floor{3i/5} )!\over (i)!(i-1)!}\right) = i + v_5\left({5^{\floor{3i/5}} (3i-1)!(\floor{3i/5} )!\over (i)!(i-1)!}\right) \end{equation} But for any integer $n\geq 1$ and prime $p$ we have $v_p(\floor{n/p}!) = v_p(n!) - \floor{n/p}$. Thus \begin{equation}\label{eqn:p=5loeffler-2} v_5\left({5^{\floor{3i/5}} (3i-1)!(\floor{3i/5} )!\over (i)!(i-1)!}\right) = v_5\left({ (3i-1)!(3i)!\over (i)!(i-1)!}\right) = v_5\left({(3i)!^2\over (i)!^2}\right). \end{equation} Combining \eqref{eqn:p=5loeffler-1} and \eqref{eqn:p=5loeffler-2}, we deduce our claim. \end{proof} \begin{remark} For $p=7$ and $N=1$ the $i$-th slope of $\oper{NP}(G_0)$ is \begin{equation*} i + v_7\left({(2i)!(2i-1)!\over \floor{i/2}!\floor{(i-1)/2}!}\right) = i + 2v_7\left((2i)!\over \floor{(i-1)/2}!\right) - v_7(i) - \begin{cases} 1 & \text{if $i \congruent 0 \bmod 14$}\\ 0 & \text{otherwise.} \end{cases} \end{equation*} (compare with the comments of Loeffler in the final paragraph prior to Section 4 of \cite{Loeffer-SpectralExpansions}). \end{remark} \section{Distributions of slopes} \label{subsec:distribution} For a fixed integer $k$ we write $s_1(k) \leq s_2(k) \leq \dotsb$ for the slopes of $\oper{NP}(G_k)$. Recall our conventions for $O$-notation (Section \ref{subsec:notations}). Throughout this section, functions of $i$ and $k$ are restricted to $i\geq 1$ and $k\geq 2$. The main theorem of this section is:\ \begin{theorem} \label{thm:asymptotic} $$ s_i(k) = \begin{cases} \displaystyle \frac{12i}{\mu_0(N)(p+1)} + O(\log(k), \log(i)) & \text{if~}i\leq d_k \text{~or~}i > d_k +d_k^{\new}, \\ ~\\ \displaystyle \frac{k}{2} + O(\log(k)) & d_k < i \leq d_k + d_k^{\new}. \end{cases} $$ \end{theorem} Before beginning the proof of Theorem \ref{thm:asymptotic}, we state two corollaries (Theorems \ref{theorem:buzzard-gouvea-intro} and \ref{thm:gouvea-dist} from the introduction). \begin{corollary} \label{cor:highest} $ \displaystyle s_{d_k}(k)= \frac{k}{p+1} + O(\log(k)). $ \end{corollary} \begin{proof} Note that $\displaystyle d_k = {k\mu_0(N)\over 12} + O(1)$ and take $i=d_k$ in Theorem \ref{thm:asymptotic}. \end{proof} Recall that $d_{k,p} := \dim S_k(\Gamma_0(Np))$. Then, consider the set $$ \mathbf x_k = \left\{ \frac{s_i(k)}{k-1} \st 1\leq i \leq d_{k,p} \right\} \subseteq [0,\infty). $$ Let $\mu^{(p)}_k$ be the probability measure on $[0,\infty)$ uniformly supported on $\mathbf x_k$. We refer to \cite[Sections 1.1--1.2]{Serre-Equidistribution} for the notion of weak convergence and its relationship to equidistribution. \begin{corollary}\label{corollary:gouvea-distribution} As $k\goto \infty$, the measures $\mu^{(p)}_k$ weakly converge to a probability measure $\mu^{(p)}$ on $[0,1]$ which is supported on $[0,\frac{1}{p+1}] \cup \{\frac{1}{2}\} \cup [\frac{p}{p+1},1]$. Explicitly, $\mu^{(p)}(\set{1\over 2}) = {p-1\over p+1}$ and the remaining mass is uniformly distributed over $[0,\frac{1}{p+1}] \cup [\frac{p}{p+1},1]$. \end{corollary} \begin{proof} This is clear from Theorem \ref{thm:asymptotic} and the asymptotics for $d_k, d_k^{\new}$ and $d_{k,p}$ (for example, see the proof of Proposition \ref{prop:approximate-zero-location} for $d_k$ and $d_k^{\new}$; an asymptotic for $d_{k,p}$ is easily obtained from those two). \end{proof} \begin{remark} The measures $\mu_k^{(p)}$ clearly depend on $N$, even if $N$ is suppressed from our notation. However, it is interesting that the limit $\mu^{(p)}$ does not. \end{remark} \begin{remark} The key point in the proof of Theorem \ref{thm:asymptotic} is the analysis in Proposition \ref{prop:ith_newtonslope} below. If one could prove an analog of Proposition \ref{prop:ith_newtonslope} for the Fredholm series of $U_p$ then the proof of Theorem \ref{thm:asymptotic} would go through as written. \end{remark} The rest of this section is devoted to proving Theorem \ref{thm:asymptotic}. Our strategy is to prove an analoge of Theorem \ref{thm:asymptotic} for ghost $\Delta$-slopes first and, from this, make conclusions about ghost slopes. We will need two short lemmas. \begin{lemma}\label{lemma:valuations} Suppose that $y,\lambda > 0$ are integers and $p$ is a prime number. Then \begin{equation*} v_p(\lambda!) \leq \sum_{i=0}^{\lambda-1} v_p(y+i) \leq v_p((\lambda-1)!) + \floor{\log_p(y+\lambda)} + \min(v_p(y),v_p(\lambda)). \end{equation*} \end{lemma} \begin{proof} Write \begin{equation*} s(y) = \sum_{i=0}^{\lambda-1} v_p(y+i) = v_p\left({y + \lambda -1 \choose \lambda}\lambda!\right). \end{equation*} Since binomial coefficients are integers we immediately get the lower bound $v_p(\lambda!) \leq s(y)$. On the other hand, we can also write \begin{equation*} {y+\lambda-1\choose \lambda}\lambda! = {(y+\lambda-1)!\over (y-1)!(\lambda-1)!}(\lambda-1)!, \end{equation*} so for the upper bound it suffices to see \begin{equation}\label{eqn:to-show-valuation-lemma} v_p\left( {(y+\lambda-1)!\over (y-1)!(\lambda-1)!} \right) \leq \floor{\log_p(y+\lambda)} + \min(v_p(y),v_p(\lambda)). \end{equation} Since \eqref{eqn:to-show-valuation-lemma} is symmetric in $\lambda$ and $y$, we may assume that $v_p(y) \leq v_p(\lambda)$. In that case, the classical estimate $v_p\left({n\choose k}\right) \leq \floor{\log_p(n+1)}$ yields \begin{equation*} v_p\left( {(y+\lambda-1)!\over (y-1)!(\lambda-1)!} \right) = v_p(y) + v_p\left({y+\lambda-1 \choose y}\right) \leq v_p(y) + \floor{\log_p(y + \lambda)}. \end{equation*} This completes the proof. \end{proof} Now set $\delta$ be the size of the torsion subgroup in $\Z_p^\times$. Thus $\delta = p-1$ if $p$ is odd and $\delta = 2$ if $p = 2$. If $k_0 \in \Z$, $\lambda > 0$ and $p$ is a prime then we define \begin{equation*} P_{k_0,\lambda}(w) = (w-w_{k_0})(w-w_{k_0-\delta})\dotsb (w-w_{k_0 - (\lambda-1)\delta}). \end{equation*} Thus $P_{k_0,\lambda} \in \Z[w]$ has $\lambda$-many zeros, the highest zero is $k_0$, and the zeros are an arithmetic progression of difference $p-1$ if $p$ is odd and $2$ if $p$ is even (compare with Proposition \ref{prop:change_in_zeroes}). Write $q=p$ if $p$ is odd and $q = 4$ if $p=2$ \begin{lemma}\label{lemma:john-fixed-up} Assume that $k \congruent k_0 \bmod \delta$ and $P_{k_0,\lambda}(w_{k}) \neq 0$. Then \begin{equation*} v_p(P_{k_0,\lambda}(w_{k})) = {q \lambda \over p-1} + O(\log \lambda,\log\abs{k-k_0})) \end{equation*} \end{lemma} \begin{proof} For any $k,k'$ we have $v_p(w_{k} - w_{k'}) = v_p(2p) + v_p(k-k')$. Since $P_{k_0,\lambda}(w_{k}) \neq 0$ and $k \congruent k_0 \bmod \delta$ we have either $k < k_0-(\lambda-1)\delta$ or $k_0 < k$. Thus we deduce that \begin{equation}\label{eqn:first-eqn} v_p(P_{k_0,\lambda}(w_{k})) = v_p(2p)\lambda + \sum_{i=0}^{\lambda-1} v_p(x + i\delta) \end{equation} where $x = k-k_0$ or $x=k_0 - (\lambda-1)\delta - k$ depending on which choice makes $x > 0$. Note that $x \congruent 0 \bmod \delta$. So, replacing $x$ by $y = x/\delta$, \eqref{eqn:first-eqn} becomes \begin{equation}\label{eqn:second} v_p(P_{k_0,\lambda}(w_{k})) = \vartheta\lambda + \sum_{i=0}^{\lambda-1} v_p(y+i) \end{equation} where $\vartheta = 1$ if $p$ is odd and $\vartheta = 3$ otherwise. By \eqref{eqn:second} and Lemma \ref{lemma:valuations} we see \begin{equation}\label{eqn:kanye-west} \vartheta\lambda + v_p(\lambda!) \leq v_p(P_{k_0,\lambda}(w_{k}))\\ \leq \vartheta\lambda + v_p((\lambda-1)!) + \floor{\log_p(y+\lambda)} + \min(v_p(y),v_p(\lambda)). \end{equation} On the left-hand side of \eqref{eqn:kanye-west} we have \begin{equation*} \vartheta\lambda + v_p(\lambda!) \geq {(\vartheta(p-1) + 1)\lambda \over p-1} - \ceil{\log_p(\lambda)} = {q\lambda \over p-1} - \ceil{\log_p(\lambda)}, \end{equation*} and on the right-hand side \eqref{eqn:kanye-west} we have \begin{equation*} \vartheta\lambda + v_p((\lambda-1)!) \leq \vartheta\lambda + {\lambda - 1 \over p-1} \leq {q \lambda \over p-1}. \end{equation*} (Here we've used the classical formula of Legendre for $v_p(n!)$.) By \eqref{eqn:kanye-west} we get \begin{equation*} -\ceil{\log_p(\lambda)} \leq v_p(P_{k_0,\lambda}(w_{k})) - {q\lambda \over p-1} \leq \floor{\log_p(y+\lambda)} + \min(v_p(y),v_p(\lambda)). \end{equation*} Since $y = \abs{k-k_0} + O(\lambda)$ (and $\log_p x = O(\log x)$), we're finished. \end{proof} Now fix $\cal W_{\varepsilon}$ and write $G(w,t) = 1 + \sum g_i(w)t^i$ for the ghost series over $\cal W_{\varepsilon}$. We assume all weights $k$ are in $\cal W_{\varepsilon}$ in what follows. Recall that $\Delta_i = g_i/g_{i-1}$, and if $\Delta_i(w_{k})$ is well-defined then $v_p(\Delta_i(w_{k}))$ is the $i$-th $\Delta$-slope in weight $k$. We define $$ \Delta_i^*(w_{k}) := \begin{cases} (w-w_{k}) \Delta_i(w_{k}) & \text{if~}\Delta_i \text{~has~a~pole~at~}w_{k},\\ \displaystyle \frac{\Delta_i(w_{k})}{w-w_{k}} & \text{if~}\Delta_i \text{~has~a~zero~at~}w_{k},\\ \Delta_i(w_{k}) & \text{otherwise}. \end{cases} $$ Since $\Delta_i$ only has simple zeros or poles, $\Delta_i^*$ has no zeros or poles. \begin{proposition}\label{prop:ith_newtonslope} We have \begin{equation*} v_p\left(\Delta_i^{\ast}(w_{k})\right) = {12 i \over \mu_0(N)(p+1)} + O( \log k, \log i) \end{equation*} \end{proposition} \begin{proof} Recall our standard practice of writing $\Delta_i = \Delta_i^+/\Delta_i^-$ in lowest terms. Write $\lambda_i^+$ for the number of zeros of $\Delta_i^+$ and $\lambda_i^-$ for the number of zeros of $\Delta_i^-$. Write $k_i^+ = \HZ(\Delta_i^+)$ and $k_i^- = \HZ(\Delta_i^-)$. We note it suffices to prove the result separately for pairs $(i,k)$ ranging over a finite number of disjoint domains. With this in mind, we will focus only on the pairs $(i,k)$ such that $w_{k}$ is a zero of $\Delta_i^+$ and leave the other possible pairs for the reader. We will also assume that $p > 3$ or $N > 1$ for simplicity.\footnote{The proof below can easily be modified to handle $p=2,3$ and $N=1$. For example, the corrected formula for $\Delta_i^+$ is $\Delta_i^+(w_{k}) = P_{k_1^+-4,\lambda^+-1}(w_{k})\cdot(w_{12i+2}-w_{k})$, and so the estimates that follow will only be off by $O(\log k, \log i)$.\label{footnote:p=23N=1-asy-fix}} By \eqref{eqn:Deltaizeros}, if $\Delta_i^+(w_{k}) = 0$ then $k = O(i)$ and thus our goal is to show that \begin{equation*} v_p(\Delta_i^{\ast}(w_{k})) = {12 i \over \mu_0(N)(p+1)} + O(\log i). \end{equation*} By Proposition \ref{prop:change_in_zeroes}(a), using that either $p > 3$ or $N > 1$, we have \begin{equation*} \Delta_i^+(w_{k}) = P_{k_i^+,\lambda'}(w_{k})\cdot (w-w_{k}) \cdot P_{k - \delta,\lambda''}(w_{k}) \end{equation*} where $\lambda' + \lambda'' = \lambda^+ - 1$. So, by definition of $\Delta_i^{\ast}$ we have \begin{equation*} v_p\left(\Delta_i^{\ast}(w_{k})\right) = v_p\left(P_{k_i^+,\lambda'}(w_{k})\right) + v_p\left(P_{k - \delta,\lambda''}(w_{k})\right) - v_p\left(P_{k_i^-,\lambda^-}(w_{k})\right) \end{equation*} and $k \congruent k_i^{\pm} \bmod \delta$. Next, by Proposition \ref{prop:change_in_zeroes}(b,c) we have $k_i^+ = O(i)$ and $k_i^- = O(i)$; by Lemma \ref{lemma:zeropole_count} we have $\lambda^+ = O(i)$ and $\lambda^- = O(i)$. Since $k = O(i)$ as well, Lemma \ref{lemma:john-fixed-up} implies \begin{equation*} v_p(\Delta_i^{\ast}(w_{k})) = {q\over p-1}(\lambda ' + \lambda'' - \lambda^-) + O(\log i) = {q\over p-1}(\lambda^+ - \lambda^-) + O(\log i). \end{equation*} Finally by Lemma \ref{lemma:zeropole_count} we have \begin{equation*} {q \over p-1}(\lambda^+ - \lambda^-) = {12i \over \mu_0(N)(p+1)} + O(1). \end{equation*} This completes the proof. \end{proof} To pass from asymptotic control of ghost $\Delta$-slopes as in Proposition \ref{prop:ith_newtonslope} to asymptotic control of ghost slopes, we need to show that $i=d_k$ and $i=d_k+d_k^{\new}$ are asymptotically indices of points on $\oper{NP}(G_k)$ (Lemma \ref{lemma:break} below). First, we give asymptotic control of the ghost slopes over ``oldform'' and ``newform'' ranges. \begin{lemma} \label{lemma:Gbounds} If $x > 0$ then there exists a $k'$ such that if $k \geq k'$ then \begin{enumerate} \item $\displaystyle v_p(\Delta_i(w_{k})) \displaystyle < \left(\frac{1}{p+1} + x\right) k$ for all $i\leq d_k$, and \item $\displaystyle v_p(\Delta_{i}(w_{k})) \displaystyle > \left(\frac{p}{p+1} - x\right)k$ for all $i \geq d_k + d_k^{\new}$. \end{enumerate} \end{lemma} \begin{proof} We check the claim (a) of the lemma as (b) is handled similarly. Note that if $i\leq d_k$ then $\Delta_i^{\ast}(w_{k}) = \Delta_i(w_{k})$, and $i = O(k)$. So, Proposition \ref{prop:ith_newtonslope} implies there is a constant $A\geq0$ such that if $i\leq d_k$ then $$ v_p(\Delta_i(w_{k})) \leq \frac{12i}{\mu_0(N)(p+1)} + A \log k. $$ Since $i\leq \displaystyle d_k = \frac{k\mu_0(N)}{12} + O(1)$, $$ v_p(\Delta_i(k)) \leq \frac{k}{p+1} + A \log k + B $$ for some $B>0$. The lemma clearly follows now. \end{proof} \begin{lemma} \label{lemma:ss_slope} Set $y_i(k) = v_p(g_i(w_{k}))$. Then, $$ \frac{y_{d_k+d_k^{\new}}(k) - y_{d_k}(k)}{d_k^{\new}} = \frac{k}{2} + O(\log k). $$ \end{lemma} \begin{proof} By Proposition \ref{prop:ith_newtonslope}, we have \begin{align} y_{d_k}(k) \label{eqn:ydk-shorter} &= \sum_{i=1}^{d_k} v_p(\Delta_i (w_{k})) = \sum_{i=1}^{d_k} \frac{12i}{\mu_0(N)(p+1)} + O(\log(k)) \\ &= \frac{12}{\mu_0(N)(p+1)} \binom{d_k}{2} + O(k\log k) = \frac{6(d_k)^2 }{\mu_0(N)(p+1)} + O(k\log k) \nonumber \end{align} Among $d_k < i < d_k + d_k^{\new}$, $w_{k}$ is a zero of $\Delta_i$ exactly as many times as it is a pole (by construction), and so $$ \prod_{j=d_k+1}^{d_k+d_k^{\new}} \Delta_i(w_{k}) = \prod_{j=d_k+1}^{d_k+d_k^{\new}} \Delta_i^*(w_{k}). $$ Arguing as above, using Proposition \ref{prop:ith_newtonslope}, gives \begin{equation}\label{eqn:ydk-longer} y_{d_k+d_k^{\new}}(k) = \sum_{i=1}^{d_k+d_k^{\new}} v_p(\Delta_i^*(w_{k})) = {6(d_k+d_k^{\new})^2 \over \mu_0(N)(p+1)} + O(k\log k) \end{equation} Combining \eqref{eqn:ydk-shorter} and \eqref{eqn:ydk-longer}, we deduce that \begin{align*} \frac{y_{d_k+d_k^{\new}(k)} - y_{d_k}(k)}{d_k^{\new}} &= {1 \over d_k^{\new}}\cdot\left(\frac{6((d_k+d_k^{\new})^2-(d_k)^2)}{\mu_0(N)(p+1)} +O(k\log k)\right) \\ &= \frac{6d_{k,p}}{\mu_0(N)(p+1)} + O(\log k) = \frac{k}{2} + O(\log k), \end{align*} as desired. \end{proof} \begin{lemma} \label{lemma:break} For $k\gg0$, $i=d_k$ and $i=d_k+d_k^{\new}$ are indices of break points on $\oper{NP}(G_k)$. \end{lemma} \begin{proof} This is immediate from the two previous lemmas, and the next lemma whose proof we leave to the reader. \end{proof} \begin{lemma} \label{lemma:three_part} Consider a collection $\mathcal P = \{(i,y_i) : i\geq 0\}$ such that $y_i \in \R_{\geq 0} \union \set{\infty}$ and $y_i = \infty$ if and only if $N_1<i<N_2$ for some $N_i \geq 0$. If $i < j$, set $\Delta_{i,j} = {y_j-y_i\over j-i}$, and set $\Delta_i := \Delta_{i-1,i}$. Assume that there are constants $\gamma_i$ such that: \begin{enumerate} \item If $i \leq N_1$ then $\Delta_i \leq \gamma_1$; \item If $N_2 < i$ then $\Delta_i \geq \gamma_2$; and \item $\gamma_1 < \Delta_{N_1,N_2} < \gamma_2$. \end{enumerate} Then, $N_1$ and $N_2$ are indices of break points of $NP(\mathcal P)$. \end{lemma} We now prove the main theorem of this section. \begin{proof}[Proof of Theorem \ref{thm:asymptotic}] Let $C = {12 \over \mu_0(N)(p+1)}$. We need to show that there exists a constant $A > 0$ such that \begin{enumerate}[(i)] \item if $d_k < i \leq d_k + d_k^{\new}$ then $-A \log k \leq s_i(k) - k/2 \leq A \log k$, and \item If $i \leq d_k$ or $i > d_k + d_k^{\new}$ then $-A\max(\log k, \log i) \leq s_i(k) - Ci \leq A \max( \log k, \log i)$ \end{enumerate} It suffices to check $A$ exists independently for each of the four bounds. For (i), if $k$ is fixed then only finitely many $i$ satisfy $i \leq d_k + d_k^{\new}$ and so we may, without loss of generality, assume that $k$ is sufficiently large. In that case, Lemma \ref{lemma:break} implies that the indices $d_k$ and $d_k+d_k^{\new}$ are indices of break points on $\oper{NP}(G_k)$, and so Lemma \ref{lemma:ss_slope} proves (i) holds for some $A>0$. For case (ii), we write $y_i(k) = v_p(g_i(w_{k}))$. To compute an asymptotic for $s_i(k)$ it suffices to assume that $(i-1,y_{i-1}(k))$ is a break point of the Newton polygon. In that case, by definition of Newton polygon, we know that $s_i(k) \leq v_p(\Delta_i(w_{k}))$ for all such $i$ and all $k$ and so Proposition \ref{prop:ith_newtonslope} gives us upper bounds for (ii). Now we deal with lower bounds. We may separately assume that $i \leq d_k$ and $i > d_k + d_k^{\new}$. First assume that $i \leq d_k$. Then $i = O(k)$, so we can choose a constant $A > 0$ such that if $m \leq d_k$ then $v_p(\Delta_m(w_{k})) \geq Cm - A \log k$ (Proposition \ref{prop:ith_newtonslope}). In particular, if $j\geq 0$ and $i + j \leq d_k$ then \begin{multline}\label{eqn:ileqdk} y_{i+j}(k) = y_{i-1}(k) + \sum_{m=i}^{i+j} v_p(\Delta_m(w_{k})) \geq y_{i-1}(k) + \sum_{m=i}^{i+j} (Cm - A \log k)\\ \geq y_{i-1}(k) + (Ci + {Cj\over 2})(j+1) - (A \log k)(j+1). \end{multline} Thus, \begin{equation}\label{eqn:ileqdk2} {y_{i+j}(k) - y_{i-1}(k)\over j+1} \geq Ci + {Cj\over 2} - A \log k. \end{equation} The right-hand side of \eqref{eqn:ileqdk2} is minimized at $j = 0$ and so we deduce \begin{equation}\label{eqn:nearlyileqdkslope} {y_{i+j}(k) - y_{i-1}(k) \over j+1} \geq Ci - A \log k \end{equation} for all $i \leq d_k$ and $j \geq 0$ with $i + j \leq d_k$. Finally by Lemma \ref{lemma:break}, except for finitely many $k$, and thus finitely many $i \leq d_k$, $s_i(k)$ is the slope of line segment connecting index $i-1$ to index $i+j$ for some $i + j \leq d_k$. Thus we conclude $s_i(k) - Ci \geq - A \log k$ for some $A > 0$ and all $i \leq d_k$. Now consider the case where $i > d_k + d_k^{\new}$. If $i$ is fixed then $i \leq d_k$ except for finitely many $k$ and so we may also suppose in what follows that $i$ is sufficiently large (to be determined). Continuing, since $i > d_k + d_k^{\new}$, we have $k = O(i)$ and so the analog of \eqref{eqn:ileqdk2} is \begin{equation}\label{eqn:igreaterdk} {y_{i+j} - y_{i-1} \over j+1} \geq Ci + {Cj\over 2} - A \log(i+j) \end{equation} for all $j\geq 0$. The right-hand side of \eqref{eqn:igreaterdk}, as a function of $j$, has a unique local minimum at $j = 2A/C - i$ and so if we suppose that $i > 2A/C$ then the right-hand side of \eqref{eqn:igreaterdk} is minimized at $j=0$ on the domain $j \geq 0$. The proof is now completed just as before. \end{proof} \section{Halos and arithmetic progressions} \label{sec:global} The goal of this section is to prove that for weights $\kappa$ with $w_{\kappa} \nin \Z_p$, the slopes of $\oper{NP}(G_{\kappa})$ are, except for a finite number of terms, a finite union of arithmetic progressions whose common difference can be explicitly determined. Throughout we will {{\em assume that $p$ is odd}}. See Remark \ref{rmk:p=2-aps-remark} for $p=2$. Fix a component $\cal W_{\varepsilon}$ of $p$-adic weight space, and we implicitly assume all weights lie within $\cal W_{\varepsilon}$ in what follows. Set $\overline{G(w,t)} = \overline{G^{(\varepsilon)}(w,t)} \in \F_p[[w,t]]$ for the reduction modulo $p$ of the ghost series. We write $\oper{NP}(\bar{G})$ for the Newton polygon of $\overline{G(w,t)}$ computed with respect $w$-adic valuation on the coefficients in $\F_p[[w]]$. Write \begin{equation*} C_{p,N} := {p(p-1)(p+1)\mu_0(N)\over 24}, \end{equation*} and if $r \geq 0$ is an integer write $C_{p,N,r} = p^r C_{p,N}$. Since $p$ is odd, $p(p-1)(p+1) \congruent 0 \bmod 24$, so $C_{p,N}$ is an integer divisible by $\mu_0(N)$. Recall that if $w_{\kappa} \nin \Z_p$ then we write $\alpha_{\kappa} = \sup_{w' \in \Z_p} v_p(w_{\kappa}-w') \in (0,\infty)$. \begin{theorem}\label{theorem:global-halo-progressions} Assume that $w_{\kappa} \nin \Z_p$ and write $r = \floor{\alpha_\kappa}$. Then, the slopes of $\oper{NP}(G_{\kappa})$ form a finite union of $C_{p,N,r}$-many arithmetic progressions with common difference \begin{equation*} {(p-1)^2\over 2}\left(\alpha_{\kappa} + \sum_{v=1}^r (p-1)p^{r-v}\cdot v \right) \end{equation*} up to finitely many exceptional slopes contained within the first $C_{p,N,r}$ slopes. \end{theorem} In Theorem \ref{theorem:global-halo-progressions}, the condition that $r = 0$ is equivalent to $0 < v_p(w_{\kappa}) < 1$, and in that case $\alpha_{\kappa} = v_p(w_{\kappa})$. The conclusion is the the slopes of $\oper{NP}(G_{\kappa})$ are, up to a finite number of exceptions, a finite union of $C_{p,N}$-many arithmetic progressions of common difference ${(p-1)^2\over 2}\cdot v_p(w_{\kappa})$. From the ghost spectral halo (Theorem \ref{thm:gsh} in the introduction) we deduce: \begin{corollary} \label{cor:progs} The slopes of $\oper{NP}(\overline{G})$ are a finite union of $C_{p,N}$-many arithmetic progressions whose common difference is $\frac{(p-1)^2}{2}$ up to finitely many exceptional slopes contained within the first $C_{p,N}$ slopes. \end{corollary} \begin{remark} The exceptional slopes in $\oper{NP}(\bar G)$ should not exist (see the comments after Theorem \ref{theorem:intro-aps-boundary} in the introduction) but we have not pursued proving this stronger statement. \end{remark} The remainder of the section is devoted to proving Theorem \ref{theorem:global-halo-progressions}. Our method, as in Section \ref{subsec:distribution}, is to first verify a corresponding statement for ghost $\Delta$-slopes, and, from this, deduce our result about ghost slopes. To this end, here is a general lemma on Newton polygons. \begin{lemma} \label{lemma:newton_slopes} Consider a collection $\mathcal P = \{(i,y_i) : i\geq 0\}$ such that $y_i \in \R_{>0}$. If the $\Delta$-slopes of $\mathcal P$ form a union of $C$ arithmetic progressions with common difference $\delta$, then the same holds for the slopes of $\oper{NP}(\mathcal P)$ up to finitely many exceptional slopes contained within the first $C$ slopes. \end{lemma} \begin{proof} This follows immediately from observing that if $x\geq C$ is the index of a breakpoint of $\oper{NP}(\mathcal P)$, then $x-C$ is also the index of a breakpoint of $\oper{NP}(\mathcal P)$. \end{proof} To deduce Theorem \ref{theorem:global-halo-progressions} from Lemma \ref{lemma:newton_slopes}, we need to verify that the Newton slopes of $G_\kappa$ are a finite union of arithmetic progressions. This is not quite true, but we will show it is true after excluding the weight $w = w_{2}$ from ever appearing as a zero of a ghost coefficient. This modification has no effect on $\oper{NP}(G_\kappa)$. Specifically: \begin{lemma}\label{lemma:remove-wt2-zero} For each $i\geq 1$, write $g_{i}^{\sharp}(w) = g_i(w)(w-w_{2})^{-m}$ where $m = \ord_{w = w_{2}} g_i(w)$. Set $G^{\sharp}(w,t) = 1 + \sum g_i^{\sharp}(w)t^i$ and $\bar{G^{\sharp}}$ as its reduction modulo $p$. Then, $\oper{NP}(G_\kappa) =\oper{NP}(G_{\kappa}^{\sharp})$ for all $\kappa \in \cal W$, and $\oper{NP}(\bar G) = \oper{NP}(\bar{G^{\sharp}})$. \end{lemma} \begin{proof} This follows from (the equality in) Lemma \ref{lemma:explicit-ghost-ordinary-dims}(b). \end{proof} \begin{framed} \vspace{-.20cm} \textbf{Convention:} for the rest of this section {\em we replace the $g_i(w)$ by $g_i^{\sharp}(w)$}. \vspace{-.20cm} \end{framed} We now aim to show that the $\Delta$-slopes of (the newly defined) $G_\kappa$ form a finite union of arithmetic progressions. Recall, $\Delta_i = g_i/g_{i-1}$ and $\Delta_i = \Delta_i^+/\Delta_i^-$ with $\Delta_i^{\pm} \in \Z_p[[w]]$ and $\gcd(\Delta_i^+,\Delta_i^-) = 1$. As preparation, we will compare the zeros of $\Delta_i^{\pm}$ to those of $\Delta^{\pm}_{i+C_{p,N,r}}$. We write $\HZ(-)$ and $\LZ(-)$ for the highest and lowest zeros as in Section \ref{section:explicit-analysis} (if they exist). \begin{lemma} \label{lemma:lambdazp_change} \leavevmode \begin{enumerate} \item If $\lambda(\Delta_i^+)>0$, then \begin{equation*} \HZ(\Delta^+_{i+C_{p,N}}) = \HZ(\Delta_i^+) + \frac{p(p+1)(p-1)}{2} ~\text{~~~and~~~}~ \LZ(\Delta^+_{i+C_{p,N}}) = \LZ(\Delta^+_i) + p(p-1). \end{equation*} \item If $\lambda(\Delta_i^-)>0$, then $$ \HZ(\Delta^-_{i+C_{p,N}}) = \HZ(\Delta^-_i) + p(p-1) ~\text{~~~and~~~}~ \LZ(\Delta^-_{i+C_{p,N}}) = \LZ(\Delta^-_i) + \frac{(p+1)(p-1)}{2}. $$ \end{enumerate} \end{lemma} \begin{proof} We prove the assertions for $\Delta_i^+$ and leave part (b) for the reader (the proofs are analogous). We recall that by \eqref{eqn:Deltaizeros}, for each $i$, $\HZ(\Delta_i^+)$ is the largest $k \in \cal W_{\varepsilon}$ such that $d_k < i$ (with $k\geq 4$, convention in this section). Next, write $C = {p(p+1)(p-1)\over 2} = {12 C_{p,N}\over \mu_0(N)} \congruent 0 \bmod p-1$. Thus $k \mapsto k+C$ preserves the component of weight space. Moreover, since $d_{k+12} = d_k + \mu_0(N)$ we have $d_{k + C} = d_k + C_{p,N}$. Now let $k = \HZ(\Delta_i^+)$ and $k' = \HZ(\Delta_{i+C_{p,N}}^+)$. The previous paragraph implies that $k + C \leq k'$. Write $k'-C = k + j(p-1)$ for some $j\geq 0$. If $j > 0$ then $k'-C > k$ and so by definition of highest zero, $i \leq d_{k'-C}$. But the previous paragraph then implies that $i + C_{p,N} \leq d_{k'}$, which is a contradiction to the definition of $k'$. Proving the formula for $\LZ(\Delta_{i+C_{p,N}}^+)$ is slightly more tedious. Set $k = \LZ(\Delta_i^+)$ and $k' = \LZ(\Delta_{i+C_{p,N}}^+)$. Then $k' \leq k + p(p-1)$ because Lemma \ref{lemma:robs-tedious-calculation} implies that \begin{equation}\label{eqn:lemmaa6-equation} d_{k+p(p-1)} + \left\lfloor \frac{d^{\new}_{k+p(p-1)}}{2} \right\rfloor = d_k + \left\lfloor \frac{d^{\new}_k}{2} \right\rfloor + C_{p,N}. \end{equation} If $p=3$ and $N=1$ then $k' = k+6$ by Table \ref{table:least-zeros}. Thus we assume that $p>3$ or $N>1$. In particular, Lemma \ref{lemma:dim-formulas-increasing} then implies that since we already showed that $k' \leq k + p(p-1)$ we may finish by showing \begin{equation}\label{eqn:newest-realization} d_{k+(p-1)(p-1)} + \bfloor{{d_{k+(p-1)(p-1)}^{\new}\over 2}} < i + C_{p,N}. \end{equation} By definition of $k = \LZ(\Delta_i^+)$, $i \leq d_k +\floor{{d_k^{\new}\over 2}}$ and either \begin{enumerate}[i)] \item $k - (p-1) < 4$, or \item $k-(p-1) \geq 4$ but $d_{k-(p-1)} + \floor{{d_{k-(p-1)}^{\new}\over 2}} < i$. \end{enumerate} If (ii) holds then \eqref{eqn:newest-realization} is immediate from Lemma \ref{lemma:robs-tedious-calculation} and the assumption in (ii). It remains to handle case (i): $4 \leq k \leq p+1$. Then, $k$ is the lowest integer weight $k\geq 4$ on our fixed component and so the assumption that $\lambda(\Delta_i^+) > 0$ and $k = \LZ(\Delta_i^+)$ implies that \begin{equation}\label{eqn:gold-star} d_k < i \leq d_k + \left\lfloor{{d_k^{\new}\over 2}}\right\rfloor. \end{equation} First assume $p\neq 3$. Then Lemma \ref{lemma:robs-tedious-calculation} reduces \eqref{eqn:newest-realization} to showing \begin{equation*} d_k + \bfloor{{d_k^{\new}\over 2}} < i + {(p-1)(p+1)\over 24}\mu_0(N) \end{equation*} instead. But by \eqref{eqn:gold-star}, this reduces to checking for $i=d_k+1$, and in that case checking \begin{equation*} \bfloor{d_k^{\new}\over 2} < 1 + {(p-1)(p+1)\over 24}\mu_0(N). \end{equation*} We leave this final point for the reader. Now assume that $p=3$, so that our assumption now is that $k=4$ is the lowest zero of $\Delta_i^+$. We have $C_{3,N} = \mu_0(N)$. By \eqref{eqn:gold-star} it is enough to show \eqref{eqn:newest-realization} when $i = d_4+1$, and thus we need to check \begin{equation*} d_8 + \bfloor{{d_8^{\new}\over 2}} < d_4 + 1 + \mu_0(N), \end{equation*} which we also leave for the reader. \end{proof} \begin{proposition} \label{prop:lambda_change} If $i\geq 1$ then \begin{enumerate} \item $\lambda(\Delta^+_{i + C_{p,N}}) = \lambda(\Delta^+_i) + \frac{p(p-1)}{2}$, and \item $\lambda(\Delta^-_{i + C_{p,N}}) = \lambda(\Delta_i^-) + \frac{p-1}{2}$. \end{enumerate} \end{proposition} \begin{proof} If $p=3$ and $N=1$, then this proposition follows from Table \ref{table:least-zeros}. Otherwise, for each $i$, Proposition \ref{prop:change_in_zeroes}(a) (valid by our exclusion of $p=3$ and $N=1$) implies that \begin{equation}\label{eqn:lambda-difference} \lambda(\Delta_i^{\pm}) = 1 + {1\over p-1}(\HZ(\Delta_i^{\pm}) - \LZ(\Delta_i^{\pm})). \end{equation} If $\lambda(\Delta_i^+) > 0$ then (a) follows \eqref{eqn:lambda-difference} and Lemma \ref{lemma:lambdazp_change}(a), and if $\lambda(\Delta_i^-) > 0$ then (b) follows from \eqref{eqn:lambda-difference} and Lemma \ref{lemma:lambdazp_change}(b). If $\lambda(\Delta_i^+) = 0$ we proceed as follows (and leave the reader to deal with $\lambda(\Delta_i^-) = 0$). First, if $k < 4 $ is even then we re-define $d_k$ and $d_k^{\new}$ using the formulas \eqref{eqn:dim-formula-old} and \eqref{eqn:dim-formula-new} in Appendix \ref{app:dimension-formulae}. We then define $\HZ(\Delta_i^+)$ and $\LZ(\Delta_i^+)$ by insisting that \eqref{eqn:Deltaizeros} holds, i.e.\ $\HZ(\Delta_i^+)$ is the largest $k \in \cal W_{\varepsilon}$ such that $d_k < i$ and $\LZ(\Delta_i^+)$ is the least $k \in \cal W_{\varepsilon}$ such that $i\leq d_k + \bfloor{{d_k^{\new}/ 2}}$. Continue to suppose that $\lambda(\Delta_i^+) = 0$. Then we must have $\HZ(\Delta_i^+) < \LZ(\Delta_i^+)$ (since otherwise $\Delta_i^+$ would have a zero). Moreover, by definition, $\HZ(\Delta_i^+) \congruent \LZ(\Delta_i^+) \bmod p-1$. We claim that $\LZ(\Delta_i^+) - \HZ(\Delta_i^+) = p-1$. We will prove our claim by contradiction. First, we observe that $\HZ(\Delta_i^+) \geq 0$. Indeed, if $N > 1$ then it is easy to see that if $k\leq 0$ then $d_k < 0 \leq i$, and if $N=1$ then $d_2 = -1 < i$ for any $i\geq 0$, and thus $\HZ(\Delta_i^+) \geq 2$ in fact. Next, if $\LZ(\Delta_i^+) - \HZ(\Delta_i^+) > p-1$ then since $\HZ(\Delta_i^+) \congruent \LZ(\Delta_i^+) \bmod p-1$ (by definition) we can find a $k \in \cal W_{\varepsilon}$ such that $\HZ(\Delta_i^+) < k < \LZ(\Delta_i^+)$. This implies $i \leq d_k$ and $d_k + \floor{{d_k^{\new}\over 2}} < i$, whence $d_k^{\new} < 0$. But this implies that $k \leq 0$, which is a contradiction.\footnote{If $k\geq 4$ then obviously $d_k^{\new}\geq 0$ and the reader may check that $d_2^{\new} \geq 0$ for any $p$ and $N$, given our overwritten definition.} Finally, the reader may check that the proof of Lemma \ref{lemma:lambdazp_change} extends to the new definitions of $\HZ(\Delta_i^+)$ and $\LZ(\Delta_i^+)$, and thus \begin{equation*} \lambda(\Delta^+_{i+C_{p,N}}) = {\HZ(\Delta_i^+)-\LZ(\Delta_i^+)\over p-1} + 1 + {p(p-1)\over 2} = {p(p-1)\over 2}. \end{equation*} This completes the proof. \end{proof} \begin{remark} The $i$-th $w$-adic $\Delta$-slope of $\bar G$ is $\lambda(g_i) - \lambda(g_{i-1}) = \lambda(\Delta_i^+) - \lambda(\Delta_i^-)$. Thus, Proposition \ref{prop:lambda_change} together with Lemma \ref{lemma:newton_slopes} implies Corollary \ref{cor:progs}. \end{remark} We briefly unwind the condition $w_{\kappa} \nin \Z_p$. We thank Erick Knight for pointing out the equivalence in Lemma \ref{lemma:equivalent-conditons-inZp} below. Write $\Z_p^{\nr}$ for the ring of integers in the maximal unramified extension of $\Q_p$ contained in $\bar \Q_p$, and $\omega: \bar \F_p^\times \goto ( \Z_p^{\nr})^\times$ for the Teichm\"uller lift. \begin{lemma}\label{lemma:Zpnr} If $x_0 \in \cal O_{\C_p}$ and there exists $x' \in \Z_p^{\nr} - \Z_p$ such that $v_p(x_0 - x') > v_p(x')$ then $v_p(x_0 - x) = \min(v_p(x_0),v_p(x))$ for all $x \in \Z_p$. \end{lemma} \begin{proof} Suppose $x \in \Z_p$ and $v_p(x_0) = v_p(x)$. We will show $v_p(x_0 - x) = v_p(x_0)$. If $x_0 \in \Z_p^{\nr} - \Z_p$ then the result is clear. Indeed, the assumption $x_0 \nin \Z_p$ implies that the reductions of $p^{-v_p(x_0)}x_0$ and $p^{-v_p(x_0)}x$ are distinct in $\bar \F_p^\times$. Now assume $x_0$ is general. Since $v_p(x_0 - x') > v_p(x')$, we have $v_p(x_0) = v_p(x')$. By the previous paragraph applied to $x'$, we know that $v_p(x' - x) = v_p(x')$. Thus, $v_p(x_0 - x') > v_p(x' - x)$ as well. But then, the ultrametric inequality implies \begin{equation*} v_p(x_0 - x) = v_p(x' - x) = v_p(x') = v_p(x_0), \end{equation*} as promised. \end{proof} \begin{lemma}\label{lemma:equivalent-conditons-inZp} If $x_0 \in \cal O_{\C_p}$ then $x_0 \nin \Z_p$ if and only if either: \begin{enumerate} \item there exists $x' \in \Z_p$ such that $v_p(x_0-x') \nin \Z \union \set{\infty}$, or \item there exists $x' \in \Z_p^{\nr} - \Z_p$ such that $v_p(x_0 - x') > v_p(x')$. \end{enumerate} \end{lemma} \begin{proof} We assume either (a) or (b) holds and we show $\sup_{x \in \Z_p} v_p(x_0-x) < \infty$. If (b) holds, then this is done by Lemma \ref{lemma:Zpnr}. Suppose that (a) holds, and choose such an $x'$ and let $x \in \Z_p$. Then, since $v_p(x_0-x') \nin \Z\union \set{\infty}$ and $v_p(x' - x) \in \Z$ we have \begin{equation*} v_p(x_0 - x) = \min(v_p(x_0-x'),v_p(x'-x)) \leq v_p(x_0-x'). \end{equation*} We now show the converse. Specifically, we show that if $x_0 \in \bar \Z_p - \Z_p$ and $v_p(x_0 - x') \in \Z$ for all $x' \in \Z_p$ then (b) holds. By assumption, $v_p(x_0) \in \Z$, and so $x_0' = p^{v_p(x_0)}\omega(\overline{p^{-v_p(x_0)}x_0}) \in \Z_p^{\nr}$ satisfies $v_p(x_0') = v_p(x_0)$ and $v_p(x_0 - x_0') > v_p(x_0)$. If $x_0' \nin \Z_p$ then we are done. Otherwise set $x_1 = x_0 - x_0'$. Then $x_1$ satisfies all the hypotheses imposed on $x_0$ in this paragraph, and $v_p(x_1) > v_p(x_0)$. Thus we can repeat the construction of $x_1$ from $x_0$, and by induction we can construct an infinite sequence $x_0,x_1,\dotsc$ and $x_0',x_1',\dotsc$ such that $x_{i+1} = x_i - x_i'$ with \begin{enumerate}[i)] \item $x_i \in \bar \Z_p - \Z_p$ with $v_p(x_0) < v_p(x_1) < \dotsb$, \item $x_i' \in \Z_p^{\nr}$ with $v_p(x_i - x_i') > v_p(x_i) = v_p(x_i')$ for all $i\geq 0$, \end{enumerate} In particular $x_0 = \sum x_i'$. Since $x_0 \nin \Z_p$ there exists a smallest $i\geq 1$ such that $x_i' \in \Z_p^{\nr} - \Z_p$. We claim $x_0'-x_i'$ witnesses that (b) is true for $x_0$. To see that, set $x' = x_0' - x_i' \nin \Z_p$. Since $x_0' \in \Z_p$, by Lemma \ref{lemma:Zpnr}, we have \begin{equation*} v_p(x') = \min(v_p(x_i'),v_p(x_0')) = \min(v_p(x_i),v_p(x_0)) = v_p(x_0). \end{equation*} Then \begin{equation*} v_p(x' - x_0) = v_p(-x_1 - x_i') \geq v_p(x_1) > v_p(x_0) = v_p(x'), \end{equation*} as promised. \end{proof} \begin{lemma} \label{lemma:tire_vals} Suppose that $h \in \Z_p[[w]]$ and the zeros of $h$ are all in $p\Z_p$. Let $w' \in \ideal m_{\C_p}$ such that either \begin{enumerate} \item $v_p(w') \nin \Z$, or \item there exists a $\twid w \in \Z_p^{\nr} - \Z_p$ such that $v_p(w' - \twid w) > v_p(\twid w)$. \end{enumerate} If $r = \floor{v_p(w')}$, then \begin{multline*} v_p(h(w')) = v_p(w')\cdot \sizeof \set{w'' \st h(w'') = 0 \text{ and } v_p(w'') \geq r+1} \\ + \sum_{v=1}^r v\cdot \sizeof\set{w'' \st h(w'') = 0 \text{ and } v_p(w'') = v}. \end{multline*} \end{lemma} \begin{proof} In either case, if $w'' \in p\Z_p$ then $v_p(w'-w'') = \min(v_p(w'),v_p(w''))$ (see Lemma \ref{lemma:Zpnr} for case (b)). From this, the statement is immediate. \end{proof} The proof of one final lemma is left to the reader. \begin{lemma} \label{lemma:easy} Suppose that $(k_i)$ is an ordered list of integers which form an arithmetic progression of length $M = p^e u$, with $(u,p)=1$, and difference $\delta$ with $(\delta,p)=1$. Then, \begin{enumerate} \item $\sizeof\set{k_i \st v_p(k_i) \geq e} = u$, and \item if $0 \leq v < e$ then $\sizeof \set{k_i \st v_p(k_i) = v} = u\varphi(p^{e-v}) = u(p-1)p^{e-v-1}$. \end{enumerate} \end{lemma} We're now in position to prove Theorem \ref{theorem:global-halo-progressions}. \begin{proof}[Proof of Theorem \ref{theorem:global-halo-progressions}] Recall that we assume $w_{\kappa} \nin \Z_p$, we write $\alpha_{\kappa} = \sup_{w \in \Z_p} v_p(w_{\kappa} - w)$, and $r = \floor{\alpha_{\kappa}}$. For notational ease, write $C = C_{p,N,r} = p^r C_{p,N}$. Since $w_{\kappa}$ is not an integer, $\Delta_i(w_{\kappa})$ is well-defined for each $i\geq 1$. Our goal is to compare $v_p(\Delta_i(w_{\kappa}))$ to $v_p(\Delta_{i+C}(w_{\kappa}))$ and then apply Lemma \ref{lemma:newton_slopes}. Write $\Delta_i = \Delta_i^+/\Delta_i^-$ as before. We first focus on $\Delta_i^+$. By Proposition \ref{prop:lambda_change}(a) we have $\lambda(\Delta_{i+C}^+) = \lambda(\Delta_i^+) + p^{r+1} \cdot {p-1\over 2}$. By Proposition \ref{prop:change_in_zeroes}(a), the zeros of $\Delta_i^+$ (and $\Delta_{i+C}^+$) are of the form $w_{k}$ with $k$ lying in an arithmetic progression of integers whose difference is $p-1$ (save for possibly one zero when $p=3$ and $N=1$). Write $\Delta_{i+C}^+ = a\cdot b$ where $a,b \in \Z_p[[w]]$, $\lambda(b) = p^{r+1}\cdot {p-1\over 2}$, and where the zeros of $a(w)$ are the highest zeros $w = w_{k}$ of $\Delta_{i+C}^+$ for the highest $\lambda(\Delta_i^+)$-many $k$. Since $w_{\kappa} \nin \Z_p$, $w_{\kappa}$ must satisfy one of the two conditions of Lemma \ref{lemma:equivalent-conditons-inZp}. If (a) is true then choose an integer $k_0$ such that $\alpha_{\kappa} = v_p(w_{\kappa} - w_{k_0}) \nin \Z$, and if (ii) is true then set $w_{k_0} = k_0 = 0$. For each $h \in \set{a,b,\Delta_i^+}$ we then apply Lemma \ref{lemma:tire_vals} to $h(w + w_{k_0})$ and $w' = w_{\kappa} - w_{k_0}$. We deduce (remember $v_p(w_{k}-w_{k_0}) = 1 + v_p(k-k_0)$) that \begin{multline}\label{eqn:h-equation-halo} v_p(h(w_{\kappa})) = \underlabel{\alpha_{\kappa}}{v_p(w_{\kappa}-w_{k_0})} \cdot \sizeof\set{k \st h(w_{k}) = 0 \text{ and } v_p(k-k_0) \geq r}\\ + \sum_{v=0}^{r-1} (v+1) \cdot \sizeof \set{k \st h(w_{k}) = 0 \text{ and } v_p(k-k_0) = v}. \end{multline} Now we claim that $v_p(a(w_{\kappa})) = v_p(\Delta_i^+(w_{\kappa}))$. If $\lambda(\Delta_i^+) = 0$ then there is nothing to show. Otherwise, if $\lambda(\Delta_i^+) > 0$ then Lemma \ref{lemma:lambdazp_change}(a) implies that \begin{equation*} \HZ(a) = \HZ(\Delta_{i+C}^+) \congruent \HZ(\Delta_i^+) \bmod p^{r+1}. \end{equation*} Since the $k$ for which $w_{k}$ is a zero of either $\Delta_i^+$ or $a$ is an arithmetic progression, and the last terms are congruent modulo $p^{r+1}$ (as we just checked), we see that the right-hand side of \eqref{eqn:h-equation-halo} is the same for $h = a$ and $h=\Delta_i^+$. (The reader can check that the single missing zero when $p=3$ and $N=1$ does not affect this argument.) On the other hand, the zeros of $b$ are $w = w_{k}$ with $k$ lying in an arithmetic progression of length $M = p^{r+1}{p-1\over 2}$ and difference $p-1$. Thus it follows from the previous paragraph, Lemma \ref{lemma:easy} and \eqref{eqn:h-equation-halo} that \begin{align*} v_p\left(\Delta_{i+C}^+(w_{\kappa})\over \Delta_i^+(w_{\kappa})\right) &= v_p(b(w_{\kappa})) \\ &= \alpha_{\kappa}\cdot \left({p-1\over 2} +{p-1\over 2}(p-1)\right) + \sum_{v=0}^{r-1} (v+1) \cdot {p-1\over 2}(p-1) p^{r-v}\\ &= p\cdot {p-1\over 2}\left(\alpha_{\kappa} + \sum_{v=1}^r v\cdot (p-1)p^{r-v}\right). \end{align*} An analogous computation shows that \begin{equation*} v_p\left({\Delta_{i+C}^-(w_{\kappa})\over \Delta_i^-(w_{\kappa})}\right) = {p-1\over 2}\left(\alpha_{\kappa} + \sum_{v=1}^r v\cdot (p-1)p^{r-v}\right) \end{equation*} Combining the previous two equations, we deduce \begin{equation*} v_p\left({\Delta_{i+C}(w_{\kappa})\over \Delta_i(w_{\kappa})}\right) = {(p-1)^2\over 2}\left(\alpha_{\kappa} + \sum_{v=1}^r (p-1)p^{r-v}\cdot v\right) \end{equation*} This shows that the $\Delta$-slopes form a union of $C$ arithmetic progressions whose common difference is our claimed one. Our theorem then follows from Lemma \ref{lemma:newton_slopes}. \end{proof} \begin{remark}\label{rmk:p=2-aps-remark} One can ask for a version of Theorem \ref{theorem:global-halo-progressions} valid if $p=2$. If $N=1$ then it is not difficult to establish an analog of Theorem \ref{theorem:global-halo-progressions}. Namely, if $\alpha_{\kappa} < 3$ then the slopes of $\oper{NP}(G_\kappa)$ is $\set{i \cdot v_2(w_{\kappa}) \st i = 1,2,\dotsc}$ (Theorem \ref{theorem:agree-BK-p=2}) and thus a single arithmetic progression with common difference $v_2(w_{\kappa})$. If $\alpha_{\kappa} \geq 3$ and $r = \floor{\alpha_\kappa}$ then one may also show:\ except for finitely many exceptional slopes, the slopes of $\oper{NP}(G_{\kappa})$ are a finite union of $2^{r - 2}$-many arithmetic progressions whose common difference is \begin{equation*} \alpha_\kappa + \sum_{v=3}^{r} v \cdot 2^{r-v}. \end{equation*} The proof is analogous to the above, using Proposition \ref{proposition:explicit-p=2-info} for explicit analogs of Lemma \ref{lemma:lambdazp_change}, Proposition \ref{prop:lambda_change}, etc. One could also ask about $N > 1$. But, since the ghost series requires modification in that case (Section \ref{sec:modification} below) we did not pursue this. \end{remark} \section{A $2$-adic modification for the ghost series}\label{subsec:modification}\label{sec:modification} In this section we construct a modification of the ghost series which we conjecture determines slopes when $p=2$ is $\Gamma_0(N)$-regular (Conjecture \ref{conj:p=2} below). The theme of this section is that non-integral slopes are forced to be repeated and this should be taken into account in the ghost series. We emphasize that $N$ is an odd positive integer in this section. Recall \cite[Definition 1.3]{Buzzard-SlopeQuestions}: \begin{definition}\label{defn:2regular} The prime $p=2$ is called $\Gamma_0(N)$-regular if \begin{enumerate} \item The eigenvalues $T_2$ acting on $S_2(\Gamma_0(N))$ are all $2$-adic units and \item The slopes of $T_2$ acting on $S_4(\Gamma_0(N))$ are all either zero or one. \end{enumerate} \end{definition} Our definition is equivalent to \cite[Definition 1.3]{Buzzard-SlopeQuestions} by Hida theory. Also by Hida theory, \begin{multline}\label{eqn:inequality-ord-dime} \dim S_2(\Gamma_0(2N))^{\ord} \leq \dim S_2(\Gamma_0(N)) + \dim S_{2}(\Gamma_0(2N))^{2-\new}\\ = \dim S_2(\Gamma_0(2N)) - \dim S_2(\Gamma_0(N)) \end{multline} with equality if $p=2$ is $\Gamma_0(N)$-regular. We now produce non-integral slopes for $U_2$ acting on certain spaces with quadratic character regardless of a regularity hypothesis.\footnote{Note:\ not in spaces $S_k(\Gamma_0(2N))$ which would contradict Buzzard's conjecture.} Write $\eta_8^{\pm}$ for the Dirichlet characters of conductor $8$ with sign $\pm$. The character $\eta_8^{\pm}$ is quadratic, so the slopes of $U_2$ acting on $S_k(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^{\pm})$ are symmetric around ${k-1\over 2}$. Hida theory implies that \begin{equation}\label{eqn:p=2_hidacons} \dim S_2(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^+)^{\set{0,1}} = 2\dim S_2(\Gamma_0(2N))^{\ord}. \end{equation} (Here and below, if $S$ is a set of cuspforms and $X$ is a set of real numbers then we write $S^X$ for the subspace spanned by eigenforms whose slope lies in $X$.) \begin{proposition}\label{prop:fractional-slopes-p=2} If $N > 1$ is odd then $\dim S_2(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^+)^{(0,1)} > 0$. \end{proposition} \begin{proof} By \eqref{eqn:inequality-ord-dime} and \eqref{eqn:p=2_hidacons}, we see \begin{multline*} \dim S_2(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^+)^{(0,1)} \geq \\ \dim S_2(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^+) - 2\bigl(\dim S_2(\Gamma_0(2N)) - \dim S_2(\Gamma_0(N))\bigr). \end{multline*} The final expression is positive if $N > 1$ (see Lemma \ref{app-lemma:second-dim-wt-2-cond-8}). \end{proof} Since the characters $\eta_8^{\pm}$ have values only $\pm 1$, any non-integral slope appearing in a space $S_k(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^{\pm})$ must be repeated. In particular, Proposition \ref{prop:fractional-slopes-p=2} implies that for $N > 1$, there exists non-integral repeated slopes in $S_2(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^+)$. The ghost series defined thus far {\em does not} see these slopes:\ \begin{example}\label{example:p=2N=3} $p=2$ is $\Gamma_0(3)$-regular since there are no forms of weight two or four. The space $S_2(\Gamma_0(3)\intersect \Gamma_1(8), \eta_8^+)$ is two-dimensional with slope $1/2$ repeated twice. On the other hand, the ghost series predicts slopes zero and one (see Example \ref{example:issue-p=2}). \end{example} Our goal now is to salvage the ghost conjecture for $p=2$ by including the fractional (repeated) slopes appearing in the spaces $S_k(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^{\pm})$ as $k$ varies and $\pm = (-1)^k$ (we use this implicit notation throughout). Specifically, for each integer $k\geq 2$, we are going to define a second multiplicity pattern $m^{\circ}(k) = (m_i^{\circ}(k))$ which will describe the multiplicity of the weight $z^k\eta_8^{\pm}$ as a zero of a modified ghost series. Our model will be \begin{multline}\label{eqn:mi-circ-hope} m_i^{\circ}(k) > 0 \iff \text{the $i$-th and $(i+1)$-st slope in $S_k(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^{\pm})$}\\\text{are the same and strictly between $k-2$ and $k-1$.} \end{multline} In fact, for each $i$ there will be at most one $k$ such that $m_i^{\circ}(k)$ is positive. Granting the definition of $m_i^{\circ}(k)$, we then define \begin{equation*} g_i^{\circ}(w) = g_i(w) \cdot \prod_{k=2}^\infty (w - w_{z^k\eta_8^{\pm}})^{m_i^{\circ}(k)} \end{equation*} and the {\em modified ghost series} $G^{\circ}(w,t) = 1 + \sum g_i^{\circ}(w)t^i$. It is still an entire series over $\Z_2[[w]]$ (since we've only added more zeros). If $N = 1$ then $G = G^{\circ}$. \begin{conjecture}\label{conj:p=2} If $p=2$ is $\Gamma_0(N)$-regular then $\oper{NP}(G_\kappa^{\circ}) = \oper{NP}(P_\kappa)$ for each $\kappa \in \cal W$. \end{conjecture} We briefly give the evidence we have for Conjecture \ref{conj:p=2}. Recall we write $\oper{BA}(k)$ for the output of Buzzard's algorithm in weight $k$. The levels $N$ in Theorem \ref{theorem:p=2_evidence} below are all the levels $N\leq 167$ such that $p=2$ is $\Gamma_0(N)$-regular. The next $N$ is $191$.\footnote{If $p=2$ is $\Gamma_0(N)$-regular then must $N$ be either 1,3 or be a prime congruent to $7 \bmod 8$? Anna Medvedovsky tells us that $p=2$ is not $\Gamma_0(\ell)$-regular when $\ell > 3$ is a prime $3\bmod 8$.} \begin{theorem}\label{theorem:p=2_evidence} If $N = 3,7,23,31$ then $\oper{NP}((G_k^{\circ})^{\leq d_k}) = \oper{BA}(k)$ for all even $k\leq 5000$, or if $N = 47, 71, 103, 127, 151, 167$ then $\oper{NP}((G_k^{\circ})^{\leq d_k}) = \oper{BA}(k)$ for all even $k\leq 2050$. \end{theorem} \begin{remark} One could ask about the asymptotic results in Section \ref{subsec:distribution}. As we will see, for each $i$, the total multiplicity $\sum_k m_i^{\circ}(k)$ of zeros of $g_i^{\circ}$ which were not a zero of $g_i$ is bounded, and the extra zeros added are at weights $\kappa$ which satisfy $v_2(w_{k} - w_{\kappa}) = 1$ for all $k \in \Z$. Thus the estimates in Section \ref{subsec:distribution} will only be effected by $O(1)$ terms and so Corollary \ref{cor:highest} and Corollary \ref{corollary:gouvea-distribution} should still hold with $G(w,t)$ replaced by $G^{\circ}(w,t)$. \end{remark} The rest of this section is devoted to describing the multiplicity $m_i^{\circ}(k)$ of $w_{z^k\eta_8^{\pm}}$ as a zero of $g_i^{\circ}$. The idea is to {\em force} the issue for $m_i^{\circ}(2)$ by insisting that \eqref{eqn:mi-circ-hope} holds, and that the precise value of $m_i^{\circ}(2)$ follows the up-down pattern within the indices which realize each fractional slope. We then extend the pattern to $k > 2$ ``using the spectral halo'' (Section \ref{subsec:ghost-spectral-intro}). More precisely, if $k\geq 2$ write \begin{equation*} d_k^{\circ} := \dim S_k(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^{\pm}). \end{equation*} Write $\nu_1^{\circ}(2) \leq \nu_2^{\circ}(2) \leq \dotsb \leq \nu_{d_2^{\circ}}^{\circ}(2)$ for the list of slopes of $U_2$ acting on $S_2(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^+)$. Write $\nu_{i_1}^{\circ}(2) < \nu_{i_2}^{\circ}(2) < \dotsb < \nu_{i_t}^{\circ}(2)$ for the {\em distinct} slopes appearing in this list where $i_j$ is the least $i$ such that $\nu_{i_j}^{\circ}(2) = \nu_i^{\circ}(2)$ (so $i_1 = 1$). Also set $i_0 = 0$, $i_{t+1} = d_2^{\circ}$, and $\mu_{j}$ for the multiplicity of $\nu_{i_j}^{\circ}(2)$ among the $\nu_i^{\circ}(2)$. Then set \begin{equation*} m_i^{\circ}(2) = \begin{cases} s_{i}(\mu_{j}-1,i_{j}-1) & \text{if $i_{j} \leq i < i_{j+1}$ for some $1 \leq j \leq t$ \text{and $\nu_{i_j}^{\circ}(2)\neq 0,1$}}\\ 0 & \text{otherwise.}\\ \end{cases} \end{equation*} where $s_i(\ast,\ast)$ is the up-down pattern from Section \ref{subsec:ghost-conjecture}. We give three examples ($p=2$ is $\Gamma_0(N)$-regular for each $N$ below): \begin{example} Let $N = 3$. Then the slopes are computed in Example \ref{example:p=2N=3}, and we have $\nu_{1}^{\circ}(2) = \nu_{2}^{\circ}(2) = {1\over 2}$. Thus $t = 1$, $i_t = i_1 = 1$ and $(m_i^{\circ}(2) \st i \geq 1) = (1,0,0,0,\dotsc)$. \end{example} \begin{example} Let $N = 7$. The slopes of $U_2$ acting on $S_2(\Gamma_0(7)\intersect \Gamma_1(8),\eta_8^+)$ are $[0,{1\over 2},{1\over 2},{1\over 2},{1\over 2},1]$. We have \begin{equation*} \nu_1^{\circ}(2) = 0 < \nu_2^{\circ}(2) = \dotsb \nu_5^{\circ}(2) = {1\over 2} < \nu_6^{\circ}(2) = 1. \end{equation*} Thus $t = 3$, $(i_1,i_2,i_3) = (1,2,6)$ and $(m_i^{\circ}(2) \st i \geq 1) = (0,1,2,1,0,0,\dotsc)$. \end{example} \begin{example} Let $N = 23$. Then the slopes are $[0_3, ({1\over 3})_6, ({1\over 2})_4, ({2 \over 3})_6, 1_3]$ (the subscripts refer to the multiplicity). The sequence $m_i^{\circ}(2)$ is given by \begin{equation*} (m_i^{\circ}(2) \st i \geq 1) = (0,0,0,1,2,3,2,1,0,1,2,1,0,1,2,3,2,1,0,0,0,0,0,0,0,\dotsc). \end{equation*} \end{example} Now, if $k > 2$ then we will set \begin{equation}\label{eqn:p=2-mults} m_i^{\circ}(k) = \begin{cases} m_{d_k^{\circ}-i}^{\circ}(2) & \text{if $1 \leq i < d_k^{\circ}$}\\ 0 & \text{otherwise.} \end{cases} \end{equation} This completes the definition of the $m_i^{\circ}(k)$ and completes the statement of Conjecture \ref{conj:p=2}. The rest of this section is devoted to expanding on the definition of $m_i^{\circ}(k)$ when $k>2$. First, the authors believe that a version of the spectral halo will imply that $k \mapsto \oper{NP}(P_{z^k\eta_8^{\pm}})$ is independent of $k$. In particular, if our {\em modus operandi} is to predict the fractional slopes appearing in $S_k(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^{\pm})$ then we should restrict to slopes between $k-2$ and $k-1$ (the lower slopes being correctly predicted ``by induction'' on $k$; compare with Remark \ref{remark:p2zerosremark}). Now, when is the $i$-th and $(i+1)$-st slope going to be more than $k-2$ and not more than $k-1$? First, ``by the spectral halo'' we should certainly have $d_{k-1}^{\circ} < i$. But, there are also the $c_0(N)$-many $\theta^{k-2}$-critical Eisenstein series which are overconvergent $p$-adic cuspforms of weight $z^{k-1}\eta_8^{\pm}$ and slope $k-2$ ($c_0(N)$ being the number of cusps of $X_0(N)$). Thus if we want the $i$-th and $(i+1)$-st slope to be larger than $k-2$, we should expect $d_{k-1}^{\circ} + c_0(N) < i$. We now note the following lemma. \begin{lemma} If $k > 2$ and $d_{k-1}^{\circ} + c_0(N) < i$ then $d_k^{\circ}-i < d_2^{\circ}$. \end{lemma} \begin{proof} By Lemma \ref{app-lemma:dim-wt-2-cond-8} we have $d_2^{\circ} = \mu_0(N) - c_0(N)$ and if $k > 2$ then $d_k^{\circ} = \mu_0(N) + d_{k-1}^{\circ}$. The lemma clearly follows then. \end{proof} Now write $\nu_1^{\circ}(k) \leq \nu_2^{\circ}(k) \leq \dotsb$ for the slopes of $U_2$ acting on $S_k(\Gamma_0(N)\cap \Gamma_1(8),\eta_8^{\pm})$. Since $\eta_8^{\pm}$ is quadratic, the Atkin--Lehner involution implies that \begin{equation*} \nu_i^{\circ}(k) = \nu_{i+1}^{\circ}(k) \text{ is in $(k-2,k-1)$} \iff \nu_{d_k^{\circ}-i}^{\circ}(k) = \nu_{d_k^{\circ}-i+1}^{\circ}(k) \text{ is in $(0,1)$.} \end{equation*} ``By the spectral halo'', we have an equivalence \begin{equation*} \nu_{d_k^{\circ}-i}^{\circ}(k) = \nu_{d_k^{\circ}-i+1}^{\circ}(k) \text{ is in $(0,1)$} \iff \nu_{d_k^{\circ}-i}^{\circ}(2) = \nu_{d_k^{\circ}-i+1}^{\circ}(2) \text{ is in $(0,1)$.} \end{equation*} We just justified that our natural constraint on $i$ should force $d_k^{\circ} - i < d_2^{\circ}$, so that the right-hand side of the previous equivalence exactly describes when $m_{d_k^{\circ}-i}(2) > 0$, and strongly suggests the definition \eqref{eqn:p=2-mults} is natural. \begin{remark}\label{remark:p2zerosremark} It is not hard to see that if $1 \leq i < \infty$ then there exists at most one $k$ for which $m_i^{\circ}(k) > 0$, so we can then write $m_i^{\circ\circ}$ for this non-zero value, if it exists. Based on our heuristic of using the spectral halo, one could also form an alternate modification \begin{equation*} g_i^{\circ\circ}(w) = g_i(w)(w-w_{z^2\eta_8^{\pm}})^{m_i^{\circ\circ}} \end{equation*} by adding a zero at the single weight $\kappa = z^2\eta_8^{\pm}$ infinitely often. Then one could form an alternate modified ghost series $G^{\circ\circ}(w,t) = 1 + \sum g_i^{\circ\circ}(w)t^i$. Numerical checks suggests that $\oper{NP}(G^{\circ}_\kappa) = \oper{NP}(G^{\circ\circ}_\kappa)$ for all $\kappa$, but we will not pursue proving that here. It is certainly true if $v_2(w_\kappa) > 1$. \end{remark} \begin{appendix} \section{Dimension formulas} \label{app:dimension-formulae} The goal of this appendix is to gather together various estimates and formulas for the dimensions of spaces of cuspforms. The results for spaces with trivial character are deduced from the standard formulas in \cite[Section 6.1]{Stein-ModularForms}.\footnote{Freely available at \href{http://wstein.org/books/modform/modform/dimension_formulas.html}{{\tt http://wstein.org/books/modform/modform/dimension\textunderscore formulas.html}.}} We use the notation(s):\ $\mu_0(N)$ for the index of $\Gamma_0(N)$ in $\SL_2(\Z)$, $c_0(N)$ for the number of cusps of $X_0(N)$, $\mu_{0,2}(N)$ for the number of elliptic points of order two on $X_0(N)$, $\mu_{0,3}(N)$ for the number of elliptic points of order three on $X_0(N)$ and $g_0(N)$ for the genus of $X_0(N)$. Many proofs are asymptotically clear and we often leave the details of explicit constants to the reader. We fix $N$ and $p$ throughout the appendix and we will also assume that $p \ndvd N$ as a rule. As in the main text we write $d_k = \dim S_k(\Gamma_0(N))$, $d_{k,p} = \dim S_k(\Gamma_0(Np))$ and $d_k^{\new} = \dim S_k(\Gamma_0(Np))^{p-\new}$. For example, if $k > 2$ is even then \begin{equation}\label{eqn:dim-formula-old} d_k = (k-1)(g_0(N) - 1) + \left({k\over 2} - 1\right)c_0(N) + \bfloor{k\over 4}\mu_{0,2}(N) + \bfloor{k\over 3}\mu_{0,3}(N) \end{equation} where $g_0(N)$ may be written as \begin{equation*} d_2 = g_0(N) = 1 + {\mu_0(N)\over 12} - {\mu_{0,2}(N) \over 4} - {\mu_{0,3}(N) \over 3} - {c_0(N)\over 2}. \end{equation*} To compute $d_{k,p}$, one replaces $N$ by $Np$ everywhere in \eqref{eqn:dim-formula-old}. Then it is easy to check that for $p \ndvd N$ and $k>2$ then \begin{multline}\label{eqn:dim-formula-new} d_k^{\new} = {(k-1)(p-1)\over 12}\mu_0(N) + \left(\bfloor{{k\over 4}} - {k-1\over 4}\right)\left(-1 + \left({-4 \over p}\right)\right)\mu_{0,2}(N)\\ + \left(\bfloor{{k\over 3}} - {k-1\over 3}\right)\left(-1 + \left({-3 \over p}\right)\right)\mu_{0,3}(N), \end{multline} where $\left({a \over b}\right)$ is the Kronecker symbol. \begin{lemma}\label{mu:inequalities-need-for-dimensions} If $N > 1$ then $\displaystyle {1\over 6}\mu_0(N) - {1\over 2}\mu_{0,2}(N) - {2\over 3}\mu_{0,3}(N) \geq 0. $ \end{lemma} \begin{proof} Let $\omega(N)$ denote the number of distinct prime divisors of $N$. Then $\omega(N) \leq \log_3(N)$ if $N\geq 6$ and $\mu_{0,i}(N) \leq 2^{\omega(N)}$ for $i=2,3$. Moreover, if $N \geq 200$ then $N \geq 7\cdot 2^{\log_3(N)}$. Thus for $N\geq 200$ we conclude \begin{equation*} \mu_0(N) \geq N \geq 7 \cdot 2^{\log_3(N)} \geq 7\cdot 2^{\omega(N)} \geq 6\cdot\left({1\over 2}\mu_{0,2}(N) + {2\over 3}\mu_{0,3}(N)\right). \end{equation*} We leave checking $2 \leq N \leq 200$ for the reader (or a computer). \end{proof} \begin{lemma}\label{lemma:dim-formulas-increasing} \leavevmode \begin{enumerate} \item If $N > 1$ then $k \mapsto d_k$ is a weakly increasing function of even weights $k\geq 2$. \item If $N = 1$ and $p>3$ then $n \mapsto d_{k+n(p-1)}(\SL_2\Z)$ is increasing. \item If $N > 1$ or $p > 3$ then $d_k^{\new} \leq d_{k + \varphi(2p)}^{\new}$ \end{enumerate} \end{lemma} \begin{proof} For (a) it is clear from \eqref{eqn:dim-formula-old} that if we restrict to $k\geq 4$ and either $g_0(N) \geq 1$ or $c_0(N) \geq 2$. That leaves $N=1$, which we've excluded, and checking $d_2 \leq d_4$ (which is easy). Part (b) is also easy. If $N=1$, $j\geq 0$ and $d_k > d_{k+j}$ then $k \congruent 0 \bmod 12$ and $j = 2$. In particular, if $p$ is odd and $d_k > d_{k+(p-1)}$ then $p=3$. Let's prove (c). First, using \eqref{eqn:dim-formula-new} to compute $d_{k+\varphi(2p)}^{\new} - d_k^{\new}$, one uniformly sees that if $k\geq 4$ then \begin{equation}\label{eqn:difference-equation} d_{k+\varphi(2p)}^{\new} - d_k^{\new} \geq {\varphi(2p)(p-1) \over 2}\mu_0(N) - \mu_{0,2}(N) - {4\over 3}\mu_{0,3}(N). \end{equation} If $N > 1$ then Lemma \ref{mu:inequalities-need-for-dimensions} implies the right-hand side is $\geq 0$ as long as $p \geq 7$. We leave the remaining cases of $p=2,3,5$ and $N>1$, $N = 1$ and $p>3$, and $k=2$ to the reader (one just needs to make the lower bound \eqref{eqn:difference-equation} more explicit.) \end{proof} \begin{lemma}\label{lemma:weight2-inequality} If $p$ is odd and $n\geq 1$ then $d_{2 + n(p-1)} \geq d_2 + d_2^{\new}$ with equality if $n = 1$. \end{lemma} \begin{proof} Let's first show equality for $n=1$. One computes \begin{multline}\label{eqn:difference-blah} d_{2+(p-1)} - (d_2 + d_2^{\new})\\= \left(\bfloor{p+1\over 4} - {p \over 4} + {1\over 4}\left({-4 \over p}\right)\right)\mu_{0,2}(N) + \left(\bfloor{p+1\over 3} - {p \over 3} + {1\over 3}\left({-3 \over p}\right)\right)\mu_{0,3}(N). \end{multline} The right-hand side of \eqref{eqn:difference-blah} clearly vanishes for all odd $p$. Next, Lemma \ref{lemma:dim-formulas-increasing} allows us to finish except if $N=1$ and $p=3$, where the result is trivial anyways because $d_2 + d_2^{\new} = 0$. \end{proof} \begin{lemma}\label{lemma:appendix-2adic-wt-2} Let $p=2$. If $n\geq 1$ then $d_{2(n+1)} \geq d_2 + d_2^{\new}$ with equality when $n=1$ only if $N=1,3,7$. \end{lemma} \begin{proof} One checks explicitly that \begin{equation*} d_4 - (d_2 + d_2^{\new}) = {1 \over 12}\mu_0(N) + {1\over 4}\mu_{0,2}(N) - {1\over 3}\mu_{0,3}(N), \end{equation*} and this is equal to zero if and only if $N = 1,3,7$ by an argument similar to Lemma \ref{mu:inequalities-need-for-dimensions}. If $N > 1$ then we are finished by Lemma \ref{lemma:dim-formulas-increasing}. If $N = 1$ then $d_2 + d_2^{\new} = 0$ so the result is trivial in that case. \end{proof} \begin{lemma}\label{app-lemma:shift-wt-p(p-1)} If $p\geq 5$ is odd, $k\geq 4$ is even and $j \geq 0$ then $$ \displaystyle d_{k+j(p-1),p} - d_{k,p} = {j(p-1)(p+1)\over 12}\mu_0(N). $$ If $p=3$ then the same holds for $j \congruent 0 \bmod 3$. \end{lemma} \begin{proof} Let $p\geq 3$. Then, \begin{equation*} \left(\bfloor{{k + j(p-1)\over 4}} - \bfloor{{k\over 4}}\right)\mu_{0,2}(Np) = {j(p-1)\over 4}\mu_{0,2}(Np). \end{equation*} If $p\congruent 1 \bmod 4$ this is clear, and if $p \congruent 3 \bmod 4$ then both sides vanish because $\mu_{0,2}(Np) = 0$. Similarly, if either $p \geq 5$ or if $j \congruent 0 \bmod 3$ then \begin{equation*} \left(\bfloor{{k + j(p-1)\over 3}} - \bfloor{{k\over 3}}\right)\mu_{0,3}(Np) = {j(p-1)\over 3}\mu_{0,3}(Np). \end{equation*} Thus, \begin{multline*} d_{k+j(p-1),p} - d_k\\ = j(p-1)(g_0(Np)-1) + {{j(p-1)\over 2}c_0(Np)} + {j(p-1)\over 4}\mu_{0,2}(Np) + {j(p-1)\over 3}\mu_{0,3}(Np)\\ = {j(p-1)\over 12}\underlabel{\mu_0(Np)}{\left(12(g_0(Np)-1) + 6c_0(Np) + 3\mu_{0,2}(Np) + 4\mu_{0,3}(Np)\right)}. \end{multline*} Since $\mu_0(Np) = (1+p)\mu_0(N)$ we are done. \end{proof} \begin{lemma}\label{lemma:robs-tedious-calculation} If $p\geq 5$, $k\geq 4$ is even and $j\geq 0$ then \begin{equation*} d_{k+j(p-1)} - d_k + \bfloor{{d_{k+j(p-1)}^{\new}\over 2}} - \bfloor{{d_k^{\new}\over 2}} = {j(p-1)(p+1)\over 24}\mu_0(N). \end{equation*} If $p = 3$ then the same is true of $j \congruent 0 \bmod 3$. \end{lemma} \begin{proof} Since $d_k^{\new} \congruent d_{k,p} \bmod 2$, Lemma \ref{app-lemma:shift-wt-p(p-1)} implies that $d_{k+j(p-1)}^{\new} \congruent d_k^{\new} \bmod 2$. Thus, \begin{multline*} d_{k+j(p-1)} - d_k + \bfloor{{d_{k+j(p-1)}^{\new}\over 2}} - \bfloor{{d_k^{\new}\over 2}} \\ = d_{k+j(p-1)} - d_k + {1\over 2}(d_{k+j(p-1)}^{\new} - d_k^{\new}) = {1\over 2}(d_{k+j(p-1),p} - d_{k,p}) \end{multline*} Thus Lemma \ref{app-lemma:shift-wt-p(p-1)} finishes the proof. \end{proof} We finish with formulas for spaces with character. For this we use Cohen--Oesterl\'e \cite{CohenOesterle-Dimensions}. \begin{lemma}\label{app-lemma:dim-wt-2-cond-8} If $N \geq 1$ is odd, $k\geq 2$ is even and $\eta_8^{\pm}$ is the primitive character modulo $8$ such that $\eta_8^{\pm}(-1) = (-1)^k$ then \begin{equation*} \dim S_k(\Gamma_0(N)\intersect \Gamma_1(8),\varepsilon) = (k-1)\mu_0(N) - c_0(N). \end{equation*} \end{lemma} \begin{proof} This is immediate from \cite[Th\'eor\`eme 1]{CohenOesterle-Dimensions}. One should take $8N$ for $N$ in the reference, $\chi=\eta_8^{\pm}$ and note remark $1^{\circ}$ in {\em loc. cit.} \end{proof} \begin{lemma}\label{app-lemma:second-dim-wt-2-cond-8} If $N\geq 1$ is odd and $\eta_8^+$ is the even primitive character modulo $8$ then \begin{equation*} \dim S_2(\Gamma_0(N)\intersect \Gamma_1(8),\eta_8^+) - 2\bigl(\dim S_2(\Gamma_0(2N)) - \dim S_2(\Gamma_0(N))\bigr) = {2 \over 3}\left(\mu_0(N) - \mu_{0,3}(N)\right). \end{equation*} In particular, it is positive if and only if $N > 1$. \end{lemma} \begin{proof} One computes explicitly that \begin{align*} \dim S_2(\Gamma_0(2N)) - \dim S_2(\Gamma_0(N)) = {\mu_0(N)\over 6} + {\mu_{0,3}(N)\over 3} - {c_0(N)\over 2}. \end{align*} The equality then follows from Lemma \ref{app-lemma:dim-wt-2-cond-8} applied with $k=2$. Regarding the positivity, if $N = 1$ then $\mu_0(N) = \mu_{0,3}(N) = 1$. To check it is positive if $N > 1$, it reduces to a finite computation (which we leave to the reader). \end{proof} \section{The story of an explicit calculation when $p=2$ and $N = 1$} \label{app:explicit-2adic} The goal of this appendix is to describe the $2$-adic calculation we made which motivated the multiplicity pattern and the use of ``ghost'' in ``the ghost conjecture''. We begin by quoting an unpublished note of Buzzard.\footnote{Page 2 of the note ``Explicit formulae..." at \href{http://wwwf.imperial.ac.uk/~buzzard/maths/research/notes/}{{\tt http://wwwf.imperial.ac.uk/\textasciitilde buzzard/maths/research/notes/}}} In it, he writes: \begin{quote} ``...the trace [of $U_2$ acting on overconvergent $2$-adic cuspforms] vanishes at weight $w = 2^3+2^5+2^6+2^7+2^8+2^{13}+2^{16}+2^{18}+2^{19}+\dotsb$, and this corresponds to $k = 2+2^2 +2^3 +2^{11} +2^{15} +2^{16} +2^{18} + \dotsb$, which, unsurprisingly, is close to 14." \end{quote} Indeed, $S_{14}(\Gamma_0(2))$ has two distinct eigenforms, both new at 2 and whose $U_2$-eigenvalues are $6$ and $-6$. Thus $\tr(\restrict{U_2}{S_{14}(\Gamma_0(2))}) = 0$ and $\tr(\restrict{U_2}{S_{14}^{\dagger}(\Gamma_0(2))}) \congruent 0 \bmod 2^{13}$. One can check that the zero $w = 2^3 + 2^5 + \dotsb$ satisfies $v_2(w_{14}-w) = 13$. In this way, $w_{14}$ is a ``ghost zero'' of the trace:\ it is an integer weight and the true zero of the trace is only a slight $2$-adic deformation. In order to investigate whether the above phenomenon generalizes, we implemented Koike's formula \cite{Koike-padicProperties} on a computer and computed the first twenty coefficients of $P_\kappa(t) = 1 + \sum a_i(w_\kappa)t^i$ (see \cite{Robwebsite}). For each $i \leq 20$ we noticed that if $a_i(w_0) = 0$ then $v_2(w_0) \in \Z$ (see \cite[Appendix B]{BergdallPollack-FredholmSlopes}). Thus, it seems possible that the roots of the $a_i$ are relatively near actual integer weights $w_{k}$. And, we conjectured that for some meaning of ``relatively near'', the $w_{k}$ could be taken so that the $i$-th and $(i+1)$-st slope in weight $k$ is ${k-2\over 2}$, i.e.\ $k=6i+8,6i+10,\dotsc,12i-2,12i+2$ (Proposition \ref{proposition:explicit-p=2-info}). Let's see how this works out. We just pointed out that the unique zero of $a_1$ lies on $v_2(w-w_{14}) = 13$. For $a_2$, the predicted ghost zeros are $w_{20}, w_{22}$ and $w_{26}$. In Table \ref{table:b2-zero-location} below we give the relative position of the zeros of $a_2$ to these three weights. We see what we want:\ the true zeros of $a_2$ are slight $2$-adic deformations of $w_{k}$ with $k=20,22,26$. Similarly, one can work out that the six weights $w_{26},w_{30},\dotsc,w_{34},w_{38}$ are ghost zeros for the third coefficient (which has six zeros). \begin{table}[htpp] \caption{Relative position of zeros of $a_2(w)$ to the weights $w_{k}$ for $k=20,22,26$. (Bold indicates the witnesses to $k$ as a ``ghost zero''.)} \begin{center} \begin{tabular}{|c|c|c|c|} \hline $k$ & $20$ & $22$ & $26$\\ \hline $v_2(w_0-w_{k}) \st a_2(w_0) = 0$ & $\mathbf{12}, 3,3$& $\mathbf{13}, 4, 3$ & $\mathbf{9}, 4, 3$\\ \hline \end{tabular} \end{center} \label{table:b2-zero-location} \end{table} A departure must occur for the fourth coefficient:\ $a_4$ has ten zeros and there are only nine predicted ghost zeros. The relative position of the ten zeros to the nine predictions are given in Table \ref{table:b4-zero-location}. What we see is that for each $k=32,34,\dotsc,46,50$ there is a small $2$-adic disc around $w_{k}$ containing at least one root of $a_4$, and that there are actually two roots in a small disc around $w_{38}$. In this sense, $38$ is a ghost zero for $a_4$ with multiplicity two and the rest of the $w_{k}$ have multiplicity one. \begin{table}[htpp] \caption{Relative position of zeros of $a_4(w)$ to the weights $w_{k}$ for $k=32,34,\dotsc,46,50$. (Bold indicates the witnesses to $k$ as a ``ghost zero''.) } \begin{center} \begin{tabular}{|c|l|} \hline $k$ & $v_2(w_0 - w_{k})$ where $a_4(w_0) = 0$\\ \hline $32 $&$\mathbf{9}, 5, 4, 4, 3,\dotsc$\\ $34 $&$ \mathbf{9}, 6, 5, 4, 4,\dotsc$\\ $36 $&$ \mathbf{15}, 5, 4, 4, 3,\dotsc$\\ $38 $&$ \mathbf{{21 \over 2}, {21 \over 2}}, 5, 4, 4,\dotsc$\\ $40 $&$ \mathbf{9}, 5, 4, 4, 3,\dotsc$\\ $42 $&$ \mathbf{11}, 5, 5, 4, 4,\dotsc$\\ $44 $&$ \mathbf{34}, 5, 4, 4, 3,\dotsc$\\ $46 $&$ \mathbf{36}, 5, 5, 4, 4,\dotsc$\\ $50 $&$ \mathbf{14}, 6, 5, 4, 4,\dotsc$\\ \hline \end{tabular} \end{center} \label{table:b4-zero-location} \end{table} Continuing then with the weight $k=38$, it was a ghost zero for $a_3$ with multiplicity one, multiplicity two for $a_4$ and one can check it should have multiplicity one for $a_5$ (see Table \ref{table:weight62-zeros}) With these computations in mind, we cataloged the relative location of the zeros of $a_5,a_6,\dotsc$ to the ghost zeros we were predicting. Seeing the data, and writing down the multiplicity $k$-by-$k$ we saw what became the multiplicity pattern:\ for each $k$, the first and last time appear of $k$ as a ghost zero it has multiplicity one, the second and second to last time it has multiplicity two, etc. To emphasize this, in Table \ref{table:weight62-zeros} below we give the relative positions of the zeros of each $a_i$ to the weights $w_{38}$ and $w_{62}$, with the multiplicity pattern emphasized through the use of bolding. \begin{table}[htpp] \caption{Relative location of zeros of $a_1(w),\dotsc,a_{10}(w)$ for $w_{38}$ and $w_{62}$. (Bold indicates the witnesses to $k$ as a ``ghost zero''.)} \begin{center} \begin{tabular}{|c|l|l|} \hline $i$ & $v_2(w_0-w_{38})$ where $a_i(w_0) = 0$ & $v_2(w_0-w_{62})$ where $a_i(w_0) = 0$ \\ \hline $1$ & $5$ &$ 6$\\ $2$ & $6, 4, 3$ & $ 5, 4, 3$\\ $3$ & $\mathbf{31}, 5, 4, 4, 3,\dotsc $ & $ 7, 5, 4, 4, 3,\dotsc$\\ $4$ & $\mathbf{{21\over 2}, {21\over 2}}, 5, 4, 4,\dotsc$ & $ 6, 5, 5, 4, 4,\dotsc$\\ $5$ & $\mathbf{22}, 6, 5, 5, 5,\dotsc$ & $ \mathbf{30}, 6, 6, 5, 5,\dotsc$\\ $6$ & $7, 6, 6, 5, 5,\dotsc$ & $ \mathbf{14, 14}, 6, 5, 5,\dotsc$\\ $7$ & $7, 7, 6, 6, 5,\dotsc$ & $\mathbf{29, {23\over 2}, {23\over 2}}, 6, 5,\dotsc$\\ $8$ & $7, 7, 7, 6, 6,\dotsc$ & $ \mathbf{14, 14}, 7, 6, 6,\dotsc$\\ $9$ & $8, 7, 7, 6, 6,\dotsc$ & $ \mathbf{30}, 7, 7, 6, 6,\dotsc$\\ $10$ & $8, 8, 7, 6, 6,\dotsc$ & $ 7, 7, 7, 6, 6,\dotsc$\\ \hline \end{tabular} \end{center} \label{table:weight62-zeros} \end{table} \end{appendix} \newpage
3f189e1dc54c53d255f59965f7bb1f42621905ca
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Robots are increasingly being used in living spaces, factories, and outdoor environments. One recent trend has been the development of co-robots (or cobots), robots that are intended to physically interact with humans in a shared workspace. In such environments, various elements or parts of the robot tend to be in close proximity to the humans or other moving objects. This proximity gives rise to two kinds of challenges in terms of motion planning. First, we have to predict the future actions and reactions of moving obstacles or agents in the environment to avoid collisions with obstacles. Therefore, the collision avoidance algorithm needs to deal with uncertain and imperfect representation of future obstacle motions efficiently. Second, the computed robot motion still needs to be reasonably efficient. It is not desired to compute a very slow or excessively diverting trajectory in order to avoid collisions. Various uncertainties arise from control errors, sensing errors, or environmental errors (i.e. imperfect environment representation) in the estimation and prediction of environment obstacles. Typically, these uncertainties are modeled using Gaussian distributions. Motion planning algorithms use probabilistic collision detection to avoid collisions with the given imperfect obstacle representation. With such obstacle representations, it can be impossible (Gaussian distributions of obstacle positions have non-zero probabilities in the entire workspace) to compute a perfectly collision-free, or results in an inefficient trajectory to avoid collisions with very low probabilities. In order to balance the safety and efficiency of planned motions, motion planning under uncertainties is desired to guarantee collision-free of the computed trajectory only in a limited probabiliy bound, which can be specified using a parameter, \emph{confidence level} (e.g. 0.99)~\cite{du2012robot}. For the probabilistic collision detection, stochastic algorithms are used to approximate the collision probability \cite{blackmore2006probabilistic,lambert2008fast}. However, such probabilistic collision detection algorithms are computationally intensive and mostly limited to 2D spaces. Most prior planning approaches for high-DOF robots perform the exact collision checking with scaled objects that enclose the potential object volumes~\cite{bry2011rapidly,van2012lqg,lee2013sigma,sun2015stochastic}. Although these approaches guarantee probabilistical safety bounds, they highly overestimate the collision probability, which result in less optimal trajectories or failure to finding feasible trajectories in a limited planning time in dynamic environments. \noindent {\bf Main Results:} In this paper, we present a novel approach to perform probabilistic collision detection. Our approach has two novel contributions. First, we provide a fast approximation of collision probability between the high-DOF robot and high-DOF obstacles. Our approach computes more accurate probabilities than approaches using exact collision checking with enlarged obstacle shapes, and the computed probability is guaranteed as the upper bound that the actual probability is always lower than the computed probability. Second, we describe a practical belief space estimation algorithm that accounts for both spatial and temporal uncertainties in the position and motion of each obstacle in dynamic environments with moving obstacles. We present a trajectory optimization algorithm for high-DOF robots in dynamic, uncertain environments based on our probabilistic collision detection and belief space estimation. We have evaluated our planner using robot arms operating in a simulation and a real environment workspace with high-resolution point cloud data corresponding to moving human obstacles, captured using a Kinect. Our approach uses a high value of the confidence level ($0.95$ or above) to perform probabilistic collision detection and can compute a smooth collision-free trajectory. The paper is organized as follows. Section~\ref{sec:related} gives a brief overview of prior work on probabilistic collision detection and motion planning. We introduce the notation and the algorithm of our probabilistic collision detection algorithm in Section~\ref{sec:pcc}. We describe the belief space estimation and trajectory planning algorithm in Section~\ref{sec:environment} and Section~\ref{sec:optimization}, respectively. We highlight planning performance in challenging human environment scenarios in Section~\ref{sec:result}. \section{Related Work} \label{sec:related} In this section, we give a brief overview of prior work on probabilistic collision detection, trajectory planning, and uncertainty handling. \subsection{Probabilistic Collision Detection} Collision checking is an integral part of any motion planning algorithm and most prior techniques assume an exact representation of the robot and obstacles. Given some uncertainty or imperfect representation of the obstacles, the resulting algorithms perform probabilistic collision detection. Typically, these uncertainties are modeled using Gaussian distributions and stochastic algorithms are used to approximate the collision probability \cite{blackmore2006probabilistic,lambert2008fast}. In stochastic algorithms, a large number of sample evaluations are required to compute the accurate collision probability. If it can be assumed that the size of the objects is relatively small, the collision probability can be approximated using the probability at a single configuration corresponds to the mean of the probability distribution function (PDF), which provides a closed-form solution~\cite{du2011probabilistic}. This approximation is fast, but the computed probability cannot provide a bound, and can be either higher or lower than the actual probability, where the error increases as the object is bigger and has high-DOFs. For high-dimensional spaces, a common approach for checking collisions for imperfect or noisy objects is to perform the exact collision checking with scaled objects that enclose the potential object volumes~\cite{van2012lqg,Park:2012:ICAPS}. Prior approaches generally enlarge an object shape, which may correspond to a robot or an obstacle, to compute the space occupied by the object for a given standard deviation. This may correspond to a sphere~\cite{bry2011rapidly} or a ‘sigma hull’~\cite{lee2013sigma}. This approach provides an upper bounding volume for the given confidence level. However, the computed volume overestimates the probability and can be much bigger than the actual volume corresponds to the confidence level, which can cause failure of finding existing feasible trajectories in motion planning. Many other approaches have been proposed to perform probabilistic collision detection on point cloud data. Bae et al.~\cite{bae2009closed} presented a closed-form expression for the positional uncertainty of point clouds. Pan et al.~\cite{pan2011probabilistic} reformulate the probabilistic collision detection problem as a classification problem and compute per point collision probability. However, these approaches assume that the environment is static. Other techniques are based on broad phase data structures that handle large point clouds for realtime collision detection~\cite{pan2013real}. \subsection{Planning in Dynamic Environments} There is considerable literature on motion planning in dynamic scenes. In some applications, the future locations or trajectories of the obstacles are known. As a result, the time dimension is added to the configuration space of the robot and classical motion planning algorithms can be applied to the resulting state space~\cite{LaValle:2006}. In many scenarios, the future positions of the obstacles are not known. As a result, the planning problem is typically solved locally using reactive techniques such as dynamic windows or velocity obstacles~\cite{Fiorini:1998}, or assuming that the obstacle trajectories are known within a short horizon~\cite{Likhachev:2009}. Other methods use replanning algorithms, which interleave planning with execution. These methods include sampling-based planners~\cite{Hauser:safety,David:2002,SMP:2005}, grid searches~\cite{Koenig:2003:PBP,Likhachev05anytimedynamic}, or trajectory optimization~\cite{Park:2012:ICAPS}. Our formulation is based on optimization-based replanning, and we take into account smoothness and dynamic constraints. Applications that require high responsiveness use control-based approaches~\cite{haschke2008line,kroger2010online}, which can compute trajectories in realtime. They compute the robot trajectory in the Cartesian space, i.e. the workspace of the robot, according to the sensor data. However, the mapping from the Cartesian trajectory to the trajectory in the configuration space of high-DOF robots can be problematic. Furthermore, control-based approaches tend to compute less optimal robot trajectories as compared to the planning approaches that incorporate the estimation of the future obstacle poses. Planning algorithms can compute better robot trajectories in applications in which a good prediction about obstacle motions in a short horizon can be provided. \subsection{Planning under Uncertainties} The problem of motion planning under uncertainty, or belief space planning, has been an active area of research for the last few decades. The main goal is to plan a path for a robot in partially-observable state spaces. The underlying problem is formally defined using POMDPs (partially-observable Markov decision processes), which provide a mathematically rigorous and general approach for planning under uncertainty~\cite{kaelbling1998planning}. The resulting POMDP planners handle the uncertainty by reasoning over the {\em belief} space. A belief corresponds to the probability distribution over all possible states. However, The POMDP formulation is regarded as computationally intractable~\cite{papadimitriou1987complexity} for problems which are high-dimentional or have a large number of actions. Therefore, many efficient approximations~\cite{silver2010monte,kurniawati2013online,somani2013despot} and parallel techniques~\cite{shani2010evaluating,lee2013gpu} have been proposed to provide a better estimation of belief space. Most approaches for continuous state spaces use Gaussian belief spaces, which are estimated using Bayesian filters (e.g., Kalman filters)~\cite{leung2006planning,platt2010belief}. Algorithms using Gaussian belief spaces have also been proposed for the motion planning of high-DOF robots~\cite{van2012lqg,sun2015stochastic}, but they do not account for environment uncertainty or imperfect obstacle information. Instead, most planning algorithms handling environment uncertainty deal with issues arising from visual occlusions~\cite{missiuro2006adapting,guibas2010bounded,kahn2015active,charrow2015information}. In terms of dynamic environments, motion planning with uncertainty algorithms is mainly limited to 2D spaces~\cite{du2012robot,bai2015intention}, where the robots are modeled as circles, or to specialized applications such as people tracking~\cite{bandyopadhyay2009motion}. \begin{comment} The resulting POMDP planners handle the uncertainty by reasoning over the {\em belief} space~\cite{kaelbling1998planning}. A belief corresponds to the probability distribution over all possible states. However, POMDP is regarded as computationally intractable~\cite{papadimitriou1987complexity} for problems which are high-dimentional or have a large number of actions. Most practical solutions are limited to static scenes or low-dimensional state spaces~\cite{silver2010monte,kurniawati2013online,somani2013despot}. POMDP is a general motion planning framework that can compute the optimal solution under different types of uncertainties~\cite{kaelbling1998planning}. Most practical solutions are limited to static scenes or low-dimensional state spaces~\cite{silver2010monte,kurniawati2013online,somani2013despot}. The POMDP formulation considers every possible state to find the optimal solution, and is regarded as computationally intractable due to high dimensionality~\cite{papadimitriou1987complexity}. The exact state of the robot or the environment is typically not known. As a result, these approaches estimate the probability distributions of states, which are defined as the \textit{belief space}. Many efficient approximations~\cite{silver2010monte,kurniawati2013online,somani2013despot} and parallel techniques~\cite{shani2010evaluating,lee2013gpu} have been proposed to provide a better estimation of belief space. Many approximate POMDP formulations have been proposed for discretized state spaces~\cite{kurniawati2013online,somani2013despot}. \end{comment} \section{Probabilistic Collision Detection} \label{sec:pcc} In this section, we first introduce the notation and terminology used in the paper and present our probabilistic collision checking algorithm between the robot and the environment. \subsection{Notation and Assumptions} \label{subsec:notation} Our goal is to compute a collision probability between a high-DOF robot configuration and a given obstacle representation of dynamic environments, where the obstacle representation is a probability distribution which accounts uncertainties in the future obstacle motion prediction. For an articulated robot with $D$ one-dimensional joints, we represent a single robot configuration as $\mathbf q$, which is a vector composed from the joint values. The $D$-dimensional vector space of $\mathbf q$ is the configuration space $\mathcal{C}$ of the robot. We denote the collision-free subset of $\mathcal{C}$ as $\mathcal{C}_{free}$, and the other configurations corresponding to collisions as $\mathcal{C}_{obs}$. We assume that the robot consists of $J$ links $R_1,...,R_J$, where $J \leq D$. Furthermore, for each robot link $R_j$, we use multiple bounding volumes $B_{j1},...,B_{jK}$ to tightly enclose $R_j(\mathbf q)$ which corresponds to a robot configuration $\mathbf q$, i.e., \begin{equation} \label{eq:robot_bv} \begin{split} \forall j : R_j(\mathbf q) \subset \bigcup_{k=1}^K B_{jk}(\mathbf q) \,\,\textrm{for}\,\, (1 \le j \le J). \end{split} \end{equation} In our experiments, bounding spheres are automatically generated along the medial axis of each robot link. We represent $L$ obstacles in the environment as $O_l \, (1 \le l \le L)$, and assume that the obstacles undergo rigid motion. The configuration of these obstacles is specified based on geometric (shape) representation and their poses. As is the case for the robot, we use the bounding volumes $S_{l1},...,S_{lM}$ to enclose each obstacle $O_l$ in the environment: \begin{equation} \label{eq:obs_bv} \begin{split} \forall l : O_l \subset \bigcup_{m=1}^M S_{lm} \,\,\textrm{for}\,\, (1 \le l \le L). \end{split} \end{equation} For dynamic obstacles, we assume the predicted position of a bounding volume $S_{lm}$ at time $t$ is estimated as a Gaussian distribution $\mathcal{N} (\mathbf p_{lm}, \mathbf \Sigma_{lm})$, which will be described in Section~\ref{sec:environment}. \subsection{Probabilistic Collision Checking} \label{subsec:pcc} The collision probability between a robot configuration $\mathbf q_i$ with the environment at time $t_i$, $P(\mathbf q_i \in \mathcal{C}_{obs}(t_i))$ can be formulated as \begin{equation} \label{eq:colspace} \begin{split} P\left(\left(\bigcup_j \bigcup_k B_{jk}(\mathbf q_i)\right) \bigcap \left(\bigcup_l \bigcup_m S_{lm}(t_i) \right) \neq \emptyset \right). \end{split} \end{equation} We assume the robot links $R_j$ and obstacles $O_l$ are independent with each other link or obstacle, as their positions depend on corresponding joint values or obstacle states. Then (\ref{eq:colspace}) can be computed as \begin{align} \label{eq:colprob} &P(\mathbf q_i \in \mathcal C_{obs}(t_i)) =1-\prod_j\prod_l \overline{P_{col}(i,j,l) }, \end{align} where $P_{col}(i,j,l)$ is the collision probability between $R_j(\mathbf q_i)$ and obstacles $O_l(t_i)$. Since positions of bounding volumes $B_{jk}$ and $S_{lm}$ are determined by joint values or obstacle states of the corresponding robot link or obstacle, bounding volumes for the same object are dependant with each other, and $P_{col}(i,j,l)$ can be approximated as \begin{align} \label{eq:colprob2} &P_{col}(i,j,l)\approx\max_{k,m} P_{col}(i,j,k,l,m)\\ &P_{col}(i,j,k,l,m)=P(B_{jk}(\mathbf q_i)\cap S_{lm}(t_i) \neq \emptyset), \end{align} where $P_{col}(i,j,k,l,m)$ denotes the collision probability between $B_{jk}(\mathbf q_i)$ and $S_{lm}(t_i)$. \begin{figure}[t] \centering \includegraphics[trim=0in 0in 0in 1.0in, clip=true, width=0.6\linewidth]{max_prob.pdf} \caption{Approximation of probabilistic collision detection between a sphere obstacle of radius $r_2$ with a probability distribution $\mathcal{N} (\mathbf p_{lm}, \mathbf \Sigma_{lm})$ and a rigid sphere robot $B_{jk}(\mathbf q_i)$ centered at $\mathbf o_{jk}(\mathbf q_i)$ with radius $r_1$. It is computed as the product of the probability at $\mathbf x_{max}$ with the volume of the sphere with the radius computed as the sum of two radii, $V=\frac{4\pi}{3}(r_1+r_2)^3$.} \label{fig:prob_approx} \end{figure} Fig.~\ref{fig:prob_approx} illustrates how $ P_{col}(i,j,k,l,m)$ can be computed for $S_{lm}(t_i) \sim \mathcal{N} (\mathbf p_{lm}, \mathbf \Sigma_{lm})$. If we assume that the robot's bounding volume $B_{jk}(\mathbf q_i)$ is a sphere centered at $\mathbf o_{jk}(\mathbf q_i)$, similar to the environment bounding volume $S_{lm}$, and denote the radii of $B_{jk}$ and $S_{lm}$ as $r_1$ and $r_2$, respectively, the exact probability of collision between them is given as: \begin{equation} \label{eq:colobj} \begin{split} P_{col}(i,j,k,l,m)=\int_{\mathbf x}I(\mathbf x,\mathbf o_{jk}(\mathbf q_i))p(\mathbf x,\mathbf p_{lm},\mathbf \Sigma_{lm})d \mathbf x,\\ \end{split} \end{equation} where the indicator function $I(\mathbf x,\mathbf o)$ and the obstacle function $p(\mathbf x,\mathbf p,\mathbf \Sigma)$ are defined as, \begin{align} \label{eq:colobj1} I(\mathbf x,\mathbf o)=\left\{\begin{matrix} 1 & \textrm{if}\: \|\mathbf x - \mathbf o\| \leq (r_1+r_2) \\ 0 & \textrm{otherwise} \end{matrix}\right. \, \textrm{and} \end{align} \begin{align} \label{eq:colobj2} p(\mathbf x,\mathbf p,\mathbf \Sigma)=\frac{e^{-0.5(\mathbf x-\mathbf p)^T\mathbf \Sigma^{-1}(\mathbf x-\mathbf p)}}{\sqrt{(2\pi)^3\|\mathbf \Sigma\|}}, \end{align} respectively. It is known that there is no closed form solution for (\ref{eq:colobj}). Toit and Burdick approximate (\ref{eq:colobj}) as $V \cdot p(\mathbf o_{jk}(\mathbf q_i),\mathbf p_{lm},\mathbf \Sigma_{lm})$, where $V$ is the volume of sphere, i.e., $V=\frac{4\pi}{3}(r_1+r_2)^3$~\cite{du2011probabilistic}. However, this approximated probability can be either smaller or larger than the exact probability. If the covariance $\mathbf \Sigma_{lm}$ is small, the approximated probability can be much smaller than the exact probability. In order to compute an upper bound on the collision probability, we compute $\mathbf x_{max}$, the position has the maximum probability of $\mathcal{N} (\mathbf p_{lm}, \mathbf \Sigma_{lm})$ in $\mathbf B_{jk}(\mathbf q_i)$, and compute the upper bound of $P_{col}(i,j,k,l,m)$ as \begin{align} \label{eq:approx} P_{approx}(i,j,k,l,m) = V \cdot p(\mathbf x_{max},\mathbf p_{lm},\mathbf \Sigma_{lm}). \end{align} Although $\mathbf x_{max}$ has no closed-form solution, it can be computed efficiently. \begin{lemma} \label{thm:lemmamax} $\mathbf x_{max}$, the position has the maximum probability of $\mathcal{N} (\mathbf p_{lm}, \mathbf \Sigma_{lm})$ in $\mathbf B_{jk}(\mathbf q_i)$, is formulated as an one-dimensional search of a parameter $\lambda$, \begin{align} \label{eq:lemma} \mathbf x_{max}&=\left\{ \mathbf x|\|\mathbf x -\mathbf o_{jk}(\mathbf q_i)\|=(r_1+r_2) \,\textrm{and}\, \mathbf x \in \mathbf x(\lambda)\right\}, \textrm{where}\\ \mathbf x(\lambda)&=(\mathbf \Sigma_{lm}^{-1}+\lambda \mathbf I)^{-1}(\mathbf \Sigma_{lm}^{-1}\mathbf p_{lm}+\lambda \mathbf o_{jk}(\mathbf q_i)). \end{align} \end{lemma} \begin{proof} The problem of finding the position with the maximum probability in a convex region can be formulated as an optimization problem with a Lagrange multiplier $\lambda$~\cite{groetsch1984theory}, \begin{align} \label{eq:lemmaproof1} \mathbf x_{max} = \argmin_{\mathbf x} \left\{ (\mathbf x-\mathbf p_{lm})^T \mathbf \Sigma_{lm}^{-1}(\mathbf x - \mathbf p_{lm})+\lambda (\mathbf x - \mathbf o_{jk})^2\right\}. \end{align} The solution of (\ref{eq:lemmaproof1}) satisfies \begin{align} \label{eq:lemmaproof2} &\triangledown \left\{ (\mathbf x-\mathbf p_{lm})^T \mathbf \Sigma_{lm}^{-1}(\mathbf x - \mathbf p_{lm})+\lambda (\mathbf x - \mathbf o_{jk})^2\right\}=0, \end{align} and can be computed as \begin{align} \label{eq:lemmaproof3} &2\mathbf \Sigma_{lm}^{-1}(\mathbf x - \mathbf p_{lm})+2\lambda(\mathbf x-\mathbf o_{jk})=0\\ &\mathbf x = (\mathbf \Sigma_{lm}^{-1}+\lambda \mathbf I)^{-1})(\mathbf \Sigma_{lm}^{-1}\mathbf p_{lm} + \lambda \mathbf o_{jk}). \end{align} \end{proof} The approximated probability (\ref{eq:approx}) is guaranteed as an upper bound of the exact collision probability (\ref{eq:colobj}). \begin{theorem} \label{thm:approx} The approximated probability $P_{approx}(i,j,k,l,m)$ (\ref{eq:approx}) is always greater or equal to the exact collision probability $P_{col}(i,j,k,l,m)$ (\ref{eq:colobj}). \end{theorem} \begin{proof} $p(\mathbf x_{max},\mathbf p_{lm},\mathbf \Sigma_{lm}) \ge p(\mathbf x,\mathbf p_{lm},\mathbf \Sigma_{lm})$ for $\{\mathbf x|\|\mathbf x - \mathbf o_{jk}(\mathbf q_i)\| \leq (r_1+r_2)\}$ from Lemma~\ref{thm:lemmamax}. Therefore, \begin{align} \label{eq:theorem3} P_{approx}(i,j,k,l,m)&=V \cdot p(\mathbf x_{max},\mathbf p_{lm},\mathbf \Sigma_{lm}) \\ &= \int_{\mathbf x}I(\mathbf x,\mathbf o_{jk}(\mathbf q_i))d \mathbf x \cdot p(\mathbf x_{max},\mathbf p_{lm},\mathbf \Sigma_{lm})\\ &=\int_{\mathbf x}I(\mathbf x,\mathbf o_{jk}(\mathbf q_i))\cdot p(\mathbf x_{max},\mathbf p_{lm},\mathbf \Sigma_{lm})d \mathbf x \\ &\ge \int_{\mathbf x}I(\mathbf x,\mathbf o_{jk}(\mathbf q_i))\cdot p(\mathbf x,\mathbf p_{lm},\mathbf \Sigma_{lm})d \mathbf x \\ &=P_{col}(i,j,k,l,m). \end{align} \end{proof} \subsection{Comparisons with Other Algorithms} \begin{figure}[ht] \centering \subfloat[Case I] { \includegraphics[width=0.2\linewidth]{analysis_1.png} } \subfloat[Case II] { \includegraphics[width=0.2\linewidth]{analysis_2.png} } \begin{tabular}{|c|p{1.5cm}|p{1.5cm}|} \hline \multirow{2}{*}{Algorithms}& \multicolumn{2}{|c|}{Collision probability}\\ \cline{2-3} & \multicolumn{1}{c|}{Case I} & \multicolumn{1}{c|}{Case II}\\ \hline Numerical integration & \multicolumn{1}{r|}{0.09\%(O)} & \multicolumn{1}{r|}{1.72\%(X)}\\ \hline \begin{tabular}[x]{@{}c@{}}Enlarged bounding volumes\\ ($\delta_{CL} = 0.99$)~\cite{van2012lqg,Park:2012:ICAPS}\end{tabular}& \multicolumn{1}{r|}{100.00\%(X)} & \multicolumn{1}{r|}{ 100.00\%(X)} \\ \hline \begin{tabular}[x]{@{}c@{}}Approximation using\\ the center point PDF~\cite{du2011probabilistic}\end{tabular} & \multicolumn{1}{r|}{0.02\%(O)} & \multicolumn{1}{r|}{0.89\%(O)} \\ \hline Our approach & \multicolumn{1}{r|}{0.80\%(O)} & \multicolumn{1}{r|}{8.47\%(X)} \\ \hline \end{tabular} \caption{{\bf Comparison of approximated collision probabilities for feasible (Case I) and infeasible (Case II) scenarios for $\delta_{CL}=0.99$:} We compare the exact collision probability (computed using numerical integration) with approximated probabilities of 1) enlarged bounding volumes (blue contour)~\cite{van2012lqg,Park:2012:ICAPS}, 2) approximation using object center point (in green)~\cite{du2011probabilistic}, and 3) our approach that uses the maximum probability point (in red). Our approach guarantees to not underestimate the probability, while the approximated probability is close to the exact probability. } \label{fig:pcc_comparison} \end{figure} \begin{comment} A common approach for checking collisions for imperfect or noisy objects is to perform the exact collision checking with scaled objects that enclose the potential object volumes~\cite{van2012lqg,Park:2012:ICAPS}. Prior approaches generally enlarge an object shape, which may correspond to a robot or an obstacle, to compute the space occupied by the object for a given standard deviation. This may correspond to a sphere~\cite{bry2011rapidly} or a ‘sigma hull’~\cite{lee2013sigma}. This approach provides an upper bounding volume for the given confidence level. The computed volume overestimates the probability and can be much bigger than the actual volume corresponding to the confidence level. This overestimation occurs because the collision only occurs in the space that is occupied by robot, not the entire bounding volume. When there is a small overlap with the robot and the bounding volume, the exact probability of collision is usually much smaller than the one computed using the given sigma. If there are multiple obstacles in the environment, the enlarged bounding volume makes the approach conservative, and can prevent it from computing collision-free trajectories, which may still exist in the bounding volume. \end{comment} In Fig.~\ref{fig:pcc_comparison}, we illustrate two cases of the collision probability computation between a circle $B$ (in gray), and a point (in black) $\mathbf x$ which has uncertainties, $\mathbf x \sim (\mathbf p, \mathbf \Sigma)$, in 2D. We evaluate the exact collision probabilities using the numerical integration of the PDF. The collision probability of Case I is $0.09\%$, which is feasible with $\delta_{CL} = 0.99$, while the probability of Case II is $1.72\%$, which is infeasible. Contours represent the bounds for different confidence levels, where the blue contour corresponds to $\delta_{CL}=0.99$. In both cases, the blue contour intersects with $B$ and approaches that use enlarged bounding volumes~\cite{van2012lqg,Park:2012:ICAPS} determine the objects are collide, while the collision probability for Case I is $0.09\%$. Du Toit and Burdick~\cite{du2011probabilistic} used the probability of the center point (shown in green in Fig.~\ref{fig:pcc_comparison}) to compute a collision probability that is close to the actual value. However, their approach cannot guarantee upper bounds, and the approximated probability can significantly smaller from the actual probability if the covariance value is small. Case II in Fig.~\ref{fig:pcc_comparison} shows that the approximated probability is $0.89\%$, that satisfies the safety with the $\delta_{CL} = 0.99$, which is not true for the exact probability $1.72\%$. Unlike~\cite{du2011probabilistic}, we approximate the probability of the entire volume using the maximum probability value of a single point (shown in red in Fig.~\ref{fig:pcc_comparison}), as described in Section~\ref{subsec:pcc}. Our approach guarantees computation of the upper bound of collision probability, while the approximated probability is close to the exact probability than the enlarged bounding volume approaches. \section{Belief State Estimation} \label{sec:environment} In this section, we describe our approach for computing the current state $\mathbf p$ of environment obstacles, and use that to estimate the current belief state $\mathbf b_t$ and future states $\mathbf b_i \: (i>t)$, which are represented as the probability distributions. We construct or update the belief state of the environment $\mathbf b = (\mathbf p, \mathbf \Sigma)$ using means and covariances $\mathbf p_{ij}$ and $\mathbf\Sigma_{ij}$ of the poses of the existing bounding volumes $S_{ij}$. That is, $\mathbf p = \begin{bmatrix} \mathbf p_{11}^T & ... & \mathbf p_{lm}^T \end{bmatrix} ^T$ and $\mathbf\Sigma = \textrm{diag}(\mathbf\Sigma_{11},...,\mathbf\Sigma_{lm})$, where $\mathbf \Sigma$ is a block diagonal matrix of the covariances. \subsection{Environment State Model} \label{subsec:env_state} \begin{comment} For simplicity, we assume that we have the exact geometric representation and position of the static obstacles in the scene. However, we may not have the exact shape representation and poses for the dynamic obstacles. Rather, we assume that we have a few candidate shapes for the obstacles; e.g., an obstacle may correspond to a known shape such as a ball or a human arm. In our approach, the assumptions of known and unknown obstacles can be relaxed without a loss of generality. For example, we may not have the exact geometric representation of the static obstacle, and we can also deal with noisy representations of static obstacles. We assume that the environment state, corresponding to the dynamic obstacles, is provided using the depth sensors and converted into point cloud data, which are captured at a high frame rate. Point clouds corresponding to known static obstacles are removed in the background segmentation step, but the planner takes into account static obstacles for collision checking. It can be computationally inefficient to estimate and predict the states of dynamic obstacles that are represented using a large number of point clouds. Therefore, we use a reduced environment state representation that is defined in terms of the positions and velocities of the dynamic obstacles and utilize the predefined shape models for the dynamic obstacles. Each shape model for an obstacle in the model database is defined with multiple bounding volume shapes and their initial poses. \end{comment} In order to compute reliable obstacle motion trajectories in dynamic environments, first it is important to gather the state of obstacles using sensors. There is considerable work on pose recognition in humans~\cite{plagemann2010real,shotton2013real} or non-human objects~\cite{lepetit2005randomized} in computer vision and related areas. \begin{figure}[ht] \centering \includegraphics[width=0.5\textwidth]{environment_state.png} \caption{{\bf Environment belief state estimation for a human obstacle:} We approximate the point cloud from the sensor data using bounding volumes. The shape of bounding volumes are pre-known in the database, and belief states are defined on the probability distributions of bounding volume poses: (a) input point clouds (blue dots) (b) the bounding volumes (red spheres)with their mean positions (black dots) (c) the probabilistic distribution of mean positions. 0\% confidence level (black) to 100\% confidence level (white).} \label{fig:ho} \end{figure} We assume that a model database is given that consists of pre-defined shape models for each moving obstacle in the environment; e.g., an obstacle may correspond to a known shape such as a ball or a human arm. Furthermore, we are also given a bounding volume approximation of each such model. In particular, we use spheres as the underlying bounding volumes (Fig.~\ref{fig:ho}), as they provide an efficient approximation for computing the collision probability (see Section~\ref{subsec:pcc}). We segment out the background pixels correspond to the known static environments from the depth map, and generate a point cloud which is used to compute the best approximating environment state $\mathbf{p}^*$. It can be computationally inefficient to estimate and predict the states of dynamic obstacles that are represented using a large number of point clouds. Therefore, we use a reduced environment state representation that is defined in terms of the positions and velocities of the dynamic obstacles and utilize the predefined shape models for the dynamic obstacles. Each shape model for an obstacle in the model database is defined with multiple bounding volume shapes and their initial poses. For the input point cloud, we perform the object recognization at the beginning frame, then optimize $\mathbf{p}^*$ using the Ray-Constrained Iterative Closest Point~\cite{ganapathi12realtime} algorithm. Given the predefined shape model for each obstacle, ICP algorithm computes the best approximating environment state $\mathbf{p}^*$ for the input point clouds $\mathbf d_1,..., \mathbf d_n$. The likelihood of $\mathbf d_k$ for an environment state $\mathbf p$ is modeled as \begin{equation} P_{pc}(\mathbf d_k | \mathbf{p}) \propto \exp \left( {-} \frac{1}{2} \min_{i,j} \|S_{ij}-\mathbf d_k\| ^2 \right), \label{eq:probability_model} \end{equation} and the optimal environment state $\mathbf{p}^*$ that maximizes the likelihood of the each point cloud is computed with two additional constraints, represented as $C_1$ and $C_2$: \begin{align} \begin{split} &\mathbf p^* = \argmax_{\mathbf{p}} = \prod_k P_{pc}(\mathbf d_k | \mathbf{p}), \\ \text{subject to} \: C_1:& \forall (\mathbf p_{ij}, \mathbf p_{ik}) : (1 - \epsilon) \leq \frac{||\mathbf{p}_{ij} - \mathbf{p}_{ih}||}{c_{dist}({{ij}, {ih})}} \leq (1 + \epsilon) \\ C_2:& \forall \mathbf S_{ij} \forall \mathbf s_i : \textrm{proj}_{\mathbf s_i} (\mathbf S_{ij}) \subset \textrm{proj}_{\mathbf s_i}(\mathbf d_1,...,\mathbf d_n), \end{split} \label{eq:maximization}, \end{align} where ${c_{dist}({ij}, {ih})}$ is the distance between $\mathbf p_{ij}$ and $\mathbf p_{ih}$ of the predefined shape model, and $\textrm{proj}(\mathbf s_i)$ represents a projection to the 2D image space of depth sensor $\mathbf s_i$. Constraint $C_1$ corresponds to the length preserving constraint for the bounding volumes belong to the same object. $C_2$ ensures that the correct point clouds are generated for $\mathbf S_{ij}$ in view of all sensors $\mathbf s_i$. \subsection{Belief State Estimation and Prediction} \label{subsec:env_belief} The optimal solution $\mathbf p^*$ computed in Section~\ref{subsec:env_state} can have erros due to the sensors (e.g., point-cloud sensors) or poor sampling. Furthermore, obstacle motion can be sudden or abrupt and this can result in various uncertainties in the prediction of future motion. At each time $t$, we use the Kalman filter to estimate the position and velocity of the bounding volume $\mathbf S_{ij}$. We estimate the current belief states $\mathbf b_t = (\mathbf p_t, \Sigma_t)$ from the history of observed environment states $\mathbf{p}^*$, and then also predict the future state of the environment that is used for probabilistic collision checking. Its state at time $t$ is represented as \begin{align} (\mathbf{x}_{ij})_t = \begin{bmatrix} (\mathbf p_{ij})_t^T & (\mathbf{\dot{p}}_{ij})_t^T\end{bmatrix}^T, \end{align} where $(\mathbf p_{ij})_t$ is the position of $\mathbf S_{ij}$ at time $t$. We will omit subscript $_{ij}$ when we refer to a single obstacle. Using the Kalman filter, we estimate $\mathbf{x}_t$ as \begin{align} \mathbf{x}_t &= \mathbf{A} \mathbf{x}_{t-1} + \mathbf{B} \mathbf{u}_t + \mathbf{w}_t, \label{eq:KF_predict} \\ \mathbf{z}_t &= \mathbf{C} \mathbf{x}_t + \mathbf v_t, \label{eq:KF_udpate} \end{align} where the matrices are defined as \begin{equation} \mathbf A = \begin{bmatrix} I _{3\times3} & \Delta tI _{3\times3}\\ 0 & I _{3\times3} \end{bmatrix}, \mathbf{B} = \begin{bmatrix} I _{3\times3} \\ \Delta tI _{3\times3} \end{bmatrix}, \mathbf{C} = \begin{bmatrix} I _{3\times3} & 0 \end{bmatrix}, \end{equation} and $\mathbf w_t$ and $\mathbf v_t$ are the process noise and observation noise, respectively. $\mathbf{z}_t$ is an observation that corresponds to $\mathbf p^*$. Although we cannot directly control the environment, we compute an hypothetical input $\mathbf{u}_t$ that is used to preserve the distances between the bounding volumes belong to the same object in the predicted result. During the estimation, if the distance of an object $\mathbf S_{ij}$ from another object $\mathbf S_{ih}$ exceeds the distance in the predefined shape model, we compute an appropriate value for $\mathbf{u}_t$ to preserve the initial distance. In order to preserve the initial distance $\|(\mathbf p_{ij})_0-(\mathbf p_{ih})_0\|$, we pull the $\mathbf S_{ij}$'s position $(\mathbf p_{ij})_t$ toward $\mathbf S_{ih}$'s position $(\mathbf p_{ih})_t$ using \begin{equation} \label{eq:length_control} \mathbf{u}_{t} = \left( (\mathbf{p}_{ih})_t - (\mathbf{p}_{ij})_t \right) \left( 1 - \frac{\|(\mathbf p_{ij})_0-(\mathbf p_{ih})_0\|}{\|(\mathbf p_{ij})_t-(\mathbf p_{ih})_t\|} \right) . \end{equation} \subsection{Spatial and Temporal Uncertainties in Belief State} \label{subsec:uncertainties} \begin{figure}[ht] \centering \subfloat[] { \includegraphics[width=0.22\linewidth]{uncertainty1.png} } \subfloat[] { \includegraphics[width=0.22\linewidth]{uncertainty2_4.png} \label{fig:sensor_error_single} } \subfloat[] { \includegraphics[width=0.22\linewidth]{uncertainty3.png} } \subfloat[] { \includegraphics[width=0.22\linewidth]{uncertainty4.png} } \caption{{\bf Spatial uncertainty:} (a) Sphere obstacle and its point cloud samples from a depth sensor. (b) Probability distribution of a sphere center state $\mathbf p$ for a single point cloud $\mathbf d_k$. (c) Probability distribution of $\mathbf p$ for a partially visible obstacle. (d) Probability distribution of $\mathbf p$ for a fully visible obstacle. } \label{fig:sensor_error} \end{figure} During the environment state estimation, spatial uncertainty or errors arise from the resolution of the sensor. It is known that the depth sensor error can be modeled as Gaussian distributions around each point $\mathbf d_k$~\cite{nguyen2012modeling}. We assume that the center of distribution is $\mathbf d_k$ itself and the covariance is isotropic and can be represented as $\sigma_{s}^2 I_{3\times3}$. Due to the sensor error, the optimal environment state $\mathbf p^*$ computed from (\ref{eq:maximization}) may differ from the true environment state $\mathbf p^{t}$. We derive the equation for the observation noise $\mathbf v_t$ in (\ref{eq:KF_udpate}) for an environment state computed using (\ref{eq:maximization}). For simplicity, we assume the environment has only one sphere with radius $r$ and its optimal state is computed from point clouds (Fig.~\ref{fig:sensor_error}(a)). For a single obstacle case, the optimization equation (\ref{eq:maximization}) can be written as \begin{align} P(\mathbf p) \propto \max_{\mathbf{p}} \quad & \prod_k \exp \left( {-} \frac{1}{2} \left( ||\mathbf{p} - \mathbf{d}_k|| - r \right)^2 \right) \notag \\ = &\prod_k P( \mathbf{p} | \mathbf{d}_k ). \label{eq:maximization2} \end{align} Here, $P( \mathbf{p} | \mathbf{d}_k )$ corresponds to the spherical probability distribution that represents the highest value at distance $r$. If $r \gg \sigma_{s}$, it can be approximated near $\mathbf{p}^{t}$ as a Gaussian distribution as shown in Fig.~\ref{fig:sensor_error}(b), \begin{equation} \label{eq:noise_approx} P( \mathbf{p} | \mathbf{d}_k ) \sim \mathcal{N}( \mathbf{p}^{t}, \sigma_{s}^2 \mathbf{n}^{t} \times (\mathbf{n}^{t}) ^T ), \end{equation} where $\mathbf{n}_k = (\mathbf{p}^t - \mathbf{d}_k) / ||\mathbf{p}^t - \mathbf{d}_k||$. $P(\mathbf p)$ is a product of these spherical probability distributions (\ref{eq:noise_approx}) for different point cloud $\mathbf d_k$, and it corresponds to another Gaussian distribution $\mathcal{N}( \mathbf{p}_t, \Sigma^* )$. Therefore, the observation error $\mathbf{v}_t$ can be represented as: \begin{align} \mathbf{v}_t \sim P(\mathbf p) - \mathbf p^t = \mathcal{N}( \mathbf{0}, \Sigma^* ). \end{align} If we are given more samples from the sensor and there is less sensor error, the error distribution becomes more centralized. Temporal uncertainty arises due to discretization of the time domain, which corresponds to approximating the velocity of dynamic obstacle using forward differencing method. Let $\mathbf{x}(t)$ be the obstacle position at time $t$. By Taylor expansion, we obtain \begin{equation} \mathbf{x}(t + \Delta t) = \mathbf{x}(t) + \dot{\mathbf{x}}(t) \Delta t + \frac{1}{2} \ddot{\mathbf{x}}(t) \Delta t^2 + O(\Delta t^3), \end{equation} and \begin{equation} \dot{\mathbf{x}}(t) \approx \frac{\mathbf{x}(t + \Delta t) - \mathbf{x}(t)}{\Delta t} + \frac{1}{2} \ddot{\mathbf{x}}(t) \Delta t + O(\Delta t^2). \label{eq:taylor_expansion} \end{equation} From the history of past environment states, we compute $\ddot{\mathbf{x}}(t)$ of each object and its covariance $\Sigma_a (t)$. Based on Equation (\ref{eq:taylor_expansion}), we get the process error $\mathbf w_t$ as \begin{equation} \mathbf w_t \sim \mathcal{N} \left( \mathbf{0}, \left[ \begin{array}{c|c} \frac{1}{4} (\Delta t)^4 \Sigma_a(t) & \mathbf{0} \\ \hline \mathbf{0} & \frac{1}{4} (\Delta t)^2 \Sigma_a(t) \end{array} \right] \right), \end{equation} which is used in our estimation framework (Section~\ref{subsec:env_belief}) to compute the environment belief states. These estimated belief states are used for collision probability computation (Section~\ref{subsec:pcc}). \begin{comment} \subsection{Analysis} The observation error, $\mathbf v_t$ is computed as the product of the noise of each point cloud belongs to the same bounding volume, as shown in (\ref{eq:maximization2}). The empirically measured error of Kinect is known as $0.014m$ for the point cloud which is $3m$ far from the sensor~\cite{nguyen2012modeling}. Furthermore, the uncertainty of the position $\mathbf p_{ij}$ is inversely proportional to the number of point cloud datasets, $\mathbf d_k$, which are used for the computation of $\mathbf p_{ij}$. However, there is another source of the observation error. We assume that there is a pre-defined shape in the model database, which exactly matches with the environment obstacles. This is not the case in many applications. For example, we may use one generic model for all human-like obstacles. There can be errors between the predefined shape and the actual object, which is part of the observation error $\mathbf v$, that affects the accuracy of the final result. For each iteration of the trajectory optimization algorithm, the collision probability is computed for all pairs of the robot bounding volumes $\mathbf O_{jk}$ and the environment objects $\mathbf S_{lm}$. Therefore, the performance of the planning algorithm is a linear function function of the number of overlaps between the robot bounding volumes and the environment objects. In order to predict future belief states, we use an independent prediction model for each object $\mathbf S_{lm}$ with a length preserving constraint, given in Equation (\ref{eq:length_control}), instead of the joint prediction model which requires $\mathcal O(m^2)$ computations, where $m$ is the number of objects, i.e., the dimension of the environment state. This allows efficient linear time computation of the future belief state. \end{comment} \section{Space-Time Trajectory Optimization} \label{sec:optimization} In this section, we describe our motion planning algorithm based on probabilistic collision detection (Section~\ref{sec:pcc}) and environment belief state estimation (Section~\ref{sec:environment}). Fig.~\ref{fig:architecture} highlights various components of our planning algorithm. The pseudo-code description is given in Algorithm 1 for a single planning step $\delta T$. \begin{figure}[t] \centering \includegraphics[trim=0in 0in 0in 0in, clip=true, width=0.9\linewidth]{architecture.pdf} \caption{{\bf Trajectory Planning:} We highlight various components of our algorithm. These include belief space estimation from the sensor data and environment description, probabilistic collision checking, and trajectory optimization.} \label{fig:architecture} \end{figure} \begin{algorithm}[t] \caption{$\mathbf Q^*=$PlanWithEnvUncertainty($\mathbf Q$, $\{\mathbf d_k\}, t_i$) \\: Compute the optimal robot trajectory $\mathbf Q^*$ during the planning step $\Delta T$ for the environment point clouds $\{\mathbf d_k\}$ at time $t_i$} \label{alg:pseudo} \begin{algorithmic}[1] \REQUIRE initial trajectory $\mathbf Q$, environment point clouds $\{\mathbf d\}$, time $t_i$ \ENSURE Optimal robot trajectory $\mathbf Q^*$ for time step $\Delta T$ \STATE $\mathbf p_i$ = EnvironmentStateComputation($\{\mathbf d_k\}$) // {\em compute the environment state of dynamic obstacles} \FOR {$k \in \{i, ..., i + \Delta T\}$} \STATE $\mathbf B_k$ = BeliefStateEstimation($\mathbf B_0, ..., \mathbf B_{k-1}$, $\mathbf p_i$) //{\em estimate the current and future belief states} \ENDFOR \WHILE {elapsed time $<\Delta T$} \STATE $P$=ProbCollisionChecking($\mathbf Q,\{\mathbf B_i,...,\mathbf B_{i+\Delta T}\}$) // {\em perform probabilistic collision detection} \STATE $\mathbf Q^*$=Optimize($\mathbf Q,P$) // {\em compute the optimal trajectory for high-DOF robot} \ENDWHILE \end{algorithmic} \end{algorithm} As described in Section.~\ref{sec:environment}, we construct or update the belief state of the environment $\mathbf b = (\mathbf p, \Sigma)$, which is the probability distribution of the poses of the existing bounding volumes. Then we predict the future belief state of the environment. We define the time-space domain $\mathcal X$, which adds a time dimension to the configuration space, i.e., $\mathcal X = \mathcal C \times T$. The robot's trajectory, $\mathbf q(t)$, is represented as a function of time from the start configuration $\mathbf q_s$ to the goal configuration $\mathbf q_g$. It is represented using the matrix $\mathbf Q$, \begin{equation} \label{eq:x} \mathbf Q=\begin{bmatrix} \mathbf q_s & \mathbf q_1&...& \mathbf q_{n-1} & \mathbf q_g\\t_0 & t_1&...&t_{n-1}&t_n\end{bmatrix}, \end{equation} which corresponds to $n+1$ configurations at discretized keyframes, $t_i=i \Delta_T$, which have a fixed interval $\Delta_T$. We denote the $i$-th column of $\mathbf Q$ as $\mathbf x_i=\begin{bmatrix}\mathbf q_i^T & t_i \end{bmatrix}^T$. Given the initial and goal positions for motion planning, our planner computes a locally optimal trajectory based on the objective function defined for the duration of the trajectory and also based on other constraints (e.g., smoothness). We use incremental trajectory optimization, which repeatedly refines a motion trajectory using an optimization formulation~\cite{Park:2012:ICAPS}. The planner initializes the robot trajectory $\mathbf Q$ as a smooth trajectory of predefined length $T$ between the start and goal configurations $\mathbf q_s$ and $\mathbf q_g$, i.e., \begin{align} \label{eq:traj_init} \mathbf Q=\argmin_{Q}\sum_{i=1}^{n-1}\|\mathbf q_{i-1}-2\mathbf q_i+\mathbf q_{i+1}\|^2. \end{align} The trajectory is refined during every planning step $\Delta T$ based on various constraints, and we add collision probability constraints which is based on the probabilistic collision described in Section~\ref{sec:pcc} as the collision-free constraints. We define the collision probability constraint of feasible robot trajectories based on the following probability computation formulation (shown as $P()$): \begin{equation} \label{eq:pcol} \begin{split} \forall \mathbf x_i : P(\mathbf q_i \in \mathcal{C}_{obs}(t_i)) < 1 - \delta_{CL}. \end{split} \end{equation} The computed trajectories that satisfy (\ref{eq:pcol}) guarantee that the probability of collision with the obstacles is bounded by the confidence level $\delta_{CL}$, i.e. the probability that a computed trajectory has no collision is higher than $\delta_{CL}$. Use of a higher confidence level computes safer, but more conservative trajectories. The use of a lower confidence level increases the success rate of planning, but also increases the probability of collision. The objective function for trajectory optimization at time $t_k$ can be expressed as the sum of trajectory smoothness cost, and collision constraint costs for dynamic uncertain obstacles and static known obstacles, \begin{small} \begin{equation} \label{eq:opt} \begin{split} f(\mathbf Q)=\min_{Q}&\sum_{i=k+m}^{n}\left(\|\mathbf q_{i-1}-2\mathbf q_i+\mathbf q_{i+1}\|^2+C_{static}(\mathbf Q_i) \right)\\ +&\sum_{i=k+m}^{k+2m} \textrm{max}(P(\mathbf q_i \in \mathcal C_{obs}(\mathbf x_i))-(1-\delta_{CL}),0), \end{split} \end{equation} \end{small} where $m$ is the number of time steps in a planning time step $\Delta T$. Furthermore, we can add additional kinematic or dynamic constraints that the robot has to satisfy, such as bounds on the joint position, velocity limits or robot balancing constraints. These can be satisfied in the trajectory optimization framework by formulating them as a constraint optimization problem, with these specific constraints. Unlike the previous optimization-based planning approaches~\cite{Park:2012:ICAPS,Zucker:IJRR:2012} which maintain and cannot change the predefined trajectory duration for the computed trajectory, our planner can adjust the duration of trajectory $T$ to avoid collisions with the dynamic obstacles. When the trajectory planning starts from $\mathbf t_i$ ($\mathbf t_i$ can be different from $\mathbf t_s$ due to replanning) and if the computed trajectory $\mathbf Q$ violates the collision probability constraint (\ref{eq:pcol}) at time $j$, i.e., $P(\mathbf q_j \in \mathcal C_{obs}(t_j)) \ge \delta_{CL}$, we repeatedly add a new time step $\mathbf x_{new}$ before $\mathbf x_{j}$ and rescale the trajectory from $\left[\mathbf t_i,...,\mathbf t_{j-1}\right]$ to $\left[\mathbf t_i,...,\mathbf t_{j-1}, \mathbf t_{new}\right]$, until $\mathbf x_{new}$ is collision-free. Moreover, the next planning step starts from $\mathbf x_{new}$. This formulation of adjusting the trajectory duration allows the planner to slow the robot down when it cannot find a safe trajectory for the previous trajectory duration due to the dynamic obstacles. The optimization problem in (\ref{eq:opt}) is solved using Covariant Hamiltonian trajectory optimization~\cite{Zucker:IJRR:2012}, which preserves the trajectory smoothness during optimization. If the optimization algorithm converges, our algorithm computes the optimal trajectory, \begin{align} \mathbf Q^*=\argmin_{\mathbf Q}f(\mathbf Q), \end{align} which provides a probabilistic collision guarantee with the given confidence level $\delta_{CL}$, for the time step $\Delta T$. \section{Results} \label{sec:result} In this section, we describe our implementation and highlight the performance of our planning algorithm on different benchmark scenarios. We tested our planning algorithm in simulated environments with models of a 6-DOF Universal Robot UR5 (Fig.~\ref{fig:experiment1}(a)(b)) and a 7-DOF KUKA IIWA robot arm (Fig.~\ref{fig:experiment1}(c)(d)). The environments have some complex static obstacles such as tools or furniture in a room. The dynamic obstacle is a human, and we assume that the robot operates in close proximity to the human, however, the human does not intend to interact with the robot. We use a Kinect as the depth sensor, which can represent a human as 25-30k point clouds. We use a commodity PC for the planner, and use OpenMP to compute the probabilistic collision checking in parallel using multi-core CPUs. \subsection{Experimental Results} \begin{table*}[ht] \centering \resizebox{\linewidth}{!}{ \begin{tabular}{|c|c|c|c|c|c|c|} \hline Benchmark & Robot DOF & \begin{tabular}[x]{@{}c@{}}\# of Robot\\Bounding\\ Volumes\end{tabular} & \begin{tabular}[x]{@{}c@{}}\# of Samples\\ in Point Cloud\end{tabular} & \begin{tabular}[x]{@{}c@{}}Environment \\ State DOF\end{tabular} & \begin{tabular}[x]{@{}c@{}}Confidence \\ Level\end{tabular} & \begin{tabular}[x]{@{}c@{}}Average \\ Planning\\ Time\end{tabular} \\ \hline Prediction 1 & 6 (UR5) & 30 & 33,000 & 336 & $0.95$ & 138.83 ms \\ \hline Prediction 2 & 6 (UR5) & 30 & 29,500 & 336 & $0.95$ & 115.55 ms \\ \hline \begin{tabular}[x]{@{}c@{}}Time-Space\\Search 1\end{tabular} & 7 (IIWA) & 35 & 35,000 & 336 & $0.95$ & 771.44 ms \\ \hline \begin{tabular}[x]{@{}c@{}}Time-Space\\Search 2\end{tabular} & 7 (IIWA) & 35 & 35,500 & 336 & $0.95$ & 552.64 ms \\ \hline \begin{tabular}[x]{@{}c@{}}Exact\\Collision\\Checking\end{tabular} & 7 (IIWA) & 35 & 35,000 & 336 & $1.0$ & 154.99 ms \\ \hline Sensing Noise 1 & 7 (IIWA) & 35 & 35,000 & 336 & $0.95$ & 720.52 ms \\ \hline Sensing Noise 2 & 7 (IIWA) & 35 & 35,000 & 336 & $0.99$ & 846.11 ms \\ \hline \end{tabular} } \caption{{\bf Complexity and planning results in our benchmarks:} We use two different robot models UR5 and IIWA, in our benchmarks. We highlight the complexity of each benchmark in terms of robot bounding volumes, the number of point cloud samples, DOF of the environment state, and the confidence level used for probabilistic collision detection. We compute the average planning time for each benchmark on a multi-core CPU.} \label{table:performance} \end{table*} \begin{figure}[t] \centering \subfloat { \includegraphics[width=0.22\linewidth]{result1.png} } \subfloat { \includegraphics[width=0.22\linewidth]{result2.png} } \subfloat { \includegraphics[width=0.22\linewidth]{result3.png} } \subfloat { \includegraphics[width=0.22\linewidth]{result4.png} } \caption{{\bf Robot Trajectory with Dynamic Human Obstacles:} Static obstacles are shown in green, the estimated current and future human bounding volumes are shown in blue and red, respectively. (a) When a human is approaching the 6-DOF robot arm (UR5), our planner changes its trajectory to avoid potential future collisions. (b) When a standing human only stretchs out an arm, our shape model-based prediction prevents unnecessary reactive motions, which results a better robot trajectory than the prediction using simple extrapolations. (c) A collision-free computed trajectory that avoids collisions with the obstacle and the environment. (d) The robot adjusts its speed or waits if there is no feasible path to the goal position due to the dynamic obstacles.} \label{fig:experiment1} \end{figure} Table~\ref{table:performance} presents the complexity of the benchmarks and the performance of our planning results. Our first benchmark tests our shape model-based environment belief state prediction. When a human is dashing onto the robot at a fast speed, the robot is aware of the potential collision with the predicted future human position and changes its trajectory (Fig.~\ref{fig:experiment1}(a)). However, if a human in standing only stretchs out an arm toward the robot, even if the velocity of the arm is fast, the model-based prediction prevents unnecessary reactive motions, which is different from the prediction models with constant velocity or acceleration extrapolations (Fig.~\ref{fig:experiment1}(b)). The second benchmark set shows our planner's collision-free path computation in the space-time domain. The planner computes a collision-free trajectory that avoids collision with the obstacles and the environments (Fig.~\ref{fig:experiment1}(c)). If there is no feasible solution due to dynamic obstacles, the planner adjusts its speed or waits until it finds a solution (Fig.~\ref{fig:experiment1}(d)). \begin{figure}[ht] \centering \subfloat { \includegraphics[width=0.3\linewidth]{confidence_noise1.png} } \subfloat { \includegraphics[width=0.3\linewidth]{confidence_noise2.png} } \subfloat { \includegraphics[width=0.3\linewidth]{confidence_noise3.png} } \caption{{\bf Robot Trajectory with Different Confidence and Noise Levels:} (a) A trajectory with exact collision checking for zero-noise obstacles. (b) A trajectory with $\delta_{CL}=0.95$ and $\mathbf v_t=0.005 I_{3 \times 3}$. (c) A trajectory with $\delta_{CL}=0.99$ and $\mathbf v_t=0.05 I_{3 \times 3}$. } \label{fig:experiment2} \end{figure} Fig.~\ref{fig:experiment2} shows a robot trajectory with different confidence levels and sensor noises. If the obstacle states are assumed as exact, the robot can follow the shortest and smoothest trajectory that is close to the obstacle (Fig.~\ref{fig:experiment2}(a)). However, as the noise of the environment state or expected confidence level becomes higher, the computed robot trajectories become longer and less smooth to avoid potential collision with the obstacles (Fig.~\ref{fig:experiment2}(b)-(c)). \subsection{Comparison with Other Algorithms} \begin{table}[ht] \centering \begin{tabular}{|c|c|c|c|} \hline Algorithms & \begin{tabular}[x]{@{}c@{}}Number of\\Collisions\end{tabular} & \begin{tabular}[x]{@{}c@{}}Trajectory\\Duration (sec)\end{tabular} & \begin{tabular}[x]{@{}c@{}}Trajectory\\Length (m)\end{tabular}\\ \hline \begin{tabular}[x]{@{}c@{}}Enlarged bounding\\ volumes~\cite{van2012lqg,Park:2012:ICAPS}\end{tabular} & 0.02 & 10.47 & 2.32\\ \hline \begin{tabular}[x]{@{}c@{}}Approximation using\\ the center point PDF~\cite{du2011probabilistic}\end{tabular} & 0.43 & 6.62 & 1.52\\ \hline Our approach & 0.03 & 7.16 & 1.87 \\ \hline \end{tabular} \caption{{\bf Planning results of different probabilistic collision detection algorithms:} Our probabilistic collision detection approach shows a high safety as the approach using enlarged bounding volumes, while computes efficient trajectories.} \label{table:performance2} \end{table} In order to compare our algorithm with other probabilistic collision detection algorithms, we plan trajectories using the different probabilistic collision detection algorithms. We choose different initial and goal configurations for each trials, and compute trajectories with $\delta_{CL}=0.99$. The trajectory durations are initialized to 5 seconds. We applied the same recorded human motion that stretches an arm that blocks the initial robot trajectory to the planning, but each trial has a different small perturbation of the human obstacle position that corresponds to the environment uncertainties. We measure the number of collisions, as well as the durations and lengths of the computed trajectories for planners with three different probabilistic collision detection algorithms. The averages of 100 trials are shown in Table~\ref{table:performance2}. The enlarged bounding volumes have less collisions, but the durations and lengths of the computed trajectories are longer than other approaches, since the overestimated collision probability makes the planner to compute trajectories which are unnecessarily far apart from the obstacles, or to wait when there is a feasible trajectory. On the other hand, the approximating approach that uses the probability of the object center point underestimate the collision probability and causes several collisions in the planned trajectories. Our approach shows a similar safety with the approach using enlarged bounding volumes, while it also computes efficient trajectories that have shorter trajectory durations and lengths. \section{Conclusions and Future Work} We present a novel algorithm for trajectory planning for high-DOF robots in dynamic, uncertain environments. This include new methods for belief space estimation and probabilistic collision detection. Our approach is quite fast, and works well in our simulated results where it can compute collision-free paths with high confidence level. Our probabilistic collision detection computes tighter upper bounds of the collision probability as compared to prior approaches. We highlight the performance of our planner on different benchmarks with human obstacles for two robot models. To the best of our knowledge, that can handle high-DOF robots in dynamic environment with imperfect obstacle representations. Our approach has some limitations. Some of the assumptions used in belief space estimation in terms of Gaussian distribution and Kalman filter may not hold. Moreover, we may have a pre-defined shape representation of the obstacle. The trajectory optimization may get stuck at a local minima and may not converge to a global optimal solution. There are many avenues for future work. Our approach only takes into account the imperfect information about the moving obstacles. In particular, we assume that a good point-cloud sample of the obstacles is given for probabilistic collision checking. Our current approach does not account for control errors or sensor errors, which are rather common with the controllers and sensors.
7a5a816724631a02e454be2b11ac72ef6965f30f
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The classical problem to determine the rational points of the modular curve $X_0(p)$, when $p$ is a prime number, has been solved by Mazur in his paper \cite{Maz76} in 1975. In his work, Mazur proved, for $p>37$, that the only rational points of the curve $X_0(p)$ are cusps and CM points. In the same paper Mazur also refers to the interesting case of the modular curve $X_0^+(p)$. This curve is the quotient of $X_0(p)$ by the Atkin-Lehner involution $w_p$. See Galbraith \cite[Section 2]{Gal99}. For these curves a similar result is expected: for sufficiently large $p$ the only rational points of the curve $X_0^+(p)$ are cusps and CM points. The genus $g$ of $X_0^+(p)$ growth with $p$, so for a fixed $g$ we have only finitely many curves. The cases $g$ equal to $0$ or $1$ are well known. If $g$ is equal to $0$, then $X_0^+(p)$ is isomorphic to $\PP^1$ and there are infinitely many rational points. If $g$ is equal to~$1$, then $X_0^+(p)$ is an elliptic curve whose Mordell-Weil group is torsion-free and has rank~$1$. Let us focus on the case $g>1$. Galbraith in his papers \cite{Gal96} and \cite{Gal99} describes a method to find explicit equations for the curves $X_0^+(p)$. If $X_0^+(p)$ is hyperelliptic he uses an \emph{ad hoc} technique. We recall that by \cite{HH96} that the curve $X_0^+(p)$ is hyperelliptic if and only if $g=2$. If $X_0^+(p)$ is not hyperelliptic, Galbraith uses the canonical embedding $\varphi\colon X_0^+(p) \hookrightarrow \PP^{g-1}$ to find equations for the curves. He finds explicit equations for all curves $X_0^+(p)$ with $g\le 5$. Here we present the following result. \begin{result} We use the method of Galbraith to find explicit equations for the canonical model of all the modular curves $X_0^+(p)$ with genus $g=6$ or $7$. \end{result} Equations for the thirteen curves of genus $6$ or $7$ are listed in Section \ref{sec:listeq}. The equations that we find have very small coefficients and have good reduction for each prime number $\ell\neq p$. The technique used here play an important role in \cite{Mer15} in finding explicit equations defining the modular curves $X_{ns}^+(p)$. These curves are the modular curves associated to the normalizer of a nonsplit Cartan subgroup of the general linear group over a finite field with $p$ elements. Few is still known about them and they are harder to study with respect to modular curves $X_0^+(p)$ considered here. Galbraith uses the equations of the curves $X_0^+(p)$ to look for rational points $P$ up to some bounds on the naive projective height $H$. See Silverman \cite[Chapter VIII, Section 5]{Sil86} for more details about the naive height on projective space. Using a projection onto a plane model, possibly singular, we check that if $X_0^+(p)$ has genus $g=6$ or $7$, then it has no rational points $P$, except the cusps and the CM points, for which $H(P)\le 10^4$. When the jacobian variety $J_0^+(p)$ of $X_0^+(p)$ has an isogeny factor over $\QQ$ of dimension 1, we improve this bound significantly. In particular we prove the following result. \begin{result} For the primes $p=163,197,229,269$ and $359$, the only rational points $P$ on $X_0^+(p)$ of naive height $H(P)\le 10^{10000}$ are cusps or CM points. \end{result} \section{The method to get explicit equations} \label{sec:algoeq} In this section we show how to get explicit equations of the canonical model for modular curves. The final output is a list of quadrics with very small integer coefficients. Let $\Gammait$ be a congruence subgroup of $\SL_2(\ZZ)$, let $\cfs_2(\Gammait)$ be the $\CC$-vector space of the cusp forms of weight $2$ with respect to $\Gammait$, and let $f_1,\ldots,f_g$ be a $\CC$-basis of $\cfs_2(\Gammait)$ where $g$ is the genus of the modular curve $X(\Gammait)$. If \mbox{$g>2$}, we use the canonical embedding to get the canonical model for $X(\Gammait)$. We know that $X(\Gammait)$ is isomorphic as compact Riemann surfaces to $\Gammait \backslash \HH^*$, where $\HH^*$ is the extended complex upper half plane. We also know that $\cfs_2(\Gammait)$ is isomorphic to the $\CC$-vector space of holomorphic differentials $\Omega^1(X(\Gammait))$ with respect to the map $f(\tau)\mapsto f(\tau)d\tau$. See Diamond and Shurman \cite{DS05} for more details about these results. Using these isomorphisms, the canonical embedding can be expressed in the following way \begin{align*} \varphi\colon X(\Gammait) & \to \PP^{g-1}(\CC) \\ \Gammait\tau & \mapsto (f_1(\tau):\ldots:f_g(\tau)), \end{align*} where $\tau\in\HH^*$. The Enriques-Petri Theorem (see Griffith, Harris \cite[Chapter 4, Section 3, pag. 535]{GH78} or Saint-Donat \cite{Sd73}), states that the canonical model of a complete nonsingular non-hyperelliptic curve is entirely cut out by quadrics and cubics. Moreover, it is cut out by quadrics if it is neither trigonal, nor a quintic plane curve with genus exactly~$6$. When $X(\Gammait)$ can be defined over $\QQ$, we can look for equations defined over~$\QQ$. The Enriques-Petri Theorem is proved over algebraically closed fields, but its proof can be suitable modified to work over $\QQ$. Alternatively, if we find equations defined over $\QQ$, then we can check by MAGMA that their zero locus $Z$ is an algebraic curve with the same genus as $X(\Gammait)$. An application of the Hurwitz genus formula for genus $g>1$, to the morphism $\varphi\colon X(\Gammait)\to Z$ implies that $\varphi$ is an isomorphism. Finding equations for $Z$ is equivalent to finding generators of the homogeneous ideal $\mathcal{I}$ defining it in $\PP^{g-1}$. Let $\mathcal{I}_d$ be the set of homogeneous elements of $\mathcal{I}$ of degree exactly $d$. We explain how to find generators of $\mathcal{I}$ that belong to $\mathcal{I}_d$ for a fixed $d$. By the Enriques-Petri Theorem and by Hasegawa and Hashimoto \cite{HH96}, we know that if $g>2$, the ideal $\mathcal{I}$ is generated by elements in $\mathcal{I}_d$ for $d=2,3$. Now, we fix the degree $d$ and we suppose that we know the first $m$ Fourier coefficients of the $q$-expansions of a basis $\mathcal{B}=\{f_1,\dots,f_g\}$ of $\Omega^1(X(\Gammait))$, where $m>d(2g-2)$. This condition on $m$ guarantees that if we have a polynomial $F$ with rational coefficients and $g$ unknowns such that $F(f_1,\ldots,f_g)\equiv 0 \pmod{q^{m+1}}$, then $F(f_1,\ldots,f_g)=0$. See \cite[Section 2.1, Lemma 2.2, pag 1329]{BGGP05} for more details. We also assume that the Fourier coefficients of the basis $\mathcal{B}$ are algebraic integers. There is a number field $K$, of degree~$D$ over~$\QQ$, which contains all the coefficients of all the elements of $\mathcal{B}$. Moreover, if the Fourier coefficients are algebraic integers, they have integer coordinates with respect to a suitable chosen basis of $K$ over $\QQ$. If $a_n(f_i)$ is the $n$-th Fourier coefficient of $f_i$, we denote the $k$-th coordinate of $a_n(f_i)$ in $K$ by $c(i,n,k)\in\ZZ$, where $k=1,\ldots,D$. We can associate to $\mathcal{B}$ a matrix $M$ whose integer entries are the coordinates of the Fourier coefficients where each row $i$ corresponds to an element $f_i$ of $\mathcal{B}$. An explicit way to set the entries of $M=(u_{ij})$ is to choose $u_{ij}=c(i,n,k)$, where $j=(n-1)D+k,$ for $i=1,\ldots,g$, for $n=1,\dots,m$ and for $k=1,\ldots,D$. The matrix $M$ has $g$ rows and $mD$ columns, where $m$ is the number of the first Fourier coefficients used and $g$ is the cardinality of the basis. Since $m>d(2g-2)$ and $g>0$, we always have $m\geq g$, hence the rank of $M$ is $g$. To find equations defining $X(\Gammait)$ we compute all the monomials $F_j$ of degree $d$ where the indeterminates are the elements of $\mathcal{B}$. The elements of a $\ZZ$-basis of the space $S$ of the solutions of the homogeneous linear system in the unknowns $F_j$ are the coefficients of the desired equations. We are interested in reducing the size of the coefficients of these equations and minimizing the number of primes $\ell$ such that the model has bad reduction modulo~$\ell$. To reduce the size of the coefficients, we apply the LLL-algorithm first to $M$ and then to the $\ZZ$-basis of $S$. We know that if the rank of $M$ modulo $\ell$ is less than $g$, then the canonical model of the curve that we find, is singular modulo $\ell$. We say that $M$ is \emph{optimal}, if the rank of $M$ modulo $\ell$ is exactly $g$ for each prime~$\ell$. In Algorithm~\ref{alg:badprimes} below, within a more general setting, we describe how to modify $\mathcal{B}$ to make $M$ optimal. Now, we describe how to find the primes $\ell$ such that the canonical model has bad reduction. Let $\mathcal{I}_{\text{jac}}$ be the ideal generated by the polynomials defining the curve and by all the determinants of order $g-2$ of the jacobian matrix of the curve. We compute the elimination ideals $\mathcal{J}_i:=\mathcal{I}_{\text{jac}}\cap \QQ[x_i]$, for $i=1,\ldots,g$. If $\mathcal{J}_i\neq 0$, it turns out to be generated by $\lambda x_i^n$ with $\lambda,n\in\ZZ_{>0}$. The curve has bad reduction modulo any prime~$\ell$ such that $\ell\mid \lambda$. \subsection{Algorithm} Let $g$ and $m$ be positive integers such that $g\leq m$ and let $v_1,\ldots,v_g\in\ZZ^m$ be linearly independent vectors over $\QQ$. We describe a method to find a basis over $\ZZ$ of the lattice $L:=\Span_{\QQ}\{v_1,\ldots,v_g\} \cap \ZZ^m$. Let $L':=\Span_{\ZZ}\{v_1,\ldots,v_g\}$, so $L'$ is a subgroup of $L$, and let $J:=[L:L']$. We want to modify $L'$ and its basis until we have $J=1$ and so $L'=L$. \begin{lemma} \label{lem:pdep} Let the notation as above and let $p$ be a prime number. We have that $v_1,\ldots,v_g$ are linearly dependent in $\ZZ^m/p\ZZ^m$ if and only if $p\mid [L:L']$. \end{lemma} \begin{proof} We have that $p\mid [L:L']=\#(L/L')$ if and only if there is an element of $L/L'$ of order exactly $p$. This is equivalent to have an element $v\in L\setminus L'$ such that $pv\in L'$. The condition $v\in L$ implies that there are $\alpha_1,\ldots,\alpha_g\in\QQ$ such that $v=\alpha_1 v_1+\ldots+\alpha_g v_g$. The condition $pv\in L'$ implies that $p\alpha_1,\ldots,p\alpha_g$ are integers. The independence over $\QQ$ implies that $p\alpha_1,\ldots,p\alpha_g$ are not all zero. Finally, the condition $v\notin L'$ implies that $p\alpha_1,\ldots,p\alpha_g$ are not all zero modulo $p$. \end{proof} Let $M\in\ZZ^{m\times g}$ be the matrix whose columns are the vectors $v_1,\ldots,v_g$. It follows from Lemma \ref{lem:pdep} above that if a prime $p$ divides the index $J$, then $p$ divides the determinant $\Delta$ of any $g\times g$ submatrix of $M$. \begin{algoritmo} \label{alg:badprimes} \begin{description} \item[Step 0] Choose a $g\times g$ submatrix of $M$ with nonzero determinant~$\Delta$. Set $\mathcal{P}:=\{\text{prime numbers dividing $\Delta$}\}$. Go to Step~1. \item[Step 1] If $\mathcal{P}=\varnothing$ the algorithm terminates. If $\mathcal{P}\neq \varnothing$ go to Step~2. \item[Step 2] Let $p\in \mathcal{P}$. If $v_1,\ldots,v_g$ are linearly dependent in $\ZZ^m/p\ZZ^m$, we have that $p\mid J$, and go to Step~3. Else set $\mathcal{P}:=\mathcal{P}\setminus \{p\}$ and go to Step~1. \item[Step 3] If $v_1,\ldots,v_g$ are linearly dependent in $\ZZ^m/p\ZZ^m$, up to reordering the $v_i$'s, there are, $\alpha_2,\ldots,\alpha_g\in\ZZ$ such that $v_1+ \alpha_2 v_2+\ldots+\alpha_g v_g=pv$. We replace $v_1$ by $v=\frac{1}{p}(v_1+ \alpha_2 v_2+\ldots+\alpha_g v_g)\in L\setminus L'$ and go to Step~2. \end{description} \end{algoritmo} \begin{remark} To reduce the number of primes to check, one can apply Gauss elimination to make $\Delta$ minimal. \end{remark} \begin{remark} The algorithm terminates in finitely many steps. In fact the substitution $v_1\mapsto \frac{1}{p}(v_1+ \alpha_2 v_2+\ldots+\alpha_g v_g)$ implies that \[ \Delta=\det(v_1,\ldots,v_g)\mapsto \det\left(\frac{1}{p}\left(v_1+\sum_{i=2}^g \alpha_i v_i\right),v_2,\ldots,v_g\right)=\frac{1}{p}\det(v_1,\ldots,v_g)=\frac{\Delta}{p}. \] Hence, after finitely many iterations of the algorithm the rank of $M$ modulo $p$ is $g$ for each prime $p$. \end{remark} \section{Computing the expected rational points} We have a simple moduli interpretation of certain rational points of the modular curves $X_0^+(p)$. The \emph{expected} rational points on $X_0^+(p)$ are the points corresponding to the unique cusp or to elliptic curves with complex multiplication such that $p$ is split or ramified inside the endomorphism ring of the elliptic curve itself. We call a rational point \emph{exceptional} if it is not one of the expected ones. Here we describe a way to find numerically the expected rational points on a modular curve $X_0^+(p)$, for $p$ an odd prime. We assume to know the first $m$ Fourier coefficients of a basis $f_1,\ldots,f_g$ of $\Omega^1(X_0^+(p))$, where $g$ is the genus of the modular curve. It is well known, see Stark \cite{Sta66}, that the orders in imaginary quadratic number fields with class number $1$ are the ones with discriminant \[ \Delta=-3,-4,-7,-8,-11,-12,-16,-19,-27,-28,-43,-67,-163. \] The corresponding orders can be written in the form $\ZZ+\ZZ\tau$, where \[ \tau=\begin{cases} \frac{1+i\sqrt{|\Delta|}}{2} & \text{if }\Delta \text{ is odd}, \\ \frac{i\sqrt{|\Delta|}}{2} & \text{if }\Delta \text{ is even}. \end{cases} \] Let $E$ be an elliptic curve with complex multiplication such that the discriminant $\Delta_E$ of its endomorphism ring $\mathcal{O}_E$ is a class number 1 discriminant. This means that $\Delta_E$ is in the list above. If $\Delta_E$ is a square modulo $p$ we can associate to $E$ a point on $X_0^+(p)$, see Galbraith \cite{Gal99}. The unique cusp is always rational. Let $\{1,\tau_E\}$ be the basis of the order $\mathcal{O}_E$ such that $\tau_E$ is defined as above. There is a suitable element $\hat{\tau}$ in the $\SL_2(\ZZ)$-orbit of $\tau_E$ in $\HH$ such that $P=(f_1(\hat{\tau}):\ldots:f_g(\hat{\tau}))$ is a rational point for the modular curve. If we know $\hat{\tau}$, we can evaluate $f_1(\hat{\tau}),\ldots,f_g(\hat{\tau})$ numerically using the $q$-expansions. If $m$ is large enough, it is quite easy recognize the rational coordinates of $P$. To compute the coordinates of the cusp it is enough take $\hat{\tau}=i\infty$, this means to set $q=0$ in the $q$-expansions. Now we explain how to compute $\hat{\tau}$ for the CM points. We recall the moduli interpretation of the points of $X_0^+(p)$. Let $E'$ be an elliptic curve and let $C$ be a cyclic subgroup of order $p$ of the $p$-torsion subgroup $E'[p]$. We know that a point on $X_0^+(p)$ is an unordered pair $\{(E',C),(E'/C,E'[p]/C)\}$, where $E'/C$ is the unique elliptic curve, up to isomorphism, that is the image of the unique isogeny of $E'$ with kernel~$C$. Over $\CC$ this is $\{(\CC/\Lambda,\frac{1}{p}\ZZ+\Lambda),(\CC/\Lambda',\frac{1}{p}\ZZ+\Lambda')\}$ for some $\tau\in\HH$ such that $\Lambda=\ZZ+\tau\ZZ$ and $\Lambda'=\ZZ+\frac{-1}{p\tau}\ZZ$. If $\Delta_E$ is a square modulo~$p$, we know there is a principal prime ideal $\mathfrak{p}$ of $\mathcal{O}_E$ such that $(p)=\mathfrak{p}\bar{\mathfrak{p}}$. Let $\alpha$ and $\bar{\alpha}$ be generators of $\mathfrak{p}$ and $\bar{\mathfrak{p}}$ respectively. The cyclic subgroups $C$ and $E'[p]/C$ are the kernels of the multiplication by $\alpha$ and $\bar{\alpha}$. Let $(E',C)=(\CC/\Lambda_E,\frac{1}{\bar{\alpha}}\ZZ+\Lambda_E)$, where $\Lambda_E=\ZZ+\tau_E\ZZ$. The point $\{(\CC/\Lambda_E,\frac{1}{\bar{\alpha}}\ZZ+\Lambda_E),(\CC/\Lambda_E,\frac{1}{\alpha}\ZZ+\Lambda_E)\}$ is rational on~$X_0^+(p)$. Now, we want to find $\hat{\tau}$ such that $(\CC/\Lambda_E,\frac{1}{\bar{\alpha}}\ZZ+\Lambda_E)=(\CC/\hat{\Lambda},\frac{1}{p}\ZZ+\hat{\Lambda})$, where $\hat{\Lambda}=\ZZ+\hat{\tau}\ZZ$. Let $c,d\in\ZZ$ such that $\alpha=c\tau_E+d$. Then we can find $a,b\in\ZZ$ such that $ad-bc=1$ and so we have a matrix $\hat{\gamma}=\begin{pmatrix} a & b \\ c & d \end{pmatrix}\in\SL_2(\ZZ)$. The transformation $\hat{\tau}=\hat{\gamma} \tau_E$ correspond to the isogeny $z+\Lambda_E\mapsto \frac{z}{\alpha}+\hat{\Lambda}$ for every $z+\Lambda_E\in\CC/\Lambda_E$. Hence $\frac{1}{\bar{\alpha}}+\Lambda_E\mapsto \frac{1}{\alpha\bar{\alpha}}+\hat{\Lambda}=\frac{1}{p}+\hat{\Lambda}$ and the group $\frac{1}{\bar{\alpha}}\ZZ+\Lambda_E$ maps to $\frac{1}{p}\ZZ+\hat{\Lambda}$ and we are done. \section{Computing the modular parametrization} \label{sec:compEmap} In this section we assume to know the first $m$ Fourier coefficients of the modular forms involved, where $m$ is "large enough" for our purposes. Let $p$ be a prime number and let $J_0(p)$ be the jacobian variety of the modular curve $X_0(p)$. If there is an isogeny factor $E$ over $\QQ$ of dimension $1$ of $J_0(p)$, we know there is a non-constant morphism $\phi\colon X_0(p)\to E$ defined over $\QQ$. Let $f$ be the normalized eigenform in $\cfs_2(\Gammait_0(p))$ associated to the isogeny class of $E$. Let $\mathds{T}$ be the Hecke algebra over $\ZZ$ and let $I_f=\{T\in\mathds{T}: Tf=0\}$. We know that if the Fourier coefficients of $f$ belong to $\QQ$, the associated abelian variety $J_0(p)/I_f J_0(p)$ is an elliptic curve called \emph{optimal curve} (or \emph{strong Weil curve}). See Diamond and Shurman \cite[Sections 6.5 and 6.6]{DS05} for more details about these topics. If $E=J_0(p)/I_f J_0(p)$, then we call $\phi$ the \emph{modular parametrization with respect to $X_0(p)$}. The degree of the morphism $\phi$ is called the \emph{modular degree}. Using the identification of $X_0(p)$ with $\Gammait_0(p)\backslash\HH^*$, we can write \begin{align*} \phi\colon \Gammait_0(p)\backslash\HH^* &\to E(\CC) \\ \Gammait_0(p)\tau &\mapsto \phi(\tau)=(\phi_x(\tau),\phi_y(\tau)), \end{align*} where $\tau\in\HH^*$ and $\phi_x(\tau)$ and $\phi_y(\tau)$ are Fourier series with rational coefficients in the indeterminate $q=e^{2\pi i \tau}$. \begin{lemma} \label{lem:factorparmod} If $E$ is an elliptic curve over $\QQ$ of conductor a prime $p$ and with negative sign of the functional equation of the associated L-function, then the modular parametrization $\phi\colon X_0(p)\to E$ factors as $\phi=\phi_+\circ\pi_p$, where $\pi_p\colon X_0(p)\to X_0^+(p)$ is the natural projection and $\phi_+\colon X_0^+(p)\to E$ is the modular parametrization with respect to $X_0^+(p)$. Moreover $\phi_+$ is defined over $\QQ$. \end{lemma} \begin{proof} Just apply Galois Theory to the associated function fields and use invariance under the Atkin-Lehner involution of the differentials of $X_0^+(p)$. \end{proof} Let $X_0^+(p)$ a modular curve of genus $g$ with an isogeny factor $E$ over $\QQ$ of dimension $1$ of the jacobian variety $J_0^+(p)$. We denote by $f_1,\ldots,f_g$ some $g$ linearly independent newforms of $\cfs_2(\Gammait_0(p))$ with eigenvalue $+1$ with respect to the Atkin-Lehner operator~$w_p$. This is a basis of $\cfs_2(\Gammait^+_0(p))$. The modular parametrization $\phi_+\colon X_0^+(p)\to E$ is locally given by four homogeneous polynomials $p_x,q_x,p_y,q_y\in\ZZ[x_1,\ldots,x_g]$, such that $p_x$ and $q_x$ have the same degree, $p_y$ and $q_y$ have the same degree and \begin{align} \label{eq:sistema} \begin{sistema} \phi_x(\tau)= \frac{p_x(f_1(\tau),\ldots,f_g(\tau))}{q_x(f_1(\tau),\ldots,f_g(\tau))} \\ \phi_y(\tau)= \frac{p_y(f_1(\tau),\ldots,f_g(\tau))}{q_y(f_1(\tau),\ldots,f_g(\tau))}, \end{sistema} \end{align} for every $\tau$ representative of $\Gammait_0(p)\tau$ chosen in a suitable open set of $X_0(p)$. To find these polynomials explicitly we use some linear algebra. The main difficulty is that we don't know exactly the degrees of the polynomials $p_x,q_x,p_y,q_y$, but we know the degree of the morphism $\phi_+$. In Cremona \cite{Cre92} or in the LMFDB online database \cite{lmfdb} we can find the value of $\deg\phi$ and we have $\deg\phi_+=\frac{1}{2}\deg\phi$. For increasing degrees from $1$ to $\deg{\phi_+}$, we apply the following procedure to look for nontrivial relations. We rewrite the first equation as \[ \label{eq:ctrimg} p_x(f_1(\tau),\ldots,f_g(\tau))- q_x(f_1(\tau),\ldots,f_g(\tau))\phi_x(\tau)=0. \] Let $r=\begin{pmatrix} g +d_x-1 \\ d_x \end{pmatrix}$ be the number of monomials of degree $d_x$. Let $\mathcal{I}=\{F_1,\ldots,F_r\}$ be the set of the monomials obtained as product of $d_x$ elements of the basis $f_1,\ldots,f_g$, where repetitions are allowed, and let $\mathcal{J}=\{G_1,\ldots,G_r\}$ be the set with the same elements of $\mathcal{I}$ multiplied by $\phi_x$. Since the Fourier coefficients of $\phi_x,f_1,\ldots,f_g$ are algebraic numbers, the coefficients of the elements of $\mathcal{I}$ and $\mathcal{J}$ are also algebraic numbers and there is a number field $K_f$ of finite degree over $\QQ$, which contains all these Fourier coefficients. Multiplying the elements of $\mathcal{I}$ and $\mathcal{J}$ by a suitable algebraic number, we can assume that the Fourier coefficients belong to the ring of integers of $K_f$ and that they have integer coordinates with respect to a suitable basis of~$K_f$. Hence, there are $\alpha_1,\ldots,\alpha_r,\beta_1,\ldots,\beta_r\in\ZZ$ such that \[ \begin{sistema} \alpha_1 F_1+\ldots+\alpha_r F_r+\beta_1 G_1+\ldots+\beta_r G_r=0, \\ \alpha_1 F_1+\ldots+\alpha_r F_r\neq 0, \\ \beta_1 G_1+\ldots+\beta_r G_r\neq 0. \end{sistema} \] These $\alpha_i$'s and $\beta_i$'s are the coefficients of $p_x$ and $q_x$ respectively. Different choices of $\alpha_1,\ldots,\alpha_r,\beta_1,\ldots,\beta_r$ correspond to the same map defined on different open sets. The same method is applied to the second equation of (\ref{eq:sistema}) to find the coefficients of $p_y$ and $q_y$. \begin{remark} We find the optimal curve with conductor $p$ in Cremona's tables in \cite{Cre92} or in the LMFDB online database \cite{lmfdb}. We use PARI to compute the $q$-expansions of the two components $\phi_x(\tau)$ and $\phi_y(\tau)$ of the modular parametrization of $X_0(p)$. \end{remark} \section{An estimation on heights} \label{sec:upbnd} Let $n$ be a positive integer and let $X$ be a complete nonsingular curve defined over $\QQ$ in $\PP^{n-1}$ and with an elliptic curve $E$ as isogeny factor over $\QQ$ of its jacobian variety. In this section we suppose to know a Weierstrass equation of $E$, an explicit formula for the morphism \mbox{$\phi\colon X \to E$} defined over $\QQ$ and generators for the Mordell-Weil group of $E$. Writing $\phi$ in more explicit terms, we have \[ \phi(x_1,\ldots,x_n)=\left(\frac{p_x(x_1,\ldots,x_n)}{q_x(x_1,\ldots,x_n)}, \frac{p_y(x_1,\ldots,x_n)}{q_y(x_1,\ldots,x_n)}\right), \] where $p_x,q_x$ are homogeneous polynomials with integer coefficients of degree $d_x$ and $p_y,q_y$ are homogeneous polynomials with integer coefficients of degree $d_y$. Let $r$ be the number of monomials of degree $d_x$ and let $\alpha_i$ and $\beta_i$, for \mbox{$i=1,\ldots, r$}, be the coefficients of $p_x$ and $q_x$ respectively. Moreover, let \mbox{$\alpha:=\log\max\left\{\sum_i\left|\alpha_i\right|,\sum_i\left|\beta_i\right|\right\}$} a constant that we use in the Proposition \ref{pro:upbnd} below. In the paper \cite{Sil90}, Silverman gives an estimate of the difference between the canonical height and the Weil height on an elliptic curve. We use the Theorem~1.1 of \cite{Sil90} to prove the following proposition. \begin{proposizione} \label{pro:upbnd} Let the notation be as above and let $Q\in\PP^{n-1}$. Let $H(Q)$ be the naive height of $Q$ and let $h(Q):=\log H(Q)$. If $Q\in X(\QQ)$, then $\phi(Q)\in E(\QQ)$ and \[ \hat{h}(\phi(Q))\le \mu(E)+1.07+\frac{1}{2}\Big(\alpha+d_x h(Q)\Big), \] where the quantity $\mu(E)$ is defined in Theorem 1.1 of \cite{Sil90}, the function $\hat{h}$ is the canonical height on $E$, the number $d_x$ is the degree of the homogeneous polynomials $p_x$ and $q_x$ defined above and $\alpha$ is the constant defined above. \end{proposizione} \begin{proof} Let $Q$ be a rational point on $X$. Let $(x_1:\ldots:x_n)$ its coprime integers coordinates in $\PP^{n-1}$. We denote by $P$ the image of this rational point under $\phi$, hence we have $\phi(x_1,\ldots,x_n)=P=(P_x,P_y)$, where \[ P_x=\frac{p_x(x_1,\dots,x_n)}{q_x(x_1,\dots,x_n)}. \] If $|x_i|\leq H(Q)$, by the definition of height we have \[ H(P_x)\leq H(Q)^{d_x} \max\left\{\sum_i\left|\alpha_i\right|,\sum_i\left|\beta_i\right|\right\}. \] Taking the logarithm in the previous inequality we get $h(P_x)\leq\alpha+d_x h(Q)$. Now, using the second inequality in Theorem 1.1 of \cite{Sil90}, we are done. \end{proof} \begin{remark} Since $\phi$ is defined over $\QQ$, every rational point on the curve $X$ must go in a rational point on the elliptic curve $E$. Therefore, to find rational points on $X$, it is enough to search in the preimage $\phi^{-1}(P)$ letting $P$ run over the rational points of~$E$. \end{remark} \begin{remark} In all our cases the Mordell-Weil group has rank $1$ and is torsion-free, so it is generated by one point. One can find a generator in Cremona's tables in \cite{Cre92} or in the LMFDB online database \cite{lmfdb}. \end{remark} \begin{remark} \label{rem:upbnd} If the Mordell-Weil group of $E$ has rank $1$ with generator $P_0$ and is torsion-free, every rational point $P$ of $E$ has the form $P=kP_0$ for some $k\in\ZZ$. In this case we have \[ \hat{h}(P)=\hat{h}(kP_0)=k^2\hat{h}(P_0)\leq\mu(E)+1.07+\frac{1}{2}\Big(\alpha+d_x h(Q)\Big), \] by properties of the canonical height. See \cite[Chapter VIII, Section 9]{Sil86} for more details about the canonical height. \end{remark} \section{Tables} \label{sec:listeq} In \ref{subsec:eq} we list the equations for the canonical model of the modular curves $X_0^+(p)$ for primes $p$ such that the genus $g$ is $6$ or $7$. We also list the expected rational points. The models all have good reduction modulo every prime $\ell\neq p$. In \ref{subsec:modpar} we give an explicit morphism $\phi_+\colon X_0^+(p)\to E$, whenever the jacobian variety $J_0^+(p)$ of $X_0^+(p)$ admits a $1$-dimensional isogeny factor $E$ over $\QQ$. \subsection{Equations}\label{subsec:eq} Here we list the genus, the equations of the canonical model and the expected rational points with the corresponding discriminants. \begin{itemize} \item Curve $X_0^+(163)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} x_1x_5+x_2x_3+x_2x_4-x_2x_5+x_2x_6=0 \\ x_1^2-x_1x_2+x_1x_5-x_2x_5+x_3^2+x_3x_4=0 \\ x_2x_4-x_2x_5+x_3x_5=0 \\ -x_1^2+x_1x_2-x_1x_4+x_2x_4 + x_3x_6=0 \\ x_1x_3+x_1x_5+x_1x_6+x_2x_3-x_2x_5+x_4x_5=0 \\ -x_1^2+x_1x_2+x_1x_3+x_1x_4-x_1x_5+x_1x_6+x_2x_5-x_3^2+x_4^2+x_5x_6=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:1) & \text{cusp} & (0:1:1:0:1:0) & -12 \\ (24:10:13:15:-50:42) & -163 & (0:1:0:0:0:0) & -11 \\ (1:0:1:-2:0:-1) & -67 & (0:0:1:-1:0:0) & -8 \\ (1:1:2:-2:2:0) & -28 & (1:1:0:0:0:0) & -7 \\ (1:1:0:1:1:-1) & -27 & (2:-1:3:-4:-1:-2) & -3 \\ (0:0:0:0:1:0) & -19 & & \\ \bottomrule \end{array} \] \item Curve $X_0^+(193)$. Genus $g=7$. Equations of the canonical model in $\PP^{6}$ \[ \begin{sistema} -x_1^2-x_1x_4-x_1x_6-x_1x_7+x_2x_7+x_3^2+x_3x_4=0 \\ x_1x_2+x_1x_4-x_1x_5+x_1x_7-x_2x_3-x_2x_4-x_2x_7+x_3x_5=0 \\ x_1x_2-x_1x_5-x_2x_7+x_3x_6=0 \\ -x_1x_2-x_1x_3+2x_1x_5+x_1x_6+2x_2x_3+2x_2x_4-x_3^2+x_4^2+x_4x_5=0 \\ x_1^2-x_1x_2+2x_1x_4+x_1x_5+2x_1x_6+x_1x_7-x_2x_7-x_3^2+x_4^2+x_4x_6=0 \\ x_1x_3-x_1x_4-x_1x_5-2x_1x_6-x_1x_7-2x_2x_3-2x_2x_4+2x_2x_7+x_3^2-x_4^2+ x_4x_7=0 \\ x_1^2+x_1x_3+x_1x_4-2x_1x_5+x_1x_7-x_2x_3-3x_2x_4-x_2x_5-x_2x_7-x_4^2+x_5^2=0 \\ -x_1^2+2x_1x_2-x_1x_4-x_1x_5-2x_1x_6-x_1x_7-x_2x_3-x_2x_4+x_2x_7+x_3^2-x_4^2+ x_5x_6=0 \\ -x_1x_3+x_1x_5+x_1x_6+2x_2x_3+2x_2x_4-x_2x_7-x_3^2+x_4^2+x_5x_7=0 \\ -x_1x_4-x_1x_5-x_1x_7+x_2x_3-x_2x_4-x_2x_6+2x_2x_7-x_3x_7-x_4^2+x_6^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (1:0:-1:0:0:0:0) & -12 \\ (1:1:1:-1:2:0:-1) & -67 & (0:1:0:0:0:1:0) & -8 \\ (0:0:0:1:-1:-1:1) & -43 & (0:1:0:0:0:0:0) & -7 \\ (0:1:2:-2:0:0:0) & -28 & (1:0:0:1:0:-1:-1) & -4 \\ (0:1:0:0:1:0:0) & -27 & (3:2:3:0:2:0:0) & -3 \\ (1:0:0:-1:0:-1:1) & -16 & & \\ \bottomrule \end{array} \] \item Curve $X_0^+(197)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} -x_1x_2+x_1x_4-x_1x_5-x_1x_6+x_2x_3=0 \\ x_1^2-x_1x_3-x_1x_4+x_1x_5+x_2x_4=0 \\ 2x_1x_2-2x_1x_3-2x_1x_4+x_1x_5+x_1x_6-x_2^2-x_2x_6+x_3^2+x_3x_4+x_4^2=0 \\ -x_1^2+x_1x_2+x_1x_3+x_1x_6-x_2^2-x_2x_6+x_4x_5=0 \\ x_1x_2-x_1x_3-x_1x_4+x_3x_5+x_4x_6=0 \\ -x_1x_2+x_1x_3+x_1x_4-x_1x_6+x_2x_5-x_2x_6-x_3x_5+x_3x_6+x_5^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:1) & \text{cusp} & (1:1:1:0:0:0) & -19 \\ (1:-1:3:2:6:-6) & -163 & (1:0:0:0:-1:1) & -16 \\ (1:1:1:1:0:1) & -43 & (1:0:0:1:0:1) & -7 \\ (1:0:2:-1:0:-1) & -28 & (1:0:2:0:1:-1) & -4 \\ \bottomrule \end{array} \] \item Curve $X_0^+(211)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} -x_1^2-x_1x_4-x_1x_5+x_2x_3-x_2x_6+x_3x_4=0 \\ -x_1x_3-x_1x_4-x_1x_5+x_2x_3-x_2x_6+x_3x_5=0 \\ x_1^2-x_1x_2+x_1x_3+x_1x_4-x_2x_3+x_3x_6=0 \\ 2x_1^2-2x_1x_2+x_1x_3+3x_1x_4+x_1x_5-2x_2x_3-x_2x_4+2x_2x_6+x_4^2=0 \\ x_1x_3+x_1x_4+2x_1x_5-x_1x_6-2x_2x_3-x_2x_5+2x_2x_6+x_4x_5=0 \\ x_1^2-2x_1x_2+x_1x_3+2x_1x_4+x_1x_5-x_1x_6-2x_2x_3-x_2x_4+x_2x_5+2x_2x_6-x_4x_6+ x_5^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:1) & \text{cusp} & (0:1:1:0:0:1) & -12 \\ (1:0:-1:-1:1:-1) & -67 & (0:0:1:0:0:0) & -8 \\ (2:1:2:0:0:-2) & -28 & (0:1:0:0:0:0) & -7 \\ (1:0:1:-1:-1:-1) & -27 & (2:-3:-1:-2:4:1) & -3 \\ \bottomrule \end{array} \] \item Curve $X_0^+(223)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} -x_1^2+x_1x_4-x_2x_3-x_2x_4+x_2x_5=0 \\ x_1^2+2x_1x_3-x_1x_4+x_1x_5+x_2^2+2x_2x_4+x_2x_6+x_3^2=0 \\ -x_1^2-x_1x_2-x_1x_3+x_1x_4-x_2x_4+x_2x_6+x_3x_4=0 \\ x_1^2-x_1x_2+2x_1x_5-x_1x_6+x_2^2+x_2x_4+2x_2x_6+x_3x_5=0 \\ -x_1^2-x_1x_2+x_1x_3+x_1x_4+x_2x_6-x_3x_6+x_4x_5=0 \\ -2x_1^2-2x_1x_2+x_1x_3+3x_1x_4+2x_1x_5+x_1x_6+x_2^2+4x_2x_6-x_3x_6-x_4^2-x_4x_6+ x_5^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:1) & \text{cusp} & (0:1:1:-1:0:0) & -12 \\ (2:-2:2:3:6:7) & -163 & (0:0:0:1:0:-1) & -11 \\ (1:0:-2:1:0:1) & -67 & (2:-3:-3:-1:-6:2) & -3 \\ (1:0:0:1:0:1) & -27 & & \\ \bottomrule \end{array} \] \item Curve $X_0^+(229)$. Genus $g=7$. Equations of the canonical model in $\PP^{6}$ \[ \begin{sistema} -x_1x_5+x_2x_4-x_2x_6+x_2x_7=0 \\ x_1x_2-x_1x_3-x_1x_4-x_1x_5+x_2x_3+x_2x_4-x_2x_6+x_3^2+x_3x_4=0 \\ -x_1x_5+x_2^2+x_2x_3+x_2x_4-x_2x_5-x_2x_6+x_3x_5=0 \\ x_1x_2-x_1x_3-x_1x_4-x_1x_6-x_2^2+x_2x_4+x_3^2+x_3x_6=0 \\ x_1x_2+x_1x_3-x_1x_6-x_1x_7-x_2^2-x_2x_3+x_2x_6+x_3x_7=0 \\ -2x_1x_2+x_1x_4+x_1x_5+x_1x_6+x_1x_7-x_2x_3-x_2x_4+x_2x_6-x_3^2+x_4^2=0 \\ -x_1x_3+x_1x_6-x_2^2-x_2x_3-x_2x_4+x_2x_5+x_4x_5=0 \\ -x_1x_2+x_1x_5+x_1x_7+x_2x_4-x_2x_5+x_5x_6=0 \\ x_1x_2-2x_1x_5-x_1x_6-x_1x_7+x_2x_3+2x_2x_4-x_2x_5-x_2x_6+x_3^2-x_4x_6+x_5x_7=0 \\ x_1x_5-x_2x_3-x_2x_4-x_3^2-x_4x_7+x_6^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (0:1:-1:0:0:0:0) & -12 \\ (1:1:0:1:1:0:0) & -43 & (1:0:0:0:0:0:0) & -11 \\ (1:0:1:0:0:1:0) & -27 & (2:0:1:-1:0:1:0) & -4 \\ (0:0:0:0:1:0:0) & -19 & (2:-3:-1:0:-6:-4:0) & -3 \\ (0:0:1:-1:0:-1:0) & -16 & & \\ \bottomrule \end{array} \] \item Curve $X_0^+(233)$. Genus $g=7$. Equations of the canonical model in $\PP^{6}$ \[ \begin{sistema} -x_1x_2-x_1x_3-x_1x_4+x_2^2-x_2x_3-x_2x_5-x_2x_6+x_2x_7=0 \\ x_1x_2+x_1x_4-x_2^2+x_2x_6+x_3x_5=0 \\ x_1^2-2x_1x_2-x_1x_3+x_1x_5+2x_2^2-2x_2x_3-x_2x_4-2x_2x_5-x_2x_6+x_3x_6=0 \\ -2x_1x_3-x_1x_6+x_2^2-x_2x_3-x_2x_4-x_2x_5+x_3x_7=0 \\ x_1^2-2x_1x_2+x_1x_4+x_1x_5+2x_2^2-2x_2x_3-2x_2x_5-x_2x_6+x_3x_4+x_4^2=0 \\ x_1x_3+x_1x_6-x_2^2+x_2x_3+x_2x_5+x_4x_5=0 \\ x_1x_2+x_1x_3-x_1x_7-x_2^2+x_2x_3-x_2x_4+x_2x_5+x_2x_6+x_4x_6=0 \\ -x_1x_2-x_1x_3-x_1x_4+x_1x_5-x_1x_6-x_1x_7+2x_2^2-x_2x_3-x_2x_4-2x_2x_5-x_2x_6+ x_5^2=0 \\ -x_1^2+x_1x_2+2x_1x_3-x_1x_5+x_1x_6+x_1x_7-2x_2^2+2x_2x_3+x_2x_4+2x_2x_5+x_4x_7+ x_5x_6=0 \\ -x_1^2+x_1x_2-x_1x_5+x_2x_4-x_2x_6+x_5x_7+x_6^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (0:0:1:0:0:0:0) & -8 \\ (1:0:2:-2:1:0:2) & -28 & (1:0:0:0:-1:0:0) & -7 \\ (0:0:1:-1:0:0:0) & -19 & (1:1:0:-2:1:2:1) & -4 \\ (1:1:0:0:1:0:1) & -16 & & \\ \bottomrule \end{array} \] \item Curve $X_0^+(241)$. Genus $g=7$. Equations of the canonical model in $\PP^{6}$ \[ \begin{sistema} x_1^2+x_1x_4-x_2^2-x_2x_4+x_2x_6=0 \\ x_1x_5-x_2x_3-x_2x_5+x_2x_7=0 \\ x_1x_2+2x_1x_3+x_1x_6-x_1x_7-x_2^2+x_3^2=0 \\ x_1x_3-x_1x_7+x_2x_4+x_3x_4=0 \\ -x_1x_4-x_1x_5-x_2x_4+x_2x_5-x_3x_5-x_4^2+x_4x_5=0 \\ -x_1^2+x_1x_2-x_1x_3+x_1x_4+x_1x_5+x_1x_7+x_2x_3-2x_2x_4-x_2x_5+x_3x_5-x_3x_6+ x_4x_6=0 \\ -x_1^2-x_1x_3+x_1x_4-x_1x_6+x_2^2+x_2x_3-x_2x_4+x_3x_5-x_3x_6-x_3x_7+x_4x_7=0 \\ -x_1x_2+x_1x_3-3x_1x_4-x_1x_5+x_1x_6+x_2^2+x_2x_4-x_3x_5+x_3x_6-x_4^2+x_5^2=0 \\ -x_1^2-x_1x_3+x_1x_4+x_1x_5-x_1x_6+x_2^2+x_2x_3-2x_2x_4-x_2x_5+x_3x_5-x_3x_6- x_3x_7+x_5x_6=0 \\ -x_1^2+x_1x_2-x_1x_3+x_1x_4+2x_1x_5+x_1x_7+x_2x_3-2x_2x_4-2x_2x_5+x_5x_7+ x_6^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (1:1:0:0:1:0:0) & -12 \\ (1:0:-2:-1:0:0:0) & -67 & (0:0:0:1:1:0:0) & -8 \\ (0:1:-1:0:0:1:-1) & -27 & (1:-1:-2:0:0:0:-2) & -4 \\ (1:1:0:0:0:0:0) & -16 & (3:1:-4:-6:-3:4:2) & -3 \\ \bottomrule \end{array} \] \item Curve $X_0^+(257)$. Genus $g=7$. Equations of the canonical model in $\PP^{6}$ \[ \begin{sistema} -x_1^2-x_1x_2+x_1x_4-x_2^2-x_2x_4+x_2x_5=0 \\ -x_1^2-2x_1x_2+x_1x_3+x_1x_4-x_1x_5-2x_2^2-x_2x_4+x_2x_6=0 \\ -x_1x_2-x_1x_6-x_2x_3+x_2x_7=0 \\ x_1x_2-x_1x_3+x_1x_6+x_2^2+x_2x_4+x_3x_4=0 \\ -x_1x_3-x_1x_4+x_1x_6-x_1x_7+x_2^2-x_2x_3+x_2x_4+x_3x_5=0 \\ x_1^2+x_1x_2-x_1x_3-x_1x_4+2x_2^2-x_2x_3+2x_2x_4-x_3x_6+x_4x_5=0 \\ x_1^2+x_1x_2-2x_1x_4+x_1x_5-x_1x_7+2x_2^2-x_2x_3+x_2x_4+x_3^2-x_3x_6-x_3x_7+ x_4x_6=0 \\ x_1^2-x_1x_3-x_1x_4-x_1x_5-x_1x_6+x_2^2-x_2x_3+x_2x_4-x_3^2-x_3x_6+x_3x_7+ x_5^2=0 \\ x_1^2+x_1x_3-2x_1x_4-2x_1x_7+x_2^2-x_2x_3+x_2x_4+x_3^2-x_3x_6-x_3x_7+x_4x_7+ x_5x_6=0 \\ -2x_1x_2+2x_1x_3-2x_1x_6-x_1x_7-2x_2^2-x_2x_4-x_3x_6+x_4x_7-x_5x_7+x_6^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (0:0:1:0:0:0:1) & -11 \\ (1:0:0:1:0:0:-1) & -67 & (0:0:0:1:0:0:0) & -8 \\ (0:1:0:-1:0:1:0) & -16 & (2:-1:0:1:0:1:0) & -4 \\ \bottomrule \end{array} \] \item Curve $X_0^+(269)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} x_1x_4+x_2^2+x_2x_3+x_3^2+x_3x_5=0 \\ -x_1x_2-x_1x_5+x_2^2+x_2x_3+x_2x_5+x_3x_6=0 \\ x_1x_2+2x_1x_4+x_1x_6+x_3^2+x_3x_4+x_4^2=0 \\ x_1x_2-x_1x_4+x_1x_5-x_2^2-x_2x_3+x_2x_4+x_3x_4+x_4x_5=0 \\ x_1x_2+x_1x_3+x_1x_4+x_1x_5+x_1x_6-x_2x_3-x_2x_5+x_4x_6=0 \\ -x_1x_2-x_1x_3-2x_1x_4-x_1x_5-x_1x_6-x_2^2+2x_2x_5-x_2x_6-x_3^2+x_5^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:1) & \text{cusp} & (1:1:0:-1:0:0) & -16 \\ (1:-1:0:-1:1:2) & -67 & (1:0:0:0:0:0) & -11 \\ (1:1:0:-1:-1:0) & -43 & (1:-1:2:-3:0:0) & -4 \\ \bottomrule \end{array} \] \item Curve $X_0^+(271)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} x_1^2+2x_1x_2+x_1x_5-x_1x_6-x_2x_4-2x_2x_5-x_2x_6+x_3x_5=0 \\ x_1^2+x_1x_2-x_1x_4-x_1x_6-x_2x_3-x_2x_4-x_2x_5-x_2x_6+x_4^2=0 \\ -x_1^2-2x_1x_2-x_1x_5+x_1x_6+x_2x_3+x_2x_4+x_2x_5+x_4x_5=0 \\ x_1x_2-x_1x_4-x_1x_6-x_2x_3+x_2x_6+x_3^2+x_4x_6=0 \\ x_1^2+3x_1x_2+2x_1x_3+x_1x_5-3x_1x_6-2x_2x_4-x_2x_5+x_5^2=0 \\ -x_1x_2-x_1x_3-x_1x_4-x_1x_5+x_1x_6-x_2x_3+x_2x_4+x_2x_5+x_2x_6+x_3x_4+x_5x_6=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:1) & \text{cusp} & (0:1:0:0:0:0) & -19 \\ (1:0:1:1:0:1) & -43 & (1:0:1:0:0:1) & -12 \\ (0:1:1:1:-1:0) & -27 & (3:-2:-5:4:-4:-3) & -4 \\ \bottomrule \end{array} \] \item Curve $X_0^+(281)$. Genus $g=7$. Equations of the canonical model in $\PP^{6}$ \[ \begin{sistema} -x_1x_2+x_1x_3+x_2x_3+x_2x_4-x_3x_4+x_3x_6=0 \\ -x_1x_6-x_2x_3-x_2x_4-x_2x_5+x_3^2+x_3x_5+x_3x_7=0 \\ x_1x_2+x_1x_7+2x_3x_4-x_3x_5+x_4^2=0 \\ -x_1x_2-2x_1x_3-x_1x_4-x_1x_5-x_1x_7-x_2x_3+2x_3x_5+x_4x_5=0 \\ x_1x_2-x_1x_3+x_2x_5+2x_3x_4+x_4x_6=0 \\ 2x_1x_3+2x_1x_4+x_1x_5+x_1x_6+x_1x_7-x_2^2+x_2x_3+x_2x_4+x_2x_5+x_3x_4-2x_3x_5+ x_4x_7=0 \\ x_1x_2+2x_1x_3+x_1x_4+x_1x_5+x_1x_6+x_1x_7+x_2x_3-x_2x_4-x_3x_5+x_5^2=0 \\ x_1x_2-x_1x_3+x_2^2-x_2x_3-x_2x_4-x_2x_5+x_3x_5+x_5x_6=0 \\ x_1x_2-x_1x_3-x_1x_4-x_1x_7+x_2^2+x_2x_6+x_3x_5+x_5x_7=0 \\ x_1x_2-x_1x_3+x_1x_6-x_2x_3-x_2x_4-x_2x_7+2x_3x_4+x_6^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (1:0:0:0:0:0:0) & -16 \\ (2:-5:-5:-1:1:5:-3) & -163 & (1:0:0:1:0:0:-1) & -8 \\ (0:1:-1:1:1:1:-1) & -43 & (0:0:1:0:0:0:-1) & -7 \\ (0:0:1:-2:0:-2:-1) & -28 & (1:2:-2:2:2:0:-2) & -4 \\ \bottomrule \end{array} \] \item Curve $X_0^+(359)$. Genus $g=6$. Equations of the canonical model in $\PP^{5}$ \[ \begin{sistema} x_1x_2 -x_1x_5 + x_1x_6 + x_3x_4=0 \\ x_1^2 + x_1x_2 + x_1x_3 -2x_1x_4 - x_1x_5 + x_2x_4 + x_3x_5=0 \\ -3x_1x_2 -x_1x_3 + 2x_1x_4 + 2x_1x_5 - x_1x_6 + x_2^2 -x_2x_3 - x_2x_4-x_3^2 + x_3x_6=0 \\ x_1^2 + 3x_1x_2 + x_1x_3 -2x_1x_4 -2x_1x_5 + x_1x_6-x_2^2 +x_2x_4 + x_3^2 + x_4x_5=0 \\ -x_1^2 -2x_1x_2 -2x_1x_3 + 3x_1x_4 + 3x_1x_5 - x_6 + x_2^2 -2x_2x_4 - x_2x_5 -x_3^2 + x_4x_6=0 \\ x_1^2 +2x_1x_3 -2x_1x_4 - 2x_1x_5 + x_2^2 +x_2x_4 - x_2x_6 + x_5^2=0. \end{sistema} \] \[ \begin{array}{|cc|cc|} \toprule \text{Rational point} & \text{Disc.} & \text{Rational point} & \text{Disc.} \\ \midrule (0:0:0:0:0:0:1) & \text{cusp} & (1:0:0:0:0:0:0) & -28 \\ (2:-5:-5:-1:1:5:-3) & -163 & (1:0:0:1:0:0:-1) & -19 \\ (0:1:-1:1:1:1:-1) & -67 & (0:0:1:0:0:0:-1) & -7 \\ (0:0:1:-2:0:-2:-1) & -43 & & \\ \bottomrule \end{array} \] \end{itemize} \subsection{Modular parametrization}\label{subsec:modpar} Here we list the isogeny factor $E$ of $J_0^+(p)$, a generator $P_0$ of the Mordell Weil group, an explicit formula for the modular parametrization~$\phi_+\colon X_0^+(p)\to E$ and the modular degree $\deg(\phi_+)$ of the morphism $\phi_+$. \begin{comment} the bound $\delta$ used and the corresponding~$k_{\delta}$. For the definition of last two values see Proposition \ref{pro:upbnd} and Remark \ref{rem:upbnd} above. \end{comment} \begin{itemize} \item Curve $X_0^+(163)$. Elliptic curve $E: y^2+y=x^3-2x+1$. \\ \noindent Mordell-Weil generator $P_0=(1,0)$. Modular degree $\deg(\phi_+)=3$. \\ \noindent Modular parametrization: $\phi_+(x_1:x_2:x_3:x_4:x_5:x_6)=(P_x,P_y)$, \[ P_x =\frac{-2x_3-x_4+x_5-x_6}{-x_3+x_5}; \, P_y =\frac{-x_1^2+x_1x_2+2x_1x_3+ 2x_1x_4+2x_1x_6-x_3^2+x_4^2}{x_1x_3-x_1x_5}. \] \item Curve $X_0^+(197)$. Elliptic curve $E: y^2+y=x^3-5x+4$. \\ \noindent Mordell-Weil generator $P_0=(1,0)$. Modular degree $\deg(\phi_+)=5$. \\ \noindent Modular parametrization: $\phi_+(x_1:x_2:x_3:x_4:x_5:x_6)=(P_x,P_y)$, \begin{align*} \num(P_x)&=-38x_1^2-25x_1x_2-6x_1x_3+13x_1x_4-61x_1x_5+11x_1x_6+22x_2^2+ \\ &+10x_2x_5+20x_2x_6+17x_3x_5-11x_3x_6, \\ \den(P_x)&=-36x_1^2-34x_1x_2+11x_1x_3+30x_1x_4-27x_1x_5-8x_1x_6+12x_2^2+ \\ &+5x_2x_5+11x_2x_6, \\ \num(P_y) &=1715x_1^3-899x_1^2x_2-2691x_1^2x_3-1013x_1^2x_4-2889x_1^2x_5-81x_1^2x_6+ \\ &+1875x_1x_2^2+2476x_1x_2x_5+1176x_1x_2x_6-669x_1x_3x_5+1224x_1x_3x_6+ \\ &+2205x_1x_5x_6-1306x_1x_6^2-384x_2^2x_5-384x_2x_5x_6, \\ \den(P_y)&= 961x_1^3-49x_1^2x_2+3875x_1^2x3-632x_1^2x_4+1195x_1^2x_5+356x_1^2x_6+ \\ &+593x_1x_2^2+655x_1x_2x_5+432x_1x_2x_6-839x_1x_3^2-1680x_1x_3x_4+ \\ &+154x_1x_3x_5+384x_1x_5^2. \end{align*} \item Curve $X_0^+(229)$. Elliptic curve $E: y^2+xy=x^3-2x-1$. \\ \noindent Mordell-Weil generator $P_0=(-1,1)$. Modular degree $\deg(\phi_+)=4$. \\ \noindent Modular parametrization: $\phi_+(x_1:x_2:x_3:x_4:x_5:x_6:x_7)=(P_x,P_y)$, \begin{align*} P_x &=\frac{-x_2-x_5+x_6}{x_2+x_5}, \\ \num(P_y)&=x_1^2+3x_1x_2+7x_1x_3+6x_1x_4+x_1x_5-x_1x_6-4x_1x_7+ \\ &-2x_2x_3-2x_2x_4+3x_2x_6-2x_3^2+x_4x_6-2x_4x_7-x_6x_7, \\ \den(P_y)&=x_1^2+5x_1x_2+x_1x_4+2x_1x_5-x_1x_6- x_1x_7-x_3^2-x_4x_7+x_6^2. \end{align*} \item Curve $X_0^+(269)$. Elliptic curve $E: y^2+y=x^3-2x-1$. \\ \noindent Mordell-Weil generator $P_0=(-1,0)$. Modular degree $\deg(\phi_+)=3$. \\ \noindent Modular parametrization: $\phi_+(x_1:x_2:x_3:x_4:x_5:x_6)=(P_x,P_y)$, \[ P_x =\frac{x_4+x_6}{x_2}, \qquad P_y =\frac{x_1x_5-x_2x_3+x_3x_4}{x_1x_2}. \] \item Curve $X_0^+(359)$. Elliptic curve $E: y^2+xy+y=x^3-x^2-7x+8$. \\ \noindent Mordell-Weil generator $P_0=(2,-1)$. Modular degree $\deg(\phi_+)=4$. \\ \noindent Modular parametrization: $\phi_+(x_1:x_2:x_3:x_4:x_5:x_6)=(P_x,P_y)$, \begin{align*} P_x &=\frac{-4x_1^2 + 3x_1x_2 -4x_1x_3 -x_1x_5 + 3x_1x_6 + x_2x_4 + x_4^2}{-2x_1^2+ x_1x_2 -2x_1x_3 + x_1x_6}, \\ P_y &=\frac{-5x_1^2 + 4x_1x_2 -6x_1x_3 + 5x_1x_4 + 4x_1x_6 -x_2x_4 -x_2x_5 - x_2x_6 + x_3^2 + 2x_4^2}{3x_1^2 + 3x_1x_3 -2x_1x_4 - 2x_1x_5 + x_2x_4}. \end{align*} \end{itemize}
075551569946c0f8f0a7f47157fb0a019d5c4a18
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{introduction} The $\rho$-meson distribution amplitudes (DAs) are key components for collinear factorization of the $\rho$-meson involved processes, such as the semi-leptonic decay $B(D)\to \rho \ell \nu_{\ell}$ and the flavor-changing-neutral-current decays $B\to \rho\gamma$ and $B\to\rho \ell^+ \ell^-$, which are important for extracting the Cabibbo-Kobayashi-Maskawa (CKM) matrix elements and for searching new physics beyond the Standard Model. Inversely, those processes could provide good test of various $\rho$-meson DA models suggested in the literature. The $\rho$-meson DAs arouse people's great interest since the initial works of Refs.\cite{Lepage:1980fj, Brodsky:1981rp, Chernyak:1981zz} on the light-meson DAs. The vector $\rho$-meson DAs have more complex structures than the light pseudoscalar DAs. There are chiral-even and chiral-odd $\rho$-meson DAs due to chiral-even and chiral-odd operators in the matrix elements. The $\rho$-meson thus has two polarization states, either longitudinal ($\|$) or transverse ($\perp$), which can be expanded over different twist structures~\cite{Ball:1998sk, Ball:2004rg}. In the paper, we shall concentrate our attention on the $\rho$-meson longitudinal leading-twist DA $\phi_{2;\rho}^\|$. The leading-twist DA $\phi_{2;\rho}^\|(x,\mu)$ can be expanded as a Gegenbauer polynomial series~\cite{Lepage:1979zb}, i.e. \begin{equation} \phi_{2;\rho}^{\|}(x,\mu)= 6x(1-x)\left(1+\sum_{n} C^{3/2}_{n}(\xi) \times a^\|_{n;\rho}(\mu)\right), \label{phiDA} \end{equation} where $\xi=2x-1$. The Gegenbauer moments $a^\|_{n;\rho}$ at any other scale can be obtained via QCD evolution equation. The evolution equation up to next-to-leading order (NLO) is available in Ref.\cite{Ball:2006nr}. Theoretically, the Gegenbauer moments have been studied via various approaches~\cite{Ball:1996tb, Pimikov:2013usa, Ball:2007zt, Arthur:2010xf, Gao:2014bca, Forshaw:2010py, Forshaw:2012im, Choi:2007yu, Dorokhov:2006xw}. Most of their predictions are consistent with each other, but are of large theoretical uncertainties. It is helpful to provide an accurate prediction for a better comparison with the forthcoming more accurate experimental data. The conventional Shifman-Vainshtein-Zakharov (SVZ) sum rules~\cite{Shifman:1978bx} provides a standard way to deal with the hadron phenomenology. Within the framework of SVZ sum rules, hadrons are represented by interpolating quark currents with certain quantum numbers taken at large virtualities. The correlation function (correlator) of those currents is introduced and treated by using the operator product expansion (OPE), where short- and long-distance quark-gluon interactions are separated. The short-distance part is perturbatively calculable, while the non-perturbative long-distance part can be parameterized into the non-perturbative but universal vacuum condensates. The SVZ sum rules is then achieved by matching to a sum over hadronic states with the help of dispersion relation. The introduction of vacuum condensates is the basic assumption of the SVZ sum rules. Those universal vacuum condensates reflect non-perturbative nature of QCD which can be fixed via a global fit of experimental data. As suggested by the background-field theory (BFT), the quark and gluon fields are composed of background fields and quantum fluctuations around them. This way, the BFT provides a self-consistent description for the vacuum condensates and provides a systematic way to derive the SVZ sum rules~\cite{Novikov:1983gd, Hubschmid:1982pa, Govaerts:1983ka, Huang:1986wm, Huang:1989gv}. The SVZ sum rules within the BFT has been applied to deal with the pseudoscalar and scalar DAs~\cite{Wang:1984rw, Xiang:1984hw, Huang:2004tp, Zhong:2011jf, Zhong:2011rg, Han:2013zg}. In those calculations, because of the complexity of high-dimensional operators and also the contribution of high-dimensional condensates are generally power suppressed, one simply adopts the reduced quark propagators $S_F(x,0)$ and the vertex operators $\Gamma (z\cdot \tensor{D})^n$, which keep only up to dimension-three operators. Such a rough treatment is theoretically incomplete, which may miss some important high-dimensional condensates in the sum rules. Their contribution may be sizable, especially to compare with the NLO QCD corrections to the perturbative part. Thus to compare with the forth-coming more and more accurate data, it is helpful to take those high-dimensional terms into consideration. We have deduced the formulas for the quark propagator $S_F(x,0)$ and the vertex operators $\Gamma (z\cdot \tensor{D})^n$ within the BFT by keeping all terms in the OPE up to dimension-six operators~\cite{Zhong:2014jla, Zhong:2014fma}. For example, the quark propagator is parameterized as~\cite{Zhong:2014jla} \begin{eqnarray} S_F(x,0) &=& S_F^0(x,0)+S_F^2(x,0)+S_F^3(x,0)+ \sum_{i=1}^{2} S_F^{4(i)}(x,0) \nonumber\\ && + \sum_{i=1}^{3} S_F^{5(i)}(x,0) + \sum_{i=1}^{5} S_F^{6(i)}(x,0), \end{eqnarray} where $S_F^{k(i)}(x,0)$ stand for the propagator parts that are proportional to the dimension-$k$ operators with type $(i)$ under the same dimension. Those formulas help us to achieve a sound and accurate SVZ sum rules up to dimension-six condensates such as $\langle g_s \bar q q\rangle^2$ and $\langle g_s^3 f G^3\rangle$. Their first applications for the heavy and light pseudoscalar DAs have been done in Refs.\cite{Zhong:2014jla, Zhong:2014fma}. Those applications show that the new propagators and vertex operators shall result in new terms proportional to the dimension-six condensates that are missing in previous studies but do have sizable contributions. We shall adopt those newly derived quark propagator and vertex operators to study the $\rho$-meson longitudinal leading-twist DA $\phi_{2;\rho}^\|$. We shall then apply $\phi_{2;\rho}^\|$ to deal with the $B\to \rho$ transition form factors (TFFs) within the light-cone sum rules (LCSR)~\cite{Balitsky:1989ry, Chernyak:1990ag, Ball:1991bs}. As a further step, we shall show their effects to the $B$-meson semi-leptonic decay width $\Gamma(B\to \rho \ell \nu_\ell)$, which has been measured by the BABAR collaboration~\cite{delAmoSanchez:2010af, Aubert:2005cd}. The remaining parts of the paper are organized as follows. In Sec.\ref{section:2}, we describe the calculation technology for deriving the moments of the leading-twist DA $\phi_{2;\rho}^\|$ within the SVZ sum rules. In Sec.\ref{section:3}, we present the numerical results for the moments $\langle \xi_{n;\rho}^\| \rangle$, the decay width $\Gamma(B\to \rho\ell\nu_\ell)$, and the CKM matrix element $|V_{\rm ub}|$. Sec.\ref{section:4} is reserved for a summary. \section{Calculation technology}\label{section:2} Within the framework of BFT, the gluon field $\mathcal{A}^A_\mu(x)$ and quark field $\psi(x)$ in QCD Lagrangian are replaced by \begin{eqnarray} \mathcal{A}^A_\mu(x) &\to& \mathcal{A}^A_\mu(x) + \phi^A_\mu(x), \label{bfrep0} \\ \psi(x) &\to& \psi(x) + \eta(x), \end{eqnarray} where $\mathcal{A}^A_\mu(x)$ with $A =(1, \cdots, 8)$ and $\psi(x)$ at the right-hand-side are gluon and quark background fields, respectively. $\phi^A_\mu(x)$ and $\eta(x)$ stand for the gluon and quark quantum fields, i.e., the quantum fluctuation on the background fields. The QCD Lagrangian within the BFT can be found in Ref.\cite{Huang:1989gv}. The background fields satisfy the equations of motion \begin{equation} (i \slash \!\!\!\! D - m)\psi(x) = 0 \end{equation} and \begin{equation} \widetilde{D}^{AB}_\mu G^{B\nu\mu}(x) = g_s \bar{\psi}(x) \gamma^\nu T^A \psi(x), \end{equation} where $D_\mu = \partial_\mu - ig_s T^A \mathcal{A}^A_\mu(x)$ and $\widetilde{D}^{AB}_\mu = \delta^{AB} - g_s f^{ABC} \mathcal{A}^C_\mu(x)$ are fundamental and adjoint representations of the gauge covariant derivative, respectively. One can take different gauges for the quantum fluctuations and the background fields. A proper choice of gauge could make the sum rules calculation much more simplified. Practically, one usually adopts the background gauge, $\widetilde{D}^{AB}_\mu \phi^{B \mu}(x) = 0$, for the gluon quantum field~\cite{Novikov:1983gd, Hubschmid:1982pa, Govaerts:1983ka}, and the Schwinger gauge or the fixed-point gauge, $x^\mu \mathcal{A}^A_\mu(x) = 0$, for the background field~\cite{Shifman:1980ui}. Using those inputs, the quark propagator $S_F(x,0)$ and the vertex operators $\Gamma (z\cdot \tensor{D})^n$ are ready to be derived up to dimension-six operators within the BFT. We refer the interested readers to Ref.\cite{Zhong:2014jla} for details. Considering the definition \begin{eqnarray} && \langle 0|\bar d(0) z\!\!\!\slash (iz \cdot \tensor D )^n u(0) |\rho (q,\lambda ) \rangle \nonumber\\ = && (e^{(\lambda)*} \cdot z) (q \cdot z)^n m_\rho f_\rho^\parallel \langle \xi_{n;\rho}^\| \rangle, \label{matrix:2} \end{eqnarray} where $q$ and $e^{(\lambda)}$ are momentum and polarization vector of the $\rho$-meson, $(z\cdot \tensor{D})^n = (z\cdot \overrightarrow{D} - z\cdot \overleftarrow{D})^n$, and $f_\rho ^\parallel$ is the decay constant. The $n_{\rm th}$-order moment of $\rho$-meson leading-twist DA $\phi_{2;\rho}^\|(x,\mu)$ at the scale $\mu$ is defined as \begin{equation} \langle \xi_{n;\rho}^\|\rangle = \int_0^1 du (2x-1)^n \phi_{2;\rho}^\|(x,\mu). \label{DAmoments} \end{equation} As a special case, the $0_{\rm th}$-moment satisfies the normalization condition \begin{equation} \langle \xi_{0;\rho}^\| \rangle = \int_0^1 dx \phi_{2;\rho}^\|(x,\mu) = 1 . \label{normalizationDA} \end{equation} To derive the SVZ sum rules for the $\rho$-meson leading-twist DA moments $\langle \xi_{n;\rho}^\|\rangle$, we introduce the following correlator, \begin{eqnarray} \Pi^{(n,0)}_\rho (z,q) &=& i \int d^4x e^{iq\cdot x}\langle 0|T \{ J_n(x) J^\dag_0(0)\}| 0 \rangle \nonumber\\ &=& (z\cdot q)^{n+2} I^{(n,0)}(q^2), \label{correlator} \end{eqnarray} where $J_n(x) = \bar{d}(x) {z\!\!\!\slash} (i z\cdot \tensor{D})^n u(x)$ and $z^2 = 0$. Here $n=(0,2,\cdots)$, i.e. only even moments are nonzero due to the isospin symmetry. The correlator (\ref{correlator}) is an analytic $q^2$-function defined at both positive and negative $q^2$-values. In physical region ($q^2>0$), the complicated hadronic content of the correlator can be quantified by applying the unitarity relation through inserting a complete set of intermediate hadronic states to the matrix element. By singling out the ground-state and introducing a compact notation for the rest of contributions, we obtain \begin{eqnarray} \textrm{Im} I^{(n,0)}_{\rm had}(q^2) & =& \pi \delta (q^2 - m_\rho^2) {f_\rho^\|}^2 \langle \xi_{n;\rho}^\|\rangle \nonumber\\ &&+ \pi \frac{3}{4\pi^2 (n+1) (n+3)} \theta (q^2 - s_\rho), \label{hadim} \end{eqnarray} where the quark-hadron duality has been adopted and the parameter $s_\rho$ is the continuum threshold of the lowest continuum state. \begin{figure}[htb] \includegraphics[width=0.40\textwidth]{feyn.eps} \caption{Schematic Feynman diagrams for the $\rho$-meson longitudinal leading-twist DA moments, where the cross $(\times)$ stands for the background quark field. The big dots stand for the vertex operators in the correlator, the left one is for $z\!\!\!\slash (iz \cdot \tensor D )^n$ and the right one is for $z\!\!\!\slash$.} \label{feyn} \end{figure} On the other hand, one can apply the OPE for the correlator (\ref{correlator}) in deep Euclidean region $(q^2<0)$. The coefficients before the operators (result in non-perturbative condensates) are perturbatively calculable. The OPE indicates that \begin{eqnarray} && \Pi^{(n,0)}_{2;\rho} (z,q) = i \int d^4x e^{iq\cdot x} \nonumber\\ && \quad \times \big\{ - {\rm Tr} \langle 0 | S^d_F(0,x) z\!\!\!\slash (iz\cdot \tensor{D})^n S^u_F(x,0) z\!\!\!\slash | 0\rangle \nonumber\\ && \quad + {\rm Tr} \langle 0 | \bar{d}(x) d(0) z\!\!\!\slash (iz\cdot \tensor{D})^n S^u_F(x,0) z\!\!\!\slash | 0\rangle \nonumber\\ && \quad + {\rm Tr} \langle 0 | S^d_F(0,x) z\!\!\!\slash (iz\cdot \tensor{D})^n \bar{u}(0) u(x) z\!\!\!\slash | 0\rangle \big\} + \cdots ~. \label{ope} \end{eqnarray} The first term corresponds to Fig.(\ref{feyn}a), the second one corresponds to Fig.(\ref{feyn}b), and the third one corresponds to permutation contribution by transforming $u\leftrightarrow d$ from Fig.(\ref{feyn}b). We adopt the dimensional regularization under the $\overline{\rm MS}$-scheme to deal with the infrared divergences at high orders, whose divergent terms shall be absorbed into the renormalized leading-twist DA~\cite{Li:2012gr}. As a combination of the correlator within different $q^2$-region, the sum rules for the moments of the $\rho$-meson leading-twist DA can be derived by using the dispersion relation \begin{eqnarray} \frac{1}{\pi} \frac{1}{M^2} \int ds e^{-s/M^2} \textrm{Im} I_{\rm had}(s) = \hat{L}_M I_{\rm QCD}(q^2), \label{bordisrel} \end{eqnarray} where $M$ is the Borel parameter and the Borel transformation operator \begin{eqnarray} \hat{L}_M = \lim_{\begin{array}{c} Q^2,n\to\infty\\ Q^2/n=M^2 \end{array}} \frac{1}{(n-1)!} (Q^2)^n \left( -\frac{d}{dQ^2} \right)^n, \end{eqnarray} where $Q^2=-q^2$. The final sum rules reads \begin{widetext} \begin{eqnarray}\label{xiSR} &&\langle \xi_{n;\rho}^\| \rangle = \frac{M^2}{f_\rho^2}e^{m_\rho^2/M^2} \Bigg\{ \frac{3}{4\pi^2(n+1)(n+3)}\left(1 + \frac{{{\alpha _s}}}{\pi }A_n'\right) \left( 1 - e^{-s_\rho/M^2} \right) + \sum\limits_{q=u,d} \left( \frac{m_q \langle \bar q q\rangle}{M^4} - \frac{8n+1}{18} \frac{m_q \langle g_s \bar q\sigma TGq\rangle}{M^6}\right. \nonumber \\ &&\quad + \left.\frac{4n+2}{81} \frac{\langle g_s\bar q q \rangle^2}{M^6}\right)+ \frac{1 + n\theta(n-2)}{12\pi(n+1)} \frac{\langle \alpha_s G^2\rangle }{M^4} + \frac{1}{16\pi}\frac{\langle g_s^3fG^3\rangle}{M^6} \left\{\frac{8\delta^{n0} + 405n + 192}{36}\ln\frac{M^2}{\mu^2} - \frac{16\delta^{n0} + 810n + 363}{72} \right. \nonumber \\ && \quad \times \gamma_E+ \frac{7}{24} \psi(n+1) + \frac{8 \delta^{n0} + 405n + 826}{72} + \theta (n - 2)\left[ \frac{16 - 22n}{72}\ln \frac{M^2}{\mu^2} - \frac{788n + 421}{72}\psi(n + 1) - \frac{766n + 437}{72} \gamma_E \right. \nonumber \\ &&\quad - \frac{68n^2 - 37n - 11}{144n} + \left.\left. \sum\limits_{k=0}^{n-2} {(-1)^k} \frac{1}{144}\left( \frac{3(135k + 128)}{n-k} + \frac{383k+399}{k-n+1} - \frac{106kn-410k+617n-415}{(k + 1)(k + 2)} + 106 \right) \right]\right\}\Bigg\}, \nonumber\\ \label{momsr} \end{eqnarray} \end{widetext} where the step function $\theta(x)=1$ for $x\geq0$, and $\theta(x)=0$ for $x<0$. $\gamma_E=0.557216$ is Euler's constant and the $0_{\rm th}$-derivative of the digamma function $\psi(n+1)=\sum_{k=1}^n 1/k -\gamma_E$. The NLO coefficients $A_n$ have been calculated by Ref.\cite{Ball:1996tb}, whose first three ones are, $A'_0 = 1$, $A'_2 = \frac{5}{3}$ and $A'_4 = \frac{59}{27}$, respectively. One can obtain relations among the Gegenbauer moments $a_{n;\rho}^\|$ and the moments $\langle \xi_{n;\rho}^\|\rangle$ by substituting Eq.(\ref{phiDA}) into Eq.(\ref{DAmoments}). For examples, we have \begin{eqnarray} a_{2;\rho}^\| &=& \frac{7}{12} \left( 5\langle \xi^\|_{2;\rho} \rangle - 1 \right), \label{gegmom1} \\ a_{4;\rho}^\| &=& -\frac{11}{24} \left( 14\langle \xi^\|_{2;\rho} \rangle - 21\langle \xi_{4;\rho}^\| \rangle - 1 \right), \label{gegmom2} \\ a_{6;\rho}^\| &=& \frac{5}{64} \left( 135\langle \xi^\|_{2;\rho} \rangle - 495\langle \xi^\|_{4;\rho} \rangle + 429\langle \xi^\|_{6;\rho} \rangle - 5 \right). \label{gegmom3} \end{eqnarray} \section{Numerical results and discussions} \label{section:3} We adopt the following parameters to do the numerical calculation. The $\rho$-meson mass and decay constant are from the Particle Data Group~\cite{Agashe:2014kda}, $m_\rho = 0.775 {\rm GeV}$ and $f_\rho^\| = 0.216 \pm 0.003 {\rm GeV}$. The non-perturbative vacuum condensates up to dimension-six have been determined in references~\cite{Nambu:1960xd, GellMann:1968rz, Narison:2010wb, Belyaev:1982sa, Narison:2009vy, Bordes:1988yr, Dominguez:1987nw, Causse:1989cr, Launer:1983ib, Bertlmann:1984ih, Narison:1995jr, Narison:2001ix, Chung:1981wm}, \begin{eqnarray} \langle\bar q q\rangle &=& -0.0138(17){\rm GeV}^3, \nonumber\\ \langle g_s \bar q q\rangle^2 &=& -0.0018(7){\rm GeV}^6, \nonumber\\ \langle\alpha_s G^2\rangle &=& 0.038(11){\rm GeV}^4, \nonumber\\ \langle g_s^3 f G^3\rangle &=& 0.013(7){\rm GeV}^6, \nonumber\\ \sum_{q=u,d} m_q \langle \bar{q}q\rangle &=& -1.656(5)\times 10^{-4} {\rm GeV}^4, \nonumber\\ ~\sum_{q=u,d}m_q \langle g_s\bar{q}\sigma TGq\rangle &=& 1.325(33) \times 10^{-4} \textrm{GeV}^4. \nonumber \end{eqnarray} The continuum threshold $s_\rho$ is usually set as the value around the squared mass of the $\rho$-meson first excited state. At present, the structure of the excited $\rho$-meson state is not yet completely clear, cf. a recent review in Ref.\cite{Agashe:2014kda}. Therefore, we use the sum rules (\ref{momsr}) with $n=0$, together with the normalization condition $\langle \xi_{0;\rho}^\|\rangle=1$, to inversely determine an effective value for $s_\rho$. We get, $s_\rho\simeq 2.8{\rm GeV}^2$, which indicates that the effective threshold continuum state is close to $\rho(1700)$. \subsection{The $\rho$-meson leading-twist DA $\phi_{2;\rho}^\|(x,\mu)$} To determine a Borel window for the sum rules (\ref{momsr}), e.g. the allowable range for $M$, we adopt two criteria: I) All continuum contributions are less than $40\%$ of the total dispersion relation; II) The contributions from the dimension-six condensates should not exceed $10\%$. By setting all other parameters to be their central values, the first two moments $\langle\xi_{2;\rho}^\|\rangle$ and $\langle\xi_{4;\rho}^\|\rangle$ up to NLO level at the scale $\mu=M$ are determined as \begin{equation} \langle\xi_{2;\rho}^\|\rangle|_{\mu=M}=0.234(23) \;\;{\rm for}\;\; M^2\in [1.72,3.00] \end{equation} and \begin{equation} \langle\xi_{4;\rho}^\|\rangle|_{\mu=M}=0.103(7) \;\;{\rm for}\;\; M^2\in [4.26,4.86], \end{equation} where the central values are for $M^2=2.185$ and $4.535$, respectively. \begin{widetext} \begin{center} \begin{table}[htb] \caption{The first two moments $\langle\xi_{(2,4);\rho}^\|\rangle|_{\mu=M}$ of the longitudinal leading-twist DA $\phi_{2;\rho}^\|$ predicted from the sum rules under the BFT. Here the perturbative contributions are calculated up to NLO level and the non-perturbative contributions are up to dimension-six condensates. The contributions from the LO-terms, the NLO-terms, the dimension-three, the dimension-four, the dimension-five and the dimension-six condensates are presented separately. The errors are obtain by varying $M^2$ within the determined Borel window. } \begin{tabular}{c| c c c c c cc } \hline & LO & NLO & Dimension-three & Dimension-four & Dimension-five & Dimension-six & Total \\ \hline $\langle \xi_{2;\rho}^\| \rangle |_{\mu=M} $ & 0.193(34)& $0.014(1)$ & $-0.0021(7)$ & 0.013(6) & 0.0007(5) & 0.015(22) & 0.234(23) \\ $\langle \xi_{4;\rho}^\| \rangle |_{\mu=M} $ & 0.081(4) & $0.008(1)$ & $-0.0008(1)$ & 0.005(2) & 0.0003(1)& 0.010(1) & 0.103(7) \\ \hline \end{tabular}\label{NLO6dim} \end{table} \end{center} \end{widetext} To show how the non-perturbative dimension-six condensates and the perturbative NLO corrections affect the moments, we list the first two moments $\langle\xi_{(2,4);\rho}^\|\rangle$ at the scale $M$ in Table \ref{NLO6dim}, where the perturbative contributions are calculated up to NLO level and the non-perturbative contributions are up to dimension-six condensates. The contributions from the LO-terms, the NLO-terms, the dimension-three, the dimension-four, the dimension-five and the dimension-six condensates are presented separately in Table \ref{NLO6dim}. It shows that the dominant contribution is from the LO-terms, which provide $\sim80\%$ contribution to $\langle\xi_{(2,4);\rho}^\|\rangle$. The NLO-terms provide $\sim 6.0\%$ contribution to $\langle\xi_{2;\rho}^\|\rangle$ and $\sim 7.8\%$ contribution to $\langle\xi_{4;\rho}^\|\rangle$. It is noted that the non-perturbative condensates do not follow the usual power counting of $1/M^2$-suppression, and the dimension-six condensates provide sizable contributions to the moments which are at the same order of the NLO-terms. Thus they are of equal importance for a precise prediction of the $\phi_{2;\rho}^\|$ moments. By using the relations among the Gegenbauer moments $a_{n;\rho}^\|$ and the moments $\langle \xi_{n;\rho}^\|\rangle$, such as Eqs.(\ref{gegmom1},\ref{gegmom2},\ref{gegmom3}), we can derive $a_{n;\rho}^\|$ at the scale $M$. Furthermore, the Gegenbauer moments $a_{n;\rho}^\|$ at any other scale can be obtained via the QCD evolution, i.e. the evolution at the NLO accuracy shows~\cite{Floratos:1977au, Mueller:1994cn, Ball:2006nr} \begin{eqnarray} a^{\|}_{n;\rho}(\mu) &=& a_{n;\rho}^\|(\mu_0) E_{n;\rho}^{\rm NLO} \nonumber\\ &+& \frac{\alpha_s(\mu)}{4\pi}\sum_{k=0}^{n-2} a_{k;\rho}(\mu_0)\, L^{\gamma_k^{(0)}/(2\beta_0)}d^{(1)}_{nk}, \end{eqnarray} where $\mu_0$ is the initial scale, $\mu$ is the required scale, and \begin{eqnarray} E_{n;\rho}^{\rm NLO} &=& L^{\gamma^{(0)}_n/(2\beta_0)} \nonumber\\ && \times\left\{1+ \frac{\gamma^{(1)}_n \beta_0-\gamma_n^{(0)}\beta_1}{8\pi\beta_0^2} \Big[\alpha_s(\mu)-\alpha_s(\mu_0)\Big]\right\}, \end{eqnarray} where $L=\alpha_s(\mu)/\alpha_s(\mu_0)$, $\beta_0=11-2n_f/3$, $\beta_1=102-38n_f/3$, $\gamma_n^{(0)}$ and $\gamma_n^{(1)}$ are LO and NLO anomalous dimensions, accordingly. \begin{widetext} \begin{center} \begin{table}[htb] \caption{The first two Gegenbauer moments $a_{2;\rho}^\|$ and $a_{4;\rho}^\|$ for the longitudinal leading-twist DA $\phi_{2;\rho}^\|$, which is predicted from the sum rules under the BFT. A comparison of predictions under various approaches~\cite{Pimikov:2013usa, Ball:2007zt, Arthur:2010xf, Gao:2014bca, Forshaw:2010py, Forshaw:2012im, Choi:2007yu, Dorokhov:2006xw} has also been presented. For easy comparison, we have set the scale $\mu_{0}=1{\rm GeV}$. The moments $\langle\xi_{2;\rho}^\|\rangle$ and $\langle\xi_{4;\rho}^\|\rangle$ and the inverse moment $\langle x^{-1}\rangle$ are also presented. The number in the parenthesis shows the uncertainties from all the input parameters. } \begin{tabular}{c| c c c c c} \hline ~ ~~~~~~&~~~~~~ $a_{2;\rho}^\|$ ~~~~~~&~~~~~~ $a_{4;\rho}^\|$ ~~~~~~&~~~~~~ $\langle \xi_{2;\rho}^\| \rangle$ ~~~~~~&~~~~~~ $\langle \xi_{4;\rho}^\| \rangle$ ~~~~~~& $\langle x^{-1} \rangle$ \\ \hline our predictions & $0.119(82)$ & $-0.035(100)$ & $0.241(28)$ & $0.109(10)$ & 3.30(34) \\ NLCSR~\cite{Pimikov:2013usa} & 0.047(58) & $-0.057(118)$ & 0.216(21) & 0.089(9) & 2.97(39) \\ BB~\cite{Ball:2007zt} & 0.150(70) & - & 0.251(24) & - & 3.45(21) \\ Lattice QCD~\cite{Arthur:2010xf} & 0.197(158) & - & 0.268(54) & - & 3.60(48) \\ BS~\cite{Gao:2014bca} & 0.111 & 0.036 & 0.238 & 0.115 & 3.44 \\ AdS/QCD~\cite{Forshaw:2010py, Forshaw:2012im} & 0.104 & 0.053 & 0.236 & 0.115 & 3.47 \\ LFQM~\cite{Choi:2007yu} & 0.014 & $-0.005$ & 0.205 & 0.088 & 3.03 \\ IM~\cite{Dorokhov:2006xw} & $-0.010$ & $-0.033$ & $0.196$ & $0.080$ & 2.87 \\ \hline \end{tabular}\label{taba2} \end{table} \end{center} \end{widetext} We present our predictions for the Gegenbauer moments $a_{2(4);\rho}^\|$, together with the moments $\xi_{2(4);\rho}^\|$ and the inverse moment $\langle x^{-1} \rangle = \int_0^1 dx x^{-1} \phi_{2;\rho}^\| (x,\mu)$, in Table~\ref{taba2}, where all uncertainty sources have been taken into consideration and have been summed up in quadrature. Because of the dominance of the LO-terms to the moments $\langle\xi_{(2,4);\rho}^\|\rangle$, the $\phi_{2;\rho}^\|$ behavior and the quantities such as the TFFs and $|V_{\rm ub}|$ shall be dominated by the LO-terms. For example, as will be shown later, if without taking the NLO-terms and the dimension-six condensates into consideration, the magnitudes of $A_1$, $A_2$ and $V_{0}$ at $q^2=0$ shall be altered by $3\%-4\%$; and the magnitude of $|V_{ub}|$ shall be altered by $\sim5\%$. \begin{figure}[htb] \centering \includegraphics[width=0.45\textwidth]{DA.eps} \caption{The $\rho$-meson leading-twist DA $\phi_{2;\rho}^\|(x,\mu_{0} =1{\rm GeV})$ predicted from the sum rules under the BFT (BFTSR). As a comparison, the NLCSR prediction~\cite{Pimikov:2013usa}, the BB prediction~\cite{Ball:2007zt}, the Lattice QCD prediction~\cite{Arthur:2010xf}, the AdS/QCD prediction~\cite{Forshaw:2010py, Forshaw:2012im}, the LFQM prediction~\cite{Choi:2007yu}, and the IM prediction~\cite{Dorokhov:2006xw} have also been presented. } \label{DA:comparation} \end{figure} As a comparison, we also present the sum rules prediction with nonlocal condensates (NLCSR)~\cite{Pimikov:2013usa}, the Ball and Brawn (BB) prediction~\cite{Ball:2007zt}, the Lattice QCD prediction~\cite{Arthur:2010xf}, the Bethe-Salpeter wavefunction (BS) prediction~\cite{Gao:2014bca}, the AdS/QCD prediction~\cite{Forshaw:2010py, Forshaw:2012im}, the Light-Front Quark Model (LFQM) prediction~\cite{Choi:2007yu}, and the Instanton Model (IM) prediction~\cite{Dorokhov:2006xw} in Table~\ref{taba2}. To compare with the other predictions, we have set the scale $\mu_{0}=1{\rm GeV}$, which is adopted in most of the references. It is noted that our present predictions on the DA moments agree with most of them within reasonable errors, and most of them prefer a double humped behavior, as explicitly shown by Fig.(\ref{DA:comparation}). \begin{figure}[htb] \includegraphics[width=0.45\textwidth]{A6.eps} \caption{The $\rho$-meson longitudinal twist-2 DA $\phi_{2;\rho}^\|(x,\mu_0=1{\rm GeV})$ for $n =2$, $4$ and $6$, respectively.} \label{fig2} \end{figure} We make a discussion on how the $\phi_{2;\rho}^\|$ behavior changes with different truncations of the Gegenbauer expansion. By taking the central values for the Gegenbauer moment $a_{n;\rho}^\|$, we put the DA $\phi_{2;\rho}^\|(x,\mu_0=1{\rm GeV})$ for $n=(2,4,6)$ in Fig.(\ref{fig2}). It shows that by including the sixth-moment $a_{6;\rho}^\| (1{\rm GeV})= 0.009 $ into the Gegenbauer expansion, the shape of $\phi_{2;\rho}^\|$ is slightly changed and close to the double humped behavior for the case of $n=4$. By including more moments into the expansion, the $\phi_{2;\rho}^\|$ behavior shall be almost unchanged. Thus it is convenient and is of high precision to keep only the first two moments in the Gegenbauer expansion. The $\rho$-meson leading-twist wavefunction $\psi_{2;\rho }^\| (x,\mathbf{k}_\perp)$ is an important component for a reliable pQCD predictions within the $k_T$ factorization formalism~\cite{Botts:1989kf, Li:1992nu}. We adopt the present DA moments to fix a $\rho$-meson wavefunction $\psi_{2;\rho}^\|(x,{\bf k}_\bot)$ that is constructed from the Wu-Huang (WH) model~\cite{Wu:2010zc} \begin{eqnarray} \psi_{2;\rho}^\|(x,{{\bf{k}}_ \bot }) = \sum\limits_{{h _1}{h _2}} {{\chi_\rho^{{h _1}{h_2}}}} (x,{{\bf{k}}_ \bot }) \psi _{2;\rho}^{R}(x,{{\bf{k}}_\bot }), \end{eqnarray} whose radial part is from the BHL-prescription~\cite{BHL}. The spin-space wavefunction ${{\chi_\rho^{{h _1}{h_2}}}} (x,{{\bf{k}}_ \bot })$ is from the Wigner-Melosh rotation~\cite{Huang:1994dy, Cao:1997hw, Huang:2004su}. The $\rho$-meson DA $\phi^\|_{2;\rho}$ can be derived from $\psi_{2;\rho}^\|(x,{{\bf{k}}_ \bot })$ via the relation \begin{eqnarray} \phi_{2;\rho}^\|(x,\mu) = \frac{ 2\sqrt{3}}{ \widetilde{f}_\rho^\|}\int_{|{\bf k}_\bot|^2\leq\mu^2}\frac{d{\bf k}_\bot}{16\pi^3}\psi_{2;\rho}^\|(x,{\bf k}_\bot)\,, \end{eqnarray} which leads to \begin{eqnarray} &&\phi _{2;\rho }^\| (x,\mu) = \frac{{A_{2;\rho}^\| \sqrt {3x\bar x} {m_q}}}{{8{\pi ^{\frac{3}{2}}}\widetilde f_\rho ^\| b_{2;\rho}^\| }}[1 + {B_{2;\rho}^\| }C_2^{\frac{3}{2}}(\xi )+ C_{2;\rho}^\| C_4^{\frac{3}{2}}(\xi )]\nonumber\\ &&\qquad\times \left[ {{\rm{Erf}}\left( {b_{2;\rho}^\| \sqrt {\frac{{{\mu^2} + m_q^2}}{{x\bar x}}} } \right) - {\rm{Erf}}\left( {b_{2;\rho}^\| \sqrt {\frac{{m_q^2}}{{x\bar x}}} } \right)} \right], \nonumber\\ \label{DA:WH} \end{eqnarray} where $\widetilde{f}_\rho^\| = f_\rho^{\|}/\sqrt{5}$, the error function, $\textrm{Erf}(x) = 2 \int^x_0 e^{-t^2} dt/ \sqrt \pi$ and the light constitute quark mass, $m_q\simeq300$ GeV. To be slightly different from the one suggested in Ref.\cite{Fu:2014cna}, we have explicitly put the newly derived fourth Gegenbauer term in the longitudinal function. Four model parameters can be fixed by the normalization condition, the average value of the squared transverse momentum $\langle {\bf k}_\bot^2 \rangle_{2;\rho}^{1/2} = 0.37 \pm 0.02 {\rm GeV}$, and the second and fourth Gegenbauer moments determined from the sum rules (\ref{momsr}). By using the central values for the input parameters, we obtain: $A_{2;\rho}^\|=24.61$, $b_{2;\rho}^\|=0.581$, $B_{2;\rho}^\| = 0.075$ and $C_{2;\rho}^\| = -0.044$. \subsection{The $B\to \rho$ transition form factors} \begin{figure}[htb] \centering \includegraphics[width=0.45\textwidth]{A1.eps} \includegraphics[width=0.45\textwidth]{A2.eps} \includegraphics[width=0.45\textwidth]{V.eps} \caption{The extrapolated $B\to \rho$ axial-vector and vector TFFs $A_{1,2}(q^2)$ and $V(q^2)$ by using the LCSRs derived in Ref.\cite{Fu:2014cna}, where the $\rho$-meson leading-twist DA $\phi_{2;\rho}^\|$ is from our present sum rules under the BFT (BFTSR). The lattice QCD predictions~\cite{Flynn:1995dc, Bowler:2004zb} are presented as a comparison. } \label{fig_TFF} \end{figure} \begin{table}[htb] \begin{tabular}{ c c c c } \hline & ~~~~$A_1$~~~~ & ~~~~$A_2$~~~~ & ~~~~$V$~~~~ \\ \hline $a_1^i$ & 0.233 & $-0.874$ & $-1.034$ \\ $a_2^i$ & 0.345 & $0.708$ & $5.257$ \\ $\Delta$ & 0.16 & 0.23 & 0.41 \\ \hline \end{tabular} \caption{The fitted parameters $a^i_{1,2}$ for the $B\to \rho$ TFFs $F_i$, in which all the LCSR parameters are set to be their central values. $\Delta$ is the measure of the quality of extrapolation. } \label{analytic} \end{table} One of the important application of $\phi_{2;\rho}^\|$ is the $B$-meson semi-leptonic decay $B\to\rho\ell\nu_\ell$. It is the key component for the vector and axial vector $B\to\rho$ TFFs $A_{1}(q^2)$, $A_{2}(q^2)$ and $V(q^2)$. By using a left-handed current $j_B^\dag (x) = i\bar b(x)(1-\gamma_5)q_2(x)$ to do the LCSR calculation, one can highlight the contributions from $\phi_{2;\rho}^\|$~\cite{Fu:2014cna}, thus showing the properties of $\phi_{2;\rho}^\|$ via a more transparent way. Following the standard LCSR procedures, one can derive the LCSRs for the mentioned TFFs, which have been presented in Ref.\cite{Fu:2014cna}. One only needs to replace the DA $\phi_{2;\rho}^\|$ used there to be our present one. At the large recoil region, $q^2 \approx 0 {\rm GeV}^2$, we obtain \begin{eqnarray} A_1(0) &=& 0.237^{+0.029}_{-0.021}~, \label{F01} \\ A_2(0) &=& 0.246^{+0.063}_{-0.043}~, \label{F02} \\ V(0) &=& 0.268^{+0.021}_{-0.017}~, \label{F03} \end{eqnarray} where the errors are squared averages of all error sources for the LCSRs. If using the $\phi_{2;\rho}^\|$ determined from the sum rules without the NLO-terms and the dimension-six condensates, we obtain $A_1(0)=0.230^{+0.028}_{-0.020}$, $A_2(0)=0.257^{+0.063}_{-0.043}$ and $V(0)=0.262^{+0.020}_{-0.016}$. Those values change from the above ones, i.e. Eqs.(\ref{F01},\ref{F02},\ref{F03}), determined from the sum rules with the NLO-terms and the dimension-six condensates by about $3\%-4\%$. We put those TFFs versus $q^2$ in Fig.(\ref{fig_TFF}), where we have extrapolated them to all allowable $q^2$-region via the rapidly converging series in the parameter $z(t)$ expansion which is suggested by Refs.\cite{Khodjamirian:2010vf, Bourrely:2008za, Straub:2015ica} \begin{eqnarray} F_i(q^2) = \frac1{1-q^2/m_{R,i}^2}\sum_{k=0,1,2}a_k^i [z(q^2)-z(0)]^k, \end{eqnarray} where \begin{eqnarray} z(t)=\frac{\sqrt{t_+ - t}-\sqrt{t_+ - t_0}}{\sqrt{t_+ - t}+\sqrt{t_+ - t_0}} \end{eqnarray} with $t_\pm=(m_B\pm m_\rho)^2$ and $t_0=t_+(1-\sqrt{1-t_-/t_+})$. The values of the resonance masses $m_{R,i}$ can be found in Ref.~\cite{Straub:2015ica}, and $F_i$ stands for the three TFFs. The parameters $a_k^i$ are fixed such that $\Delta < 1\%$, which are put in Table~\ref{analytic}. The measure of the quality of extrapolation $\Delta$ is defined as \begin{equation} \Delta=\frac{\sum_t\left|F_i(t)-F_i^{\rm fit}(t)\right|} {\sum_t\left|F_i(t)\right|}\times 100, \label{delta} \end{equation} where $t\in[0,\frac{1}{2},\cdots,\frac{27}{2},14]{\rm GeV}^2$. \begin{table}[htb] \begin{center} \begin{tabular}{c c | c } \hline \multicolumn{2}{c|}{Our prediction} & ~~~$3.19^{+0.65}_{-0.62}$~~~ \\ \hline \multicolumn{2}{c|}{Omn\`{e}s parametrization~\cite{Flynn:2008zr}} & $2.80(20)$ \\ \hline & LCSR~\cite{Ball:2004rg} & $2.75(24)$ \\ \raisebox {1.5ex}[0pt]{BABAR~\cite{delAmoSanchez:2010af}} & ISGW~\cite{Scora:1995ty} & $2.83(24)$ \\ \hline & LCSR~\cite{Ball:2004rg} & $2.85(40)$ \\ \raisebox {1.5ex}[0pt]{BABAR~\cite{Aubert:2005cd}} & ISGW~\cite{Scora:1995ty} & $2.91(40)$ \\ \hline \end{tabular} \caption{The predicted $|V_{\rm ub}|$ in unit $10^{-3}$. The estimations of the Omn\`{e}s parametrization~\cite{Flynn:2008zr} and BABAR collaboration~\cite{delAmoSanchez:2010af,Aubert:2005cd} are also presented as a comparison.}\label{Gammatota} \end{center} \end{table} We apply the extrapolated $B\to\rho$ TFFs for the semi-leptonic decays, $B^0\to\rho^-\ell^+\nu_\ell$ and $B^+\to\rho^0\ell^+\nu_\ell$. Their branching ratios and lifetimes are~\cite{Agashe:2014kda}: ${\cal B}(B^0\to\rho^-\ell^+\nu_\ell)=(2.94\pm0.21)\times 10^{-4}$ and $\tau(B^0)=1.520\pm0.004~{\rm ps}$; ${\cal B}(B^+\to\rho^0\ell^+\nu_\ell)=(1.58\pm0.11)\times 10^{-4}$ and $\tau(B^+)=1.638\pm0.004~{\rm ps}$. Those two semi-leptonic decays can be adopted to determine the CKM matrix element $|V_{\rm ub}|$, we present the results in Table~\ref{Gammatota}. Both of them lead to the same predictions, $|V_{\rm ub}| = 3.19^{+0.65}_{-0.62}$, where the errors are squared averages of the errors from $\xi_{(2,4);\rho}^\|$, the Borel window, the continuum threshold $s^{B}_0$ for the $B\to\rho$ TFFs, the $b$-quark mass, the $B$-meson decay constant and the uncertainties from the measured lifetimes and branching ratios, respectively. If using the $\phi_{2;\rho}^\|$ determined from the sum rules without the NLO-terms and the dimension-six condensates, we obtain $|V_{\rm ub}| = 3.36^{+0.66}_{-0.64}$, which changes from the one determined from the sum rules with the NLO-terms and the dimension-six condensates by about $5\%$. Table \ref{Gammatota} shows our result is consistent with the Omn\`{e}s parametrization and BABAR prediction within errors. \section{Summary} \label{section:4} The BFT provides a clean physical picture for the perturbative and non-perturbative properties of QCD and provides a systematic way to derive the SVZ sum rules for hadron phenomenology. In the paper, we have studied the moments of the $\rho$-meson leading-twist DA $\phi_{2;\rho}^\|$ via the SVZ sum rules up to dimension-six operators and by taking the NLO QCD correction to the perturbative part. Our predictions for the second and fourth moments $\langle \xi_{2;\rho}^\| \rangle$ and $\langle \xi_{4;\rho}^\| \rangle$, which lead to the Gegenbauer moments $a_{2;\rho}^{\|}|_{1{\rm GeV}} = 0.119(82)$ and $a_{4;\rho}^\| |_{1 {\rm GeV}} = -0.035(100)$. They indicate a double humped behavior for $\phi_{2;\rho}^\|$, which agrees with most of predictions done in the literature. \begin{figure}[tb] \centering \includegraphics[width=0.45\textwidth]{dG.eps} \caption{The predicted differential decay width $1/|V_{\rm ub}|^{2} \times d\Gamma/dq^2$, where the shaded band shows the squared average of all the errors from the mentioned error sources. The lattice QCD predictions~\cite{Flynn:1995dc, Bowler:2004zb} are presented as a comparison. } \label{Fig:dGamma} \end{figure} The $\rho$-meson DA is a key component for $\rho$-meson involved high-energy processes. A better determination of the $\rho$-meson DA shall be helpful for a better understanding of the $\rho$-meson physics. As an application of $\phi_{2;\rho}^\|$, we calculate the $B\to\rho\ell\nu_\ell$ semi-leptonic decays within the LCSR via a chiral correlator. It is found that the extrapolated $B\to\rho$ TFFs agrees with the lattice QCD predictions~\cite{Flynn:1995dc, Bowler:2004zb} within errors. This can be more clearly shown by Fig.(\ref{Fig:dGamma}), which shows the different decay width $1/|V_{\rm ub}|^2\times d\Gamma/dq^2$. Our present obtained $\rho$-meson DA $\phi_{2;\rho}^\|$ shall be further constrained/tested by more data available in the near future, and we hope the definite behavior can be concluded finally. \\ {\bf Acknowledgments}: This work was supported in part by the Natural Science Foundation of China under Grant No.11275280, No.11547305, No.11547015, and by Fundamental Research Funds for the Central Universities under Grant No.CDJZR305513.
8d5f5f5841cc13aeb8a9dc888a3bfca2c8a5bedf
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:intro} Hyperspectral imaging systems have become increasingly popular for a variety of applications, including remote sensing and biomedical analytics. With their dense, contiguous and narrow-band spectral sampling in visible through short-wave infrared regions of the electromagnetic spectrum, they provide rich spectral characterization of the objects that are dominant in the pixels of the hyperspectral image. The capability of hyperspectral data to accurately capture material-specific properties makes them an attractive choice for characterizing vegetation mapping invasive and endangered vegetation, (e.g., for ecological and precision agriculture applications), detection and characterization of physiological conditions and other related biomedical applications \cite{vyas2013estimating,li2012compressive,chen2011denoising,NG2011,prasad2014SMoG,PLFB2012,KPB2010Derivatives,prasad2009information,YP2015,PC2013Asilomar_Sparse,li2011multi,di2011active,PB2008a}. In addition to such applications where hyperspectral data has seen a widespread popularity and acceptance, we suggest that hyperspectral data is also particularly suited for enhanced computer vision applications as they relate to scene understanding, biometrics, and person re-dentification, by virtue of their robust characterization of material properties. Person re-identification -- the task of recognizing a person separated in location and time, has emerged as an important application of multi-camera surveillance systems. Despite algorithmic advances \cite{gheissari2006person,farenzena2010person,zheng2011person} built upon traditional computer vision imaging systems, person reidentification remains a difficult problem due to various challenges, including variations in illumination conditions, pose and viewpoints. Advances in this area include pose, viewpoint and illumination invariant feature extraction \cite{gheissari2006person,farenzena2010person}, as well as the design of appropriate similarity metrics. A majority of these methods take into account global appearance of a person, such as a weighted color histogram \cite{farenzena2010person}. Local spatial information is often extracted by analyzing cropped regions around the face \cite{dantcheva2011frontal}. In this paper, we assert that hyperspectral imagery is potentially very beneficial for the task of person re-identification in a multi-camera surveillance scenario --- specifically, spectral content can serve as a powerful descriptor that can be used by itself or in conjunction with classical spatial and statistical features derived for re-identification. The spectral reflectance \emph{signature} will demonstrate variability across materials (e.g. clothing) and across skin between different people. While it is expected that traditional computer vision imaging systems can characterize differences in clothing and other traditional descriptors, which may be highly separable between people in the visible wavelength regime, subtle differences due to individual skin physiology can be better characterized by hyperspectral imagery. With this in mind, we developed a pilot study and acquired hyperspectral imagery from fifteen subjects in the visible and very near infrared region of the electromagnetic spectrum at two different locations and at different times of the day (morning versus afternoon). We utilized a simple distance metric appropriate for hyperspectral data and reported re-identification performance using spectral features derived from superpixel patches over the skin, to demonstrate the efficacy of such data for person reidentification. This was compared with reidentification undertaken with color imagery. The outline of this paper is as follows. In section 2, we provide details of the data acquisition, and describe the data that was acquired. In section 3, we describe the approach utilized to setup and quantify re-identification performance with this data. In section 4, we provide experimental results for the re-identification task, and in section 5, we provide concluding remarks. \section{Hyperspectral Data Acquisition for Person Reidenitifcation} \label{sec:data} With the goal of demonstrating the efficacy of hyperspectral data for person re-identification, we acquired hyperspectral data from $15$ people under a traditional multi-camera surveillance scenario at two different locations and two different times of the day (morning and afternoon). The hyperspectral images were acquired using a Headwall Photonics$^\text{TM}$ hyperspectral imager --- the spatial size of each image was $1004\times400$ and each pixel had $325$ contiguous spectral bands spanning the visible and near-infrared spectrum from $400nm-1000nm$ uniformly, with a full-width half-maximum (FWHM) bandwidth of $1.8nm$. In order to evaluate the performance of a re-identification model, the dataset must represent commonly encountered confounding factors such as viewpoint and illumination variation. Hence, for each person, the images were acquired at two different times of the day (morning and afternoon), and for each of these two acquisitions, the viewpoints, and location of the camera were different. This data collection was carried out in two batches --- in the series of 15 hyperspectral images acquired in the morning, the camera orientation was fixed relative to the background in the scene, and each person was asked to arbitrarily stand somewhere in the scene and assume a natural (and arbitrary pose). Again, in the afternoon, another series of 15 images were acquired with the sample 15 people, but at a different location, and with a different orientation of the camera. For the purpose of quantifying the potential of hyperspectral images for person re-identification, we treat the morning images from the 15 people as the gallery set, and the afternoon images from those people as the probe set. Also, in addition to the camera having two different orientations and locations in the morning and afternoon, each person was asked to stand arbitrarily in the scene in order to get arbitrary viewpoints. Due to the temporal difference of the captured data, variations in physiological conditions are also represented in this dataset. Fig. \ref{fig:spectra} depicts spectral signatures from the background objects and various parts of the head from one of the $15$ people. It also depicts spectral signatures corresponding to skin pixels (from the face and hands) for two people (Person 12 and Person 13) in the gallery (AM) and probe (PM) sets. This suggests that hyperspectral data can effectively characterize variability between people when using pixels over skin objects. \begin{figure}[h]% \centering \includegraphics[width=8cm, natwidth =560, natheight = 420]{MeanSpecAllFace3AM.png} \includegraphics[width=8cm, natwidth =560, natheight = 420]{MeanSpecProbGallerySkin.png} \caption{\emph{Top}: Mean spectral signatures for various material types in one of the images, including background, clothes, skin (from face and hands), and spectra from other parts of the head; \emph{Bottom}: Mean spectra of ``skin'' from people with IDs 12 and 13 in the gallery set (AM) and the probe set (PM).} \label{fig:spectra} \end{figure} \section{Spectral Angle based Person Re-identification} \label{sec:method} In this paper, we focus on the utility of spectral information from skin pixels for person re-identification. Our approach consists of the following steps --- we manually annotated small patches of ``skin'' pixels (on the face and hands) for each person in the gallery and probe sets. We carried out superpixel segmentation to match mean spectra from skin superpixels to avoid pixel mixing that would result from rectangular windows that are traditionally employed. By averaging spectral signatures over superpixels, we stabilize spectral response from local variability and noise, while at the same time avoiding inadvertent mixing that would have resulted had we chosen rectangular windows. Specifically, we utilized the entropy rate superpixel algorithm which ensures compact and homogenous clusters of similar sizes \cite{ERSuperPixelLiu2011,priyasuperpixels2015}. We took superpixels that intersected with our ``ground-truth'' skin patches in each image and used those to match each probe image with the gallery. A spectral angle distance was used as the similarity metric for the re-identification task. If $S_i$ and $S_j$ represent superpixels in the gallery and probe set respectively, then a spectral angle distance $d_{sa} (S_i,S_j)$ between the two superpixels can be defined as : \begin{equation}\label{eq:SAD} {d_{sa} (S_i,S_j)} = {\frac{{{\hat{{\mathbf{x}}}}_{S_i}^T}{{\hat{{\mathbf{x}}}}_{S_j}}}{{\parallel {{\hat{{\mathbf{x}}}}_{S_i}} \parallel} {\parallel {{\hat{{\mathbf{x}}}}_{S_j}} \parallel}}} \end{equation} \noindent where, ${\hat{{\mathbf{x}}}}_{S_i}$ is the mean spectral signature of region ${S_i}$. In addition to facilitating a computationally simple comparison, using spectral angle as a similarity metric with spectral signature features provides us with an added advantage in that the metric exhibits illumination invariance \cite{CP2015_ADA_JSTSP,keshava2004distance,adler2001shadow}. Average spectral angle distances between skin superpixels from the probe image and every image in the gallery set are used to re-identify the probe set. Note that the superpixels that result from the entropy rate superpixel algorithm modified to use spectral angle based similarity are highly effective at oversegmenting the image into uniform sized clusters while preserving borders, and are hence appropriate for spectral matching via the spectral angle distance. In a practical implementation, matching skin superpixels between probe and gallery sets can be easily accomplished by first utilizing a face detection algorithm to identify the skin superpixels in all the images. For this closed set re-identification problem, we report performance via the cumulative matching characteristic (CMC) curve. The CMC curve represents the expectation of finding the correct match in the top \textit{{r}} matches. In other words, a rank-\textit{{r}} recognition rate shows the percentage of the test images that are correctly recognized from the top \textit{{r}} matches in the gallery set. The rank-1 value on this curve indicates the true identification performance, while the rank-$N$ score ($N$ being the number of images in the gallery) will be $100\%$ for closed-set re-identification, with the curve monotonically increasing from 1 through $N$. Comparison between hyperspectral and RGB (color) images based on this approach is reported in the CMC curve in Fig.~\ref{fig:CMC}. We observe that hyperspectral images not only provides a superior rank-1 performance compared to when using (RGB) color images, the CMC curve for hyperspectral images increases much faster as a function of the rank, with hyperspectral data providing as much as $20\%$ better identification performance. \begin{figure} \centering \includegraphics[height = 7cm]{CMC_Sk.png} \caption{CMC curves}\label{fig:CMC} \end{figure} \section{Conclusions, Caveats and Future Work} We conclude from the data and the results that hyperspectral images have potential to enhance re-identification performance in multi-camera surveillance systems --- the preliminary dataset and resulting CMC curves suggest the ``value'' added by considering the narrow-band spectral information for re-identification as opposed to 3-channel color images. By being able to distinguish subtle spectral variations between people (e.g. via the spectral signatures of their skin), it may also enable long-duration re-identification wherein significant time may have elapsed between the person reappearing in the collective field of view of the system. We acknowledge an important caveat with results presented here --- the sample size (30 images total in gallery and probe sets) is small --- overall performance will drop when the number of people are added to the re-identification problem. However, we expect that spectral signature as a descriptor to characterize skin pixels effectively is likely to result in superior performance compared to traditional (RGB) color systems. Finally, we note that the approach to re-identification can be enhanced by complementing spectral signatures with other descriptors (e.g. spatial information). In ongoing work, we are expanding the gallery and probe sets, and are exploring strategies to fuse information provided from spectral signatures with currently established color image based feature descriptors for effective re-identification. \balance \bibliographystyle{IEEEbib}
a40ff918a828c71cff8294c6a3e4fb1c61f74189
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section[#1 \debug{\fbox {#2}}]{#1 \cmd{msec} \dlabel{#2}}% \markboth{\today}{Sec. \thesection}} \newcommand{\msubsection}[2]{\subsection[#1 \debug{\fbox {#2}}] {#1 \cmd{msubsection} \dlabel{#2}}% \markboth{\today}{Sec. \thesection}} \newcommand{\msubsubsection}[2]{\subsubsection[#1 \debug{\fbox {#2}}] {#1 \cmd{msubsubsection} \dlabel{#2}}% \markboth{\today}{Sec. \thesection}} \renewcommand{{\hfill $\Box$}{\cmd{PE}}}{{\hfill qed}{\cmd{PE}}} \renewcommand{\dB}{\begin{description}} \renewcommand{\labelitemi}{$\bullet$} \setlist{leftmargin=5.5mm} \newtheorem{theorem}{Theorem} \newtheorem{lemma}{Lemma} \newtheorem{definition}{Definition} \newcommand{\textbf}{\textbf} \newcommand{\cdot}{\cdot} \begin{document} \title{ Stochastic Broadcast Control of Multi-Agent Swarms} \author{Ilana Segall and Alfred Bruckstein } \date{Center for Intelligent Systems\\MultiAgent Robotic Systems (MARS) Lab\\Computer Science Department\\ Technion, Haifa 32000, Israel} \maketitle \newpage \tableofcontents \newpage \begin{abstract} We present a model for controlling swarms of mobile agents via broadcast control, assumed to be detected by a random set of agents in the swarm. The agents that detect the control signal become ad-hoc leaders of the swarm. The agents are assumed to be velocity controlled, identical, anonymous, memoryless units with limited capabilities of sensing their neighborhood. Each agent is programmed to behave according to a linear local gathering process, based on the relative position of all its neighbors. The detected exogenous control, which is a desired velocity vector, is added by the leaders to the local gathering control. The graph induced by the agents adjacency is referred to as the \emph{visibility graph}. We show that for piecewise constant system parameters and a connected visibility graph, the swarm asymptotically aligns in each time-interval on a line in the direction of the exogenous control signal, and all the agents move with identical speed. These results hold for two models of pairwise influence in the gathering process, uniform and scaled. The impact of the influence model is mostly evident when the visibility graph is incomplete. These results are conditioned by the preservation of the connectedness of the visibility graph. In the second part of the report we analyze sufficient conditions for preserving the connectedness of the visibility graph. We show that if the visibility graph is complete then certain bounds on the control signal suffice to preserve the completeness of the graph. However, when the graph is incomplete, general conditions, independent of the leaders topology, could not be found.\end{abstract} \textbf{Keywords}: broadcast control, leaders following, linear agreement protocol, collective behavior, conditions for maintaining connectivity, neighbors influence, piecewise constant linear systems \section{Introduction}\label{Introduction} We present a system composed of a group or swarm of autonomous agents and a controller. All the agents behave according to a distributed gathering process, ensuring cohesion of the swarm, and the controller sends desired velocity controls to the cloud. The signal sent by the controller is received by a random set of agents. If all the agents receive the signal then the cloud will move with the desired velocity. If only part of the agents receive the signal then the cloud will move in the desired direction but with a fraction of the desired speed, depending on the topology of the inter-agent visibility graph. This can be viewed as representing the "inertia" or the "reluctance of the cloud to move" in the desired direction with the desired speed. We investigate two models of neighbors influence in the local control, uniform and scaled. We show that if the visibility graph is complete, then the ratio of the achieved collective speed to the desired speed, for both influence models, is the ratio of the \emph{number} of leaders to the total number of agents. However, when the graph is incomplete, the ratio of the achieved collective speed to the desired speed is a function of the influence model. If the influence is uniform the ratio stays as before, i.e. the ratio of the number of leaders to the total number of agents. but if the influence is scaled then the ratio of the achieved collective speed to the desired speed depends not only on the number of leaders but also on the exact topology of the visibility graph and on the location of the leaders within the graph. Hence, for the same number of leaders in the same incomplete visibility graph, with scaled influence, different results can be obtained for different leaders. \subsection{Statement of problem}\label{Model} We consider a system composed of $n$ homogeneous agents evolving in $\textbf{R}^2$. The agents are assumed to be homogenous, memoryless, with limited visibility (myopic) and are modeled by single integrators, namely are velocity controlled. The visibility (sensing) zone of agent $i$ is a disc of radius $R$ around its location. Agents within the sensing zone of agent $i$ are referred to as the \emph{neighbors} of $i$. If $j$ is a neighbor of $i$, we write $i \sim j$. The set of neighbors of $i$, define the \emph{neighborhood} of $i$, denoted by $N_i$. The emergent behavior of agents with unlimited visibility and stochastic broadcast control was discussed in \dcite{SB2016} Each agent can measure only the \emph{relative position} of other agents in its own local coordinate system. The orientation of all local coordinate systems is aligned to that of a global coordinate system, as illustrated in Fig.\dref{fig-frames}, i.e. agents are assumed to have compasses enabling them to align their local reference frames to a global reference frame. Here $p_i=(x_i,y_i)$ represents the position of agent $i$ in the global reference frame, \emph{unknown to the agent itself}. \begin{figure} \begin{center} \includegraphics[scale=0.4]{ReferenceFrames.jpg} \caption{Illustration of local and global reference frames alignment}\label{fig-frames} \end{center} \end{figure} We assume that the agents do not have data transmission capabilities, but all the agents are capable of detecting an exogenous, broadcast control. At any time, a random set of agents detect the broadcast control. These agents will be referred to as ad-hoc \emph{leaders}, while the remaining agents will be the \emph{followers}. The exogenous control, a velocity vector $u$, is common to all the leaders. The agents are unaware of which of their neighbors are leaders. The setup of the problem is illustrated in Fig. \dref{topology}. \begin{figure} \begin{center} \includegraphics[scale=0.5]{ProblemIllustration2.jpg} \caption{Illustration of problem topology at a certain point in time}\label{topology} \end{center} \end{figure} The sets of the leaders and of the followers are denoted by $N^l$, $N^f$ respectively. The number of leaders and followers in the system is denoted by $n_l=|N^l|$, $n_f=|N^f|$ respectively. The sum $n=n_f+n_l$ is the total number of agents. The agents are labeled $1,...,n$. \msubsection{The dynamics of the agents}{DynModel} In our model, the followers apply a local gathering control based on the relative position of all their neighbors and the leaders apply \emph{the same local control (\dref{gen-SelfDyn}) with the addition of the exogenous input $u$}. In general, the strength of the influence of neighbor $j$ on the movement of agent $i$ is some function $f(j,i)$, most often a function of the distance between $i$ and $j$, cf. \dcite{M-T}, \dcite{J-E}, \dcite{C-S}. If we denote by $\sigma_{ji}$ the strength of the \emph{influence} of agent $j$ on the movement of agent $i$, then we have: \begin{itemize} \item for each $i \in N^f$ \begin{equation}\dlabel{gen-SelfDyn} \dot{p}_i(t)=\sum\limits_{j \sim i} \sigma_{ji}(t) (p_j(t) - p_i(t)) \end{equation} where $p_i$ is the position of agent $i$ \item for $ i \in N^l$ \begin{equation}\dlabel{gen-LeadDyn} \dot{p}_i(t)=\sum\limits_{j \sim i} \sigma _{ji}(t) (p_j(t) - p_i(t)) + u \end{equation} \end{itemize} We consider two cases of influence : \begin{enumerate} \item \emph{Uniform} - The influence of all neighbors on any agent is identical and time independent, i.e. $\sigma_{ji} (t) =1 \text{ } \forall j \in N_i (t)$. \item \emph{Scaled} - The influence of an agent $j \in N_i(t)$ on $i$ is scaled by the size of the neighborhood $N_i(t)$, i.e. for each $i$, we have $ \sigma_{ji}(t) = \frac{1}{|N_i (t)|} ; \forall j \in N_i (t)$. \end{enumerate} Fig. \dref{UniScaled} illustrates an example of pairwise interaction graph with uniform influences, denoted by $G^U$, vs the corresponding graph with scaled influences, denoted by $G^S$. \begin{figure} \begin{center} \includegraphics[scale=0.4]{UniScaledGraph5n6m.jpg} \caption{Illustration of an interactions graphs with uniform vs scaled influence }\label{UniScaled} \end{center} \end{figure} In this report, we derive the emergent behavior of agents with any visibility graph, complete or incomplete, applying protocol (\dref{gen-SelfDyn}), (\dref{gen-LeadDyn}) for followers and leaders, for both influence models. We show that when the visibility graph is complete (due to a very large $R$) the two influence models will move the swarm with the same velocity to the same asymptotic (moving) gathering point but when the graph is incomplete the two influence models affect differently the collective velocity and asymptotic state of the swarm. Since $p_i(t) = [x_i(t) \quad y_i(t)]^T$ and $u =[u_x \quad u_y]^T$ and assuming that $x_i(t)$ and $y_i(t)$ are decoupled we can write \begin{eqnarray} \dot{x}_i(t)& = & \sum\limits_{j \in N_i}\sigma_{ji} (x_j(t) - x_i(t)) + b_i u_x\dlabel{x-1D}\\ \dot{y}_i(t)& = & \sum\limits_{j \in N_i}\sigma_{ji} (y_j(t) - y_i(t)) + b_i u_y\dlabel{y-1D} \end{eqnarray} and consider $x_i(t)$ and $y_i(t)$ separately, as one dimensional dynamics, (cf. Section \dref{Gen-1D}). \begin{equation}\label{bi} b_i = \begin{cases}1 ; \text{ if } i \in N^l \\ 0 ; \text{ otherwise} \end{cases} \end{equation} where $N^l$ is the set of leaders. In the piecewise constant case, when the time-line can be divided into intervals in which the system evolves as a linear time-independent dynamic system, (\dref{x-1D}) can be written in vector form as \begin{equation}\dlabel{x-1D-vec} \dot{x}( t )=-L_k\cdot x(t)+ B_k u_x(t_k) \end{equation} and similarly for (\dref{y-1D}), where \begin{itemize} \item $t \in [t_k \quad t_{k+1} )$ \item $t_k$ is a switching point, i.e. the time when either the visibility graph , the leaders or the exogenous control change \item $L_k$ is the Laplacian associated with the interactions graph $G_k$, either uniform or scaled, in the interval $[t_k \quad t_{k+1} )$ \item $B_k$ is a leaders indicator in the interval $[t_k \quad t_{k+1} )$ , i.e. a vector of dimension $n$ with $0$ entries in places corresponding to the followers and $1$ in those corresponding to the leaders \item $u_x(t_k)$ is the $x$ - component of the exogenous control $u$ in the interval $[t_k \quad t_{k+1} )$ \item $L_k, B_k, u_k $ are constant \end{itemize} The emergent behavior in the interval $t \in [t_k, t_{k+1})$ is a function of the corresponding properties of $L_k$. We show in the sequel that if the influence is uniform then the corresponding Laplacian is symmetric and its properties are independent of the topology of the graph but if the influence is scaled then the Laplacian corresponding to an incomplete graph is non-symmetric while the Laplacian corresponding to a complete graph is symmetric with the corresponding change in properties. In the sequel we treat each such interval separately, and thus it is convenient to suppress the subscript $k$. We first assume $G_k$ to be strongly connected for all $k$. In the second part of the report, we show scenarios and conditions for never losing friends, i.e. for $G_k \subseteq G_{k+1}; \quad \forall k$. We show that if $G_k$ is complete then bounding $|u_k|$ suffices to ensure that it remains complete. However, if $G_k$ is incomplete, the conditions are tightly related to the graph topology and could be derived only for specific cases. \textbf{\textit{Note}} that losing visibility to a neighbor does not necessarily mean losing connectivity. However, never losing neighbors ensures never losing connectivity. \msubsection{Literature survey and contribution}{survey} Many ways of controlling the collective behavior of self-organized multi-agent systems by means of one or more \textbf{special agents}, referred to as \emph{leaders} or \emph{shills}, have been investigated in recent years. We will be grouping the surveyed work in several broad categories and indicate the novelty of our model as compared to each. \begin{enumerate} \item \underline{Leaders that do not abide by the agreement protocol}\\ These leaders are pre-designated and their state value is fixed at a desired value. Jadbabaie et al. in \dcite{JLM} consider Vicsek's discrete model \dcite{VC95}, and introduce a leader that moves with a fixed heading . Tanner, Rahmani, Mesbahi and others in \dcite{Tanner}, \dcite{Rahmani}, \dcite{Rahmani2}, \dcite{RM2}, etc. consider static leaders (sometimes named "anchors") and show conditions on the topology that will ensure the controllability of the group. A system is controllable if for any initial state there exists a control input that transfers any initial state to any final state in finite time. Our model differs from the above in that the leaders are neither pre-designated nor static. The number of leaders and their identity is arbitrary. They do not ignore the agreement protocol, but rather add the received exogenous control to the computed local rule of motion and move accordingly. We do not require the system to reach a pre-defined final state. Our aim is to steer the swarm in a desired direction. We show the emergent dynamics for a desired velocity sent by a controller and received by random agents in the swarm. \item \underline{Leaders combining the consensus protocol with goal attraction}\\In \dcite{DGEH}, \dcite{GDEH09}, the exogenous control is a goal position, known only to the leaders. The dynamics of all agents, leaders or followers, is based on the consensus protocol. For leaders however, it includes an additional goal attraction term which aims at leading the team to the \textbf{pre-defined goal position}. The attraction term is a function of the leader's distance from the goal position, therefore varies from leader to leader. This approach is the closest to our model that we have found in the surveyed literature, but some major differences exist. In our model the exogenous control is not a goal position but a velocity vector, $u$, \textbf{common to all leaders}. Moreover, agents are not aware of their own position, but only of their relative position to their neighbors. We show that with our model, the agents, rather than gathering at a goal position, \emph{asymptotically align along a line in the direction of $u$ and move with identical speed}. \item \underline{Shills} - Intelligent agents with on-line state information of regular units. Han, Guo and Li, \dcite{HLG2006}, followed by Wang, \dcite{HW2013} introduced the notions of shill and soft control. Shills are special agents \textbf{added} to the swarm with the purpose of controlling the collective behavior. They are the exogenously controlled part of the system. The basic local rules of motion of the existing agents in the system are not changed. The existing agents treat the special agent as an ordinary agent, thus enabling it to "cheat" or "seduce" its neighbors towards the desired goals. These special agents are called "shills" \footnote{Shill is a decoy who acts as an enthusiastic customer in order to stimulate the participation of others}. As opposed to the above, in our work we study the emergent collective behavior when probabilistically selected agents, out of the existing agents, receive an exogenous control $u$. These agents become the ad-hoc leaders. The number of leaders is not predetermined, hence can be any number from $1$ to $n$. Also, we do not design $u$ in order to obtain some desired final state. Moreover, in our model the leaders do not have an entirely stand-alone control rule. All agents follow the same rule of motion, with the addition of the exogenous control, when received, i.e. while leaders. Leaders do not have on-line state information of other agents. The only available information, for leaders and all other agents, is relative position to neighbors. \item \underline{Broadcast control}\\ Recently Azuma, Yoshimura and Sugie \dcite{AYS2013} have proposed a broadcast control framework for multi-agent coordination, but in their model the control is assumed to be received by all units, i.e. there are \emph{no followers}. In this model the global controller observes the group performance, designs the information to be broadcast and sends a signal, received by all, to govern the group behavior. The agents set the local control, based on the received signal. As opposed to the above, in our model the broadcast control is the goal velocity vector, aiming to steer the swarm in some desired direction with desired speed. The detailed group performance is not directly observed by the controller, therefore the broadcast control does not depend on it. Moreover, not all units necessarily receive the broadcast control, but at least one does. \end{enumerate} \msubsection{Paper outline}{Outline} We derive the collective swarm behavior for piecewise constant systemד. We first treat each time interval separately, as a \emph{time-independent system } over an interval $[0,t)$. Section \dref{Gen-1D} presents the one dimensional case which is readily extended to two dimensions in Section \dref{LTI-dyn}. In Section \dref{SimEx} we show simulation results, illustrating the two dimensional swarm behavior over a single time interval. In Section \dref{MultiIntEx} we extend the investigation of one interval to multiple intervals, where new intervals are triggered by changes in the exogenous control, $u$, in leaders or in the visibility graph. We assume that $u$ and the leaders change randomly, but the visibility graph is state dependent, therefore, when the visibility is limited, the system may disconnect. In Section \dref{never-losing2} we derive conditions for a complete visibility graph to remain complete and in Section \dref{LimitedComplete-Ex} we illustrate the effect of the derived bounds. In Section \dref{LimitedIncomplete} we derive conditions for never losing friends, when the visibility graph is incomplete, and show that these depend on the exact, time-dependent, topology. We conclude in section \dref{future} with a short summary and directions for future research. \msec{One dimensional group dynamics}{Gen-1D} In this Section we consider a one dimensional piecewise constant system, (\dref{x-1D-vec}). In the sequel we threat each time interval, $[t_k \quad t_{k+1} )$, separately. Thus, it is convenient to suppress the subscript $k$. Moreover, it is convenient to denote by $t$ the relative time since the beginning of the interval ($t=0$) and by $x(0)$ the state of the system at this time. We then have (in each interval) \begin{equation}\dlabel{1D-vec} \dot{x}( t )=-L\cdot x(t)+ B u \end{equation} Eq. (\dref{1D-vec}) has the well known solution (ref. \dcite{TK}) \begin{equation}\dlabel{eq-pieceDyn} x(t) = e^{-L t} x(0) + \int_{0}^{t} e^{-L(t-\tau)} B u \mathrm{d}\tau \end{equation} Eq. (\dref{eq-pieceDyn}) can be rewritten as \begin{equation}\dlabel{eq-xdyn2} x(t) = x^{(h)}(t) + x^{(u)}(t) \end{equation} where \begin{itemize} \item $x^{(h)}(t)= e^{-L t} x(0)$ represents the zero input solution \item $x^{(u)}(t)=\int_{0}^{t} e^{-L (t-\tau)} B u \mathrm{d}\tau$ represents the contribution of the exogenous input to the group dynamics \end{itemize} \msubsection{Definitions}{Def} \begin{itemize} \item $G^U$ an undirected graph of uniform interactions, with vertices labeled $1,...,n$. \item ${{d}_{i}}$ the number of neighbors of vertex $i \in G^U$, i.e. the degree of $i$ \item $\Delta $ the degree matrix of the graph $G^U$, a diagonal matrix with elements $\Delta_{ii}={{d}_{i}}$, \item $A^U$ the adjacency matrix of $G^U$, a symmetric matrix with 0,1 elements, such that \begin{equation*} A^U_{ij} = \begin{cases} 1 & \text{if } i \sim{\ } j \\ 0 & \text{otherwise } \\ \end{cases} \end{equation*} \item $L^U = L(G^U)$ the Laplacian representing $G^U$, is defined by \begin{equation}\dlabel{def-Lu} L^U=\Delta -A^U \end{equation}\dlabel{Def-Gamma} \item $\Gamma$ the normalized Laplacian of $G^U$, is defined by \begin{equation}\dlabel{def-Gamma} \Gamma= \Delta^{-1/2} L^U \Delta^{-1/2} \end{equation} \item $G^S$ the directed graph of scaled interactions corresponding to $G^U$ \item $L^S=L(G^S)$ the Laplacian representing $G^S$ \begin{equation}\dlabel{def-Ls} L^S=\Delta^{-1} L^U \end{equation} \end{itemize} \emph{\underline{Note}} that Eq. (\dref{1D-vec}) and its general solution (\dref{eq-pieceDyn}) hold for both $L^U$ or $L^S$. In the following Sections, we develop explicit solutions for each case and investigate their properties. \msubsection{Zero input group dynamics }{ZeroInp} Denote by $L$ the Laplacian associated with the time-independent visibility graph, in the time interval. The zero input group dynamics is given by \begin{equation}\dlabel{eq-agree} \dot{x}^{(h)}( t )=-L\cdot x^{(h)}(t) \end{equation} We will show that for both $L=L^U$ and $L=L^S$, representing Laplacians of connected graphs and strongly connected digraphs respectively, the solution of eq. (\dref{eq-agree}) converges asymptotically to a consensus state, namely $x^{(h)}_i=x^{(h)}_j=\alpha; \forall i,j, i\neq j$ (cf. Proposition 2 in \dcite{OS-M2003}).\\ Since we consider each interval separately and $t$ is the time elapsed from the beginning of the interval, by "asymptotic state" we mean here the value of the state for large $t$. The value of the consensus state $\alpha$, in each interval, is obtained by explicitly calculating $exp(-Lt)$, for large $t$, as described below. \subsubsection{Uniform influence - Symmetric Laplacian $L^U$} \LB{L-uniConverge} The value of the consensus state for an undirected, connected, interactions graph with corresponding Laplacian, $L^U$, is the average of the initial states. \LE {\em Proof:\/\ }\cmd{ PB} \\ Using the properties of $L^U$ (cf. Appendix \dref{Graphs}), namely that: \begin{itemize} \item $L^U$ is a real symmetric positive semi-definite matrix \item all the eigenvalues of $L^U$, denoted by $\lambda^U_i$ are real and non-negative. \item if $G^U$ is connected then there is a single zero eigenvalue, denoted by $\lambda^U_1$ and the remaining eigenvalues are strictly positive. \item we can always select $n$ real orthonormal eigenvectors of $L^U$, denoted by $V^U_i$, where $V^U_i$ is the (right) eigenvector corresponding to eigenvalue $\lambda^U_i$ (cf. Theorem \dref{T-properties}d). \item the normalized eigenvector corresponding to $\lambda^U_{1}=0$ is ${V^U_{1}}\text{=}\frac{1}{\sqrt{n}}\mathbf{1}_n$. \end{itemize} it follows that $L^U$ can be diagonalized, with \begin{equation*} L^U=V^U\Lambda^U {V^U}^T \end{equation*} where $V^U$ is the matrix of (right) orthonormal real eigenvectors of $L^U$ and $\Lambda^U$ is the diagonal matrix of eigenvalues of $L^U$ (see Appendix \dref{App-decomp}). Therefore we have \begin{eqnarray*} e^{-L^U \cdot t}&=&e^{-\left( V^U \Lambda^U (V^U)^T \right) t} \\& =& V^U e^{-\Lambda^U t} (V^U)^T=e^{-\lambda^U_1 t}V^U_1 (V^U)_1^T+e^{-\lambda^U_2 t}V^U_2 (V^U_2)^T+.......+e^{-\lambda^U_n t} V^U_n (V^U_n)^T \end{eqnarray*} Since ${V^U_{1}}\text{=}\frac{1}{\sqrt{n}}\mathbf{1}_n$ we can write: \begin{equation*} x^{(h)}(t)= e^{-L t}x(0) )=\frac{1}{n}{{\mathbf{1}_n}^{T}}x(0)\mathbf{1}_n + \sum_{i=2}^n e^{-\lambda^U_i t}((V^U_i)^T x(0)) V^U_i \end{equation*} or \begin{equation}\dlabel{eq-xh} x^{(h)}(t)=\alpha \mathbf{1}_n + \sum_{i=2}^n e^{-\lambda^U_i t}((V^U_i)^T x(0)) V^U_i \end{equation} Since $\lambda^U_i>0$ forall $i>1$ we have \begin{equation}\dlabel{eq-x1} x^{(h)}_\infty=\underset{t\to \infty }{\mathop{\lim }}\,x^{(h)}(t)=\frac{1}{n}{{\mathbf{1}_n}^{T}}x(0)\mathbf{1}_n=\alpha \mathbf{1}_n \end{equation} with \ \begin{equation}\dlabel{eq-alphaG} \alpha =\frac{1}{n}\sum_{i=1}^n x_i(0) \end{equation} the \emph{average of the initial states}. {\hfill $\Box$}{\cmd{PE}} \subsubsection{Scaled influence} Let $G^S$ be a strongly connected interactions (visibility) graph with scaled influences corresponding to $G^U$. Then the Laplacian $L^S$ has the following properties (cf. Appendix \dref{ScaledGraph}). \begin{itemize} \item The eigenvalues of $L^S$ are also the eigenvalues of $\Gamma$, the normalized Laplacian of $G^U$, a real symmetric matrix, as defined in section \dref{Def}, equation (\dref{Def-Gamma}) \item All eigenvalues of $L^S$, denoted by $\lambda^S_i$, are real and non-negative \item There is a single zero eigenvalue, $\lambda^S_1=0$, and all remaining eigenvalues are strictly positive \item The eigenvectors of $L^S$ relate to the the eigenvectors of the real symmetric matrix $\Gamma$ by: \begin{equation*} V_i^S=\Delta^{-1/2} V_i^\Gamma \end{equation*} where $V_i^S$ and $V_i^\Gamma$ correspond to the eigenvalue $\lambda_i^S=\lambda_i^\Gamma$ and $\Delta$ is the degree matrix associated with the undirected graph $G^U$ \item Since $\Gamma$ is real and symmetric, one can select $V_i^\Gamma$, for all $i$, s.t $V^\Gamma$ is real and orthonormal, where $V_i^\Gamma$ is the $i'th$ column of $V^\Gamma$ (cf. Theorem \dref{T-properties}d). Since $\Delta$ is real and invertible it follows that the corresponding $V^S$ is a matrix of normalized real right eigenvectors of $L^S$ \item $V_1^S$ corresponding to $\lambda^S_1=0$ is $\displaystyle \frac{1}{\sqrt{n}}\mathbf{1}_n$ \item $L^S$ is diagonizable, thus it can be written as \begin{equation*} L^S=V^S \Lambda^S (V^S)^{-1} \end{equation*} where $\Lambda^S $ is a diagonal matrix, s.t. $\Lambda^S_{ii}=\lambda^S_i $ and $V^S$ is the matrix of normalized real right eigenvectors of $L^S$ \item If we denote $(V^S)^{-1}$ by $(W^S)^T$, i.e. $(W^S)^T=(V^S)^{-1}$, then \begin{itemize} \item Each row of $(W^S)^T$ is a left eigenvector of $L^S$ \item The first row of $(W^S)^T$, denoted by $(W^S_1)^T$, is a left eigenvector of $L^S$ corresponding to $\lambda_1^S=0$ satisfying $(W^S_1)^T V^S_1=1$, where $V^S_1$ is the normalized right eigenvector corresponding to $\lambda_1^S=0$ \end{itemize} \item According to Theorem \dref{T-WsT1} in Appendix \dref{ScaledGraph} \begin{equation}\dlabel{eq-Ws1T} (W^S_1)^T= \frac{\sqrt{n} \cdot \mathbf{d}^T}{\sum_{i=1}^n d_i} \end{equation} where $\mathbf{d}$ is a vector of degrees in the graph $G^U$, $\mathbf{d}_i=d_i$, and $d_i$ is the degree of vertex $i$ in $G^U$. \end{itemize} \LB{L-scaledConverge} The value of the \emph{asymptotic consensus state} $\alpha$ for a strongly connected digraph $G^S$ representing scaled influences, is in the convex hull of the initial states $x(0)$ and is given by \begin{equation*} \alpha= \frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i} \end{equation*} where $\mathbf{d}$ is the vector of degrees in the undirected graph $G^U$ corresponding to $G^S$. \LE {\em Proof:\/\ }\cmd{ PB} We have \begin{equation*} L^S=V^S \Lambda^S (W^S)^T \end{equation*} where $(W^S)^T=(V^S)^{-1}$ Thus \begin{eqnarray*} x^{(h)}(t)&=&e^{-L^S t} x(0)\\ & = & e^{-\lambda^S_1 t}(W^S_1)^{T}x(0)V^S_1+e^{-\lambda^S _2t}(W^S_2)^T x(0) V^S_2+.......+e^{-\lambda^S_n t}(W^S_n)^{T} x(0)V^S_n\\ & = & (W^S_1)^{T}x(0)V^S_1 + \sum_{i=2}^N e^{-\lambda^S_i t}(W^S_i)^{T}x(0)V^S_i\\ & = & \frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i} \mathbf{1}_n+ \sum_{i=2}^N e^{-\lambda^S_i t}(W^S_i)^{T}x(0)V^S_i \end{eqnarray*} where we used \begin{itemize} \item $\lambda^S_1=0$ \item $\displaystyle V^S_1 = \frac{1}{\sqrt{n}}\mathbf{1}_n$ \item $(W^S_1)^T$ from eq. (\dref{eq-Ws1T}) \end{itemize} Since $ \lambda^S_i>0 \quad \forall i\geq 2$ we have for $t \rightarrow \infty$ \begin{equation}\dlabel{eq-x1s} x^{(h)}_\infty=\underset{t\to \infty }{\mathop{\lim }}\,x^{(h)}(t)=\frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i} \mathbf{1}_n =\alpha \mathbf{1}_n \end{equation} Thus, $\displaystyle \alpha= \frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i}$ is the asymptotic consensus value for dynamics with scaled influences and no external input. {\hfill $\Box$}{\cmd{PE}} Lemma \dref{L-scaledConverge} holds for any visibility graph, $G^U$. If we let the graph be complete, then we have $d_i=n-1; \quad i=1,..,n$ and thus $\displaystyle \alpha= \frac{ 1}{n} {\sum_{i=1}^n x(0)}$, i.e. the asymptotic consensus value, for complete graphs with scaled influence, is the average of the initial states. The above results can be summarized by the following theorem: \TB{T-ConsensusValue} The value of the \emph{asymptotic consensus state}, $\alpha$, of $n$ agents with a connected visibility (interactions) graph is \begin{alphlist} \titem{a} the average of the initial states if the influence is uniform or if the influence is scaled and the visibility graph is complete. \titem{b} the weighted average of the initial states, $\displaystyle \alpha= \frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i}$, if the influence is scaled and the visibility graph is incomplete, where \begin{itemize} \item $d_i$ is the degree of vertex $i$ in $G^U$ \item $\mathbf{d}^T$ is the vector of degrees in $G^U$, i.e. $\mathbf{d}^T=(d_1 d_2 ....d_n)$ \end{itemize} \end{alphlist} \TE \msubsection{Input induced group dynamics}{MoveAgreement} Next, consider the general form of the input-related part of the group dynamics, $x^{(u)}(t)$, given by eq. (\dref{gen-xu}), where $L, B$ and $u$ are constant in the time interval $[0,t]$. \begin{equation}\dlabel{gen-xu} x^{(u)}(t)= \int_{0}^t e^{-L(t-\tau)}B u d\tau = \int_0^{t} e^{-L\nu}B u d\nu \end{equation} Eq. (\dref{gen-xu}) holds for both the uniform and the scaled influence, i.e $L=L^U$ or $L=L^S$. \begin{itemize} \item For the \textbf{uniform influence case}, since $L^U$ is symmetric we can use again the Spectral theorem and decompose (\dref{gen-xu}) into \begin{equation}\dlabel{Sym-xu} x^{(u)}(t)= \sum_{i=1}^n\left [ \int_0^{t} e^{-\lambda^U_i \nu}V^U_i (V^U_i)^T d\nu \right ] B u \end{equation} \item For \textbf{any} visibility graph with \textbf{scaled influence }, using the properties of $L^S$, we can write \begin{equation}\dlabel{Scaled-xu} x^{(u)}(t)= \sum_{i=1}^n \left [ \int_0^t e^{-\lambda^S_i \nu}V^S_i (W^S_i)^T d\nu \right ] B u \end{equation} \end{itemize} Since for both $L^U$ and $L^S$, representing connected graphs, there is a single zero eigenvalue and the remaining eigenvalues are positive, we can decompose $x^{(u)}(t)$ in two parts: \begin{equation} x^{(u)}(t)= x^{(a)}(t)+ x^{(b)}(t) \end{equation}\dlabel{eq-xu} where \begin{itemize} \item $x^{(a)}(t)$ is the zero eigenvalue dependent term, representing the movement in the agreement space \item $x^{(b)}(t)$ is the remainder, representing the deviation from the agreement space \end{itemize} \subsubsection{Movement along the agreement subspace } \paragraph{The uniform case}\mbox{} We have \begin{equation}\dlabel{Uni-xua} x^{(a)}(t) = \int_0^t e^{-\lambda^U _1 \nu} \text{} V^U_1 (V^U_1)^T B u d\nu = V^U_1 (V^U_1)^T B u t = \frac{n_l}{n} \text{} u \text{} t\mathbf{1}_n \end{equation} where $n_l$ is the number of leaders and we have used $V^U_1=\frac{1}{\sqrt{n}} \mathbf{1}$ and $\mathbf{1}^T B=n_l$. Therefore: \LB{L-firstSym} Consider a group of $n$ agents, forming a connected interactions graph, and moving according to (\dref{x-1D}) with uniform influences. If there are $n_l$ agents that receive an exogenous velocity control $u$, the entire group will move collectively with a velocity $\displaystyle \frac{n_l}{n} u $. \LE \paragraph{Scaled case}\mbox{} For the scaled case we have \begin{equation}\dlabel{Scaled-xua} x^{(a)}(t) = \int_0^t e^{-\lambda^S _1 \nu} \text{} V^S_1 (W^S_1)^T B u d\nu = V^S_1 (W^S_1)^T B u t = \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i}\text{} u \text{} t \text{} \mathbf{1}_n \end{equation} Substituting in (\dref{Scaled-xua}) $V^S_1= \displaystyle \frac{1}{\sqrt{n}} \mathbf{1}_n$ and $(W^S_1)^T$ from (\dref{eq-Ws1T}) we obtain \begin{equation}\dlabel{Scaled-beta} x^{(a)}(t) = \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i} \text{} u \text{} t \text{} \mathbf{1}_n \end{equation} \LB{L-firstScaled} Consider a group of $n$ agents, forming a strongly connected interactions graph with scaled influences, with some of the agents being leaders, i.e. detecting the exogenous velocity control $u$. If each agent moves according to (\dref{x-1D}) then the entire group will move collectively with a velocity $\displaystyle \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i} u$, where $N^l$ is the set of leaders and $d_i$ is the number of edges entering $i$. \LE We see from (\dref{Scaled-beta}) that if the the influences are scaled and \begin{enumerate} \item the visibility graph is complete then the collective velocity of the group reduces to $\displaystyle \frac{n_l}{n} u $, same as for the uniform case. \item the visibility graph is incomplete then the collective velocity of the group is a function not only of the number of leaders but also of the number of links connecting the leaders to followers \end{enumerate} \textbf{\textsf{Example}}: We illustrate the impact of leader selection on the collective velocity, when the visibility graph is incomplete and scaled influence is used, by considering the two configurations shown in Fig. \dref{ChangeLeader}, with identical $G^U$, but different leader. \begin{figure} \begin{center} \includegraphics[scale=0.5]{ChangeLeader.jpg} \caption{Same $G^u$ with different leader }\dlabel{ChangeLeader} \end{center} \end{figure} Based on Lemma \dref{L-firstScaled}, when agent 5 is the leader the group will move with velocity $\displaystyle \frac{1}{12} u$, while when the leader is agent 3 the collective velocity increases to $\displaystyle \frac{1}{4} u$. Note that when uniform influence is employed, the collective velocity depends only on the \emph{number of leaders}. Thus, in both above configurations, the collective velocity is $\displaystyle \frac{1}{5} u$ \subsubsection{Deviations from the agreement subspace }\label{DevProp} Consider now the remainder $x^{(b)}(t)$ of the input-related part, i.e. the part of $x^{(u)}(t)$ containing all eigenvalues of $L$ other than the zero eigenvalue and representing the agents' state deviation from the agreement subspace. The geometric meaning of deviations is elaborated in section \dref{Dev-mean}. In the sequel we will need the following definitions: \begin{definition}Two agents $i,j$ in a network $G$ are said to be \emph{equivalent} if there exists a Leaders-Followers Preserving Permutation $\Pi$ such that $\Pi(i) = j$,$\Pi(j) = i$ and $\Pi(G) = G$ \end{definition} \begin{definition} A Leaders-Followers Preserving Permutation $\Pi$ is a permutation of agents labeling such that $\Pi(leader)$ is a leader and $\Pi(follower)$ is a follower for all leaders and followers. \end{definition} \paragraph{Uniform case}\mbox{} We have \begin{equation*} x^{(b)}(t)= \left [ \sum_{i=2}^n \int_0^t \left (e^{-\lambda^U _i \cdot \nu} \right ) V^U_i (V^U_i)^T d\nu \right ] B u \end{equation*} Thus \begin{equation}\dlabel{Uni-xub} x^{(b)}(t)= \left[ \sum_{i=2}^n \frac{1}{\lambda^U_i}(1-e^{-\lambda^U_i t})V^U_i (V^U_i)^T \right ] B u \end{equation} Since all eigenvalues $\lambda^U_i \text{ for } i \geq 2$ are strictly positive, $x^{(b)}(t)$ converges asymptotically to a time independent vector, denoted by $\varrho$, given by: \begin{equation}\dlabel{Uni-Delta} \varrho = \left[ \sum_{i=2}^n \frac{1}{\lambda^U_i}V^U_i (V^U_i)^T \right ] B u \end{equation} The quantity $\varrho$ represents the vector of asymptotic deviations of the agents from the agreement subspace. \TB{T-DevSum} The asymptotic deviations of all agents, with uniform interactions, sum to zero \begin{equation}\dlabel{eq-sumb} \sum_{i=1}^n \varrho_i = 0 \end{equation} where $\varrho_i$ is the deviation of agent $i$. \TE {\em Proof:\/\ }\cmd{ PB} Consider eq. (\dref{1D-vec}) with $L=L^U$ and multiply it from the left by $\textbf{1}^T$. Recalling that $L^U$ has a left eigenvector $\textbf{1}^T$ corresponding to $\lambda^U_1=0$, we obtain \begin{equation*} \sum_{i=1}^n \dot{x}_i(t) = n_l u \end{equation*} and thus, for all $t$ \begin{equation}\dlabel{eq-sumt} \sum_{i=1}^n x_i(t) = n_l u t + \sum_{i=1}^n x_i(0) \end{equation} On the other hand, recalling that \begin{equation}\dlabel{eq-xt} x(t)= x^{(h)}(t)+ x^{(a)}(t) + x^{(b)}(t) \end{equation} multiplying (\dref{eq-xt}) from the left by $\mathbf{1}_n^T$ and letting $t\rightarrow \infty$, we have: \begin{equation}\dlabel{eq-sumtt} \sum_{i=1}^n x_i(t \rightarrow \infty) = n \alpha + n_l u t + \sum_{i=1}^n \varrho_i \end{equation} Substituting for $\alpha$ its value from eq. (\dref{eq-alphaG}) and comparing equations (\dref{eq-sumt}) and (\dref{eq-sumtt}) we obtain the required result (\dref{eq-sumb}). {\hfill $\Box$}{\cmd{PE}} In general, agents have non-equal deviations, but there are some special cases, detailed in Theorem \dref{T-SpecialDevUni}. \TB{T-SpecialDevUni} \begin{alphlist} \titem{a} Equivalent agents have the same deviation \titem{b} In a fully connected network, all followers have the same asymptotic deviation and all leaders have the same asymptotic deviation, with opposite sign to followers' deviation. \titem{c} If all agents are leaders, i.e. $n_l=n$, then all asymptotic deviations are zero, i.e. $\varrho_i=0, \text{ } \forall i$ \end{alphlist} \TE {\em Proof:\/\ }\cmd{ PB} \\ \tref{a} Equivalent agents follow the same equation, therefore have the same deviation.\\ \tref{b} In a fully connected network, all followers are equivalent to each other and all leaders are equivalent to each other. Thus all followers have the same asymptotic deviation and all leaders have the same asymptotic deviation (different from the followers). Since the sum of all asymptotic deviations is zero, eq. (\dref{eq-sumb}), the deviations of the followers and of the leaders have opposite signs.\\ \tref{c} If $n_l=n$, then $B=\mathbf{1}_n$. Since $V^U_k$ is an eigenvector of the Laplacian $L^U$ with eigenvalue $\lambda^U_k$, we have $L^U V^U_k = \lambda^U_k V^U_k$ or \begin{equation}\label{eq-Vuk} V^U_k = \frac {1}{\lambda^U_k}L^U V^U_k \end{equation} Substituting (\dref{eq-Vuk}) in equation (\dref{Uni-Delta}) we obtain: \begin{equation*} \varrho = \left( \sum_{k=2}^n \frac{1}{(\lambda^U_k)^3} (L^U V^U_k) ((V^U_k)^T (L^U)^T) \right) \mathbf{1}_n u = \mathbf {0}_n \end{equation*} since $(L^U)^T \mathbf{1}_n = \mathbf {0}_n$. {\hfill $\Box$}{\cmd{PE}} \paragraph{Scaled case}\mbox{}\\ Following the same procedure as above, with the corresponding decomposition of $L=L^S$, we obtain the following expression for $x^{(b)}(t)$, in the scaled case: \begin{equation}\dlabel{eq-xbtScaled} x^{(b)}(t)= \left[ \sum_{i=2}^n \frac{1}{\lambda^S_i}(1-e^{-\lambda^S_i t} ) V^S_i (W^S_i)^T \right ] B u \end{equation} and since $\lambda^S_i; \quad i \geq 2$ are positive \begin{equation}\label{eq-devScaled} \varrho=\underset{t\to \infty }{\mathop{\lim }}\,x^{(b)}(t)=\left [ \sum_{i=2}^n \frac{1}{\lambda^S_i} V^S_i (W^S_i)^T \right ] B u \end{equation} Thus, here again $x^{(b)}(t)$, converges asymptotically to a time-independent vector, $\varrho$, given by (\dref{eq-devScaled}) and representing asymptotic deviations from the agreement subspace. \TB{T-DevSumScaled} The \textbf{weighted sum} of the asymptotic deviations of all agents, with scaled pair-wise interactions, is zero \begin{equation}\dlabel{eq-sumbs} \sum_{i=1}^n d_i \varrho_i = 0 \end{equation} where $\varrho_i$ is the deviation of agent $i$ and $d_i$ is the number of edges entering $i$. \TE {\em Proof:\/\ }\cmd{ PB} Multiplying eq. (\dref{1D-vec}), where $L=L^S$, from the left by $(W^S_1)^T$ and integrating, we obtain \emph{for any $t$} \begin{equation}\dlabel{eq-xtScaled} \mathbf{d}^T \cdot x(t) = \mathbf{d}^T \cdot B u t +\mathbf{d}^T \cdot x(0) \end{equation} where $\mathbf{d}$ is the vector of degrees in the corresponding $G^U$ and we used \begin{itemize} \item $(W^S_1)^T L^S = \mathbf{0}_n^T$ \item $\displaystyle (W^S_1)^T=\frac{\sqrt{n} \mathbf{d}^T}{\sum_{i=1}^{n} d_i}$ \end{itemize} Considering now $t \rightarrow \infty$, we can write $x(t \to \infty)$ from eq. (\dref{eq-xt}) as \begin{equation}\label{xt-inf} x(t \to \infty) = \frac{\mathbf{d}^T x(0)}{\sum_{i=1}^{n} d_i} \mathbf{1}_n + \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i} u t \mathbf{1}_n + \varrho \end{equation} where we used Lemma \dref{L-scaledConverge} and Lemma \dref{L-firstScaled}.\\ Multiplying (\dref{xt-inf}) from the left by $\mathbf{d}^T$ we obtain \begin{equation}\label{xt-inf2} \begin{split} \mathbf{d}^T \cdot x(t) & = \frac{\mathbf{d}^T x(0)}{\sum_{i=1}^{n} d_i} \sum_{i=1}^{n} d_i + \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i} u t \sum_{i=1}^{n} d_i + \mathbf{d}^T \cdot \varrho \\ & = \mathbf{d}^T x(0) + \sum_{i \in N^l} d_i u t + \mathbf{d}^T \cdot \varrho\\ & = \mathbf{d}^T \cdot x(0) + \mathbf{d}^T \cdot B u t + \mathbf{d}^T \cdot \varrho \end{split} \end{equation} Comparing now (\dref{xt-inf2}) with (\dref{eq-xtScaled}) for $t\to\infty$ we immediately obtain the required result (\dref{eq-sumbs}). {\hfill $\Box$}{\cmd{PE}} Theorem \dref{T-DevPropertiesScaled} shows properties of the asymptotic deviations of agents with scaled influences in some special cases. These properties for scaled influences are identical to the corresponding ones for uniform influence. \TB{T-DevPropertiesScaled} $n$ agents with scaled interaction, out of which $n_l$ agents are leaders, satisfy the following: \begin{alphlist} \titem{a} All equivalent agents have the same asymptotic deviation \titem{b} In a fully connected network all followers have the same asymptotic deviation and all leaders have the same asymptotic deviation, with opposite sign to followers' deviation. \titem{c} If all agents are leaders, i.e. $n_l=n$, then $\varrho_i=0, \text{ } \forall i$ \end{alphlist} \TE {\em Proof:\/\ }\cmd{ PB} \\ \tref{a} Equivalent agents follow the same equation, therefore have the same deviation.\\ \tref{b} In a fully connected network, with scaled influences, \begin{itemize} \item $d_i=n-1; \quad \forall i$, thus substituting in (\dref{eq-sumbs}) we obtain $\displaystyle \sum_{i=1}^n \varrho_i=0$ \item all leaders are equivalent and all followers are equivalent, thus all leaders have the same asymptotic deviation, $\varrho_l$, and all followers have the same asymptotic deviation, $\varrho_f$ \end{itemize} Thus \begin{equation*} n_l \varrho_l + n_f \varrho_f = 0 \end{equation*} Thus, $sign(\varrho_l) = - sign(\varrho_f)$ \\ \tref{c} If $n_l=n$, then $B=\mathbf{1}_n$. Since $V^S_k$ is a right eigenvector of the Laplacian $L^S$ with eigenvalue $\lambda^S_k$ and $(W^S_k)^T$ is a left eigenvector with the same eigenvalue, we have \begin{itemize} \item $\displaystyle V^S_k = \frac {1}{\lambda^S_k}L V^S_k$ \item $\displaystyle (W^S_k)^T = \frac {1}{\lambda^S_k} (W^S_k)^T L^S$ \end{itemize} Substituting in equation (\dref{eq-devScaled}) we obtain: \begin{equation*} \varrho = \left( \sum_{k=2}^n \frac{1}{(\lambda^S_k)^3} (L^S V^S_k) ((W^S_k)^T L^S) \right) \mathbf{1}_n u = \mathbf {0}_n \end{equation*} since $L^S\mathbf{1}_n = \mathbf {0}_n$.\\ {\hfill $\Box$}{\cmd{PE}} \paragraph{Illustration of asymptotic deviations for various cases}\mbox{} In this section we illustrate by a few examples the impact of the influence model as well as of the equivalence on the obtained deviations. Due to the construction of the interaction graph with scaled influence, $G^S$, out of the the interaction graph with uniform influence, $G^U$, equivalent nodes in $G^U$ are also equivalent in $G^S$. Consider the graphs in Fig. \dref{fig-NewExamples}. \begin{figure} \begin{center} \includegraphics[scale=0.6]{NewExamples.jpg} \caption{Several Networks (Leaders are squares)}\label{fig-NewExamples} \end{center} \end{figure} Denote by $A^U$ the adjacency matrix corresponding to $G^U$ and by $A^S$ the adjacency matrix corresponding to $G^S$. Then the adjacency matrices for each interactions graph depicted in Fig. \dref{fig-NewExamples}, uniform or scaled, are: \newpage \begin{description} \item[(a)] \begin{multicols}{2} \begin{equation*} A^U=\left [ \begin{matrix} 0 & 1 & 0 & 0 & 1\\ 1 & 0 & 1 & 0 & 0\\ 0 & 1 & 0 & 0 & 1\\ 0 & 0 & 0 & 0 & 1\\ 1 & 0 & 1 & 1 & 0 \end{matrix} \right ] \end{equation*} \begin{equation*} A^S=\left [ \begin{matrix} 0 & \frac{1}{2} & 0 & 0 & \frac{1}{2}\\ \frac{1}{2} & 0 & \frac{1}{2} & 0 & 0\\ 0 & \frac{1}{2} & 0 & 0 & \frac{1}{2}\\ 0 & 0 & 0 & 0 & 1\\ \frac{1}{3} & 0 & \frac{1}{3} & \frac{1}{3} & 0 \end{matrix} \right ] \end{equation*} \end{multicols} \item[(b)] \begin{multicols}{2} \begin{equation*} A^U=\left [ \begin{matrix} 0 & 1 & 0 & 0 & 0\\ 1 & 0 & 0 & 1 & 1\\ 0 & 0 & 0 & 1 & 1\\ 0 & 1 & 1 & 0 & 0\\ 0 & 1 & 1 & 0 & 0 \end{matrix} \right ] \end{equation*} \begin{equation*} A^S=\left [ \begin{matrix} 0 & 1 & 0 & 0 & 0\\ \frac{1}{3} & 0 & 0 & \frac{1}{3} & \frac{1}{3}\\ 0 & 0 & 0 & \frac{1}{2} & \frac{1}{2}\\ 0 & \frac{1}{2} & \frac{1}{2} & 0 & 0\\ 0 & \frac{1}{2} & \frac{1}{2} & 0 & 0 \end{matrix} \right ] \end{equation*} \end{multicols} \item[(c)] \begin{multicols}{2} \begin{equation*} A^U=\left [ \begin{matrix} 0 & 1 & 1 & 1 & 1\\ 1 & 0 & 1 & 1 & 1\\ 1 & 1 & 0 & 1 & 1\\ 1 & 1 & 1 & 0 & 1\\ 1 & 1 & 1 & 1 & 0 \end{matrix} \right ] \end{equation*} \begin{equation*} A^S=\left [ \begin{matrix} 0 & 0.25 & 0.25 & 0.25 & 0.25\\ 0.25 & 0 & 0.25 & 0.25 & 0.25\\ 0.25 & 0.25 & 0 & 0.25 & 0.25\\ 0.25 & 0.25 & 0.25 & 0 & 0.25\\ 0.25 & 0.25 & 0.25 & 0.25 & 0 \end{matrix} \right ] \end{equation*} \end{multicols} \end{description} Denoting now by $\varrho^U$ the deviations vector for the uniform case, by $\varrho^S$ the deviations vector for the scaled case and using the input $u =1$ in all examples we obtain: \begin{description} \item[(a)] Node 5 is the leader, nodes 1 and 3 are equivalent, the others have no equivalents. \begin{multicols}{2} \begin{equation*} \varrho^U=\left [ \begin{matrix} -0.06\\ -0.16\\ -0.06\\ 0.04\\ 0.24 \end{matrix} \right ] \end{equation*} \begin{equation*} \varrho^S=\left [ \begin{matrix} -0.2526\\ -0.5142\\ -0.2526\\ 0.2952\\ 0.5812 \end{matrix} \right ] \end{equation*} \end{multicols} We see in this example that \begin{itemize} \item In both cases, uniform and scaled influence, the asymptotic deviations of the equivalent agents' 1 and 3, are identical \item $\displaystyle \sum_{i=1}^n \varrho^U_i = 0$ \item $\displaystyle \sum_{i=1}^n \varrho^S_i \neq 0$ \item $\displaystyle \sum_{i=1}^n d_i \varrho^S_i = 0$ where $d_i$ is the $i'th$ element of $\mathbf{d}^T=[2 \quad 2\quad 2 \quad 1 \quad 3]$ \end{itemize} \item[(b)] The leaders, nodes 4 and 5 are equivalent, the others have no equivalent \begin{multicols}{2} \begin{equation*} \varrho^U=\left [ \begin{matrix} -0.5200\\ -0.1200\\ 0.0800\\ 0.2800\\ 0.2800 \end{matrix} \right ] \end{equation*} \begin{equation*} \varrho^S=\left [ \begin{matrix} -0.6857\\ -0.3368\\ 0.0284\\ 0.4098\\ 0.4098 \end{matrix} \right ] \end{equation*} \end{multicols} In this example again \begin{itemize} \item In both cases, uniform and scaled influence, the asymptotic deviations of the equivalent agents' 4 and 5, are identical \item $\displaystyle \sum_{i=1}^n \varrho^U_i = 0$ \item $\displaystyle \sum_{i=1}^n \varrho^S_i \neq 0$ \item $\displaystyle \sum_{i=1}^n d_i \varrho^S_i = 0$ where $d_i$ is the $i'th$ element of $\mathbf{d}^T=[1 \quad 3 \quad 2 \quad 2 \quad 2]$ \end{itemize} \item[(c)] Nodes 4 and 5 are leaders. Clearly they are equivalent, and so are nodes 1, 2, 3. \begin{multicols}{2} \begin{equation*} \varrho^U=\left [ \begin{matrix} -0.0800\\ -0.0800\\ -0.0800\\ 0.1200\\ 0.1200 \end{matrix} \right ] \end{equation*} \begin{equation*} \varrho^S=\left [ \begin{matrix} -0.3200\\ -0.3200\\ -0.3200\\ 0.4800\\ 0.4800 \end{matrix} \right ] \end{equation*} \end{multicols} In this example, as before: \begin{itemize} \item equivalent nodes have identical deviations, for both uniform and scaled influences \item $\displaystyle \sum_{i=1}^n \varrho^U_i = 0$, \end{itemize} but also $\displaystyle \sum_{i=1}^n \varrho^S_i = 0$. This is due to the completeness of the graph, as stated in Theorem \dref{T-DevPropertiesScaled}. Moreover, we note that in this example, where the visibility graph is complete, the dispersion of agents along the alignment line, i.e. the distance between the position of leaders to the position of followers, is 4 times larger in the scaled case than in the uniform case. This is due to the following holding \textbf{for complete graphs}; \begin{itemize} \item $\lambda^U_i=n; \quad i=2,....,n$ and thus $\displaystyle \varrho^U = \frac{1}{n}\left [ \sum_{i=2}^n V^U_i (V^U_i)^T \right ] B u$ \item $\displaystyle \lambda^S_i=\frac{n}{n-1}; \quad i=2,....,n$; and thus $\displaystyle \varrho^S=\frac{n-1}{n}\left [ \sum_{i=2}^n V^S_i (W^S_i)^T \right ] B u$.\\ Since $V^S=V^U$ and $(V^S_i)^T=(W^S_i)^T$ we obtain, when the same $u$ is used in both cases, \textbf{$ \varrho^S =(n-1) \varrho^U $} \end{itemize} \end{description} \msec{Two dimensional group dynamics }{LTI-dyn} In this section we derive the asymptotic dynamics of a two-dimensional group of agents, in a time interval $[0,t)$, where the system is time independent and the visibility graph is connected. Denote by $p_i = (x_i,y_i)^T $ the position of agent $i$ at time $t$. Let $x(t)$ denote the $n$-dimensional vector $x(t) = (x_1(t) ... x_n(t))^T$ and similarly for $y(t)$. Let $p(t)$ be the $2n$-dimensional vector $(x^T(t)\quad y^T(t))^T$. Assuming that the two dimensions are decoupled, Eq. (\dref{1D-vec}) holds for each component and thus: \iffalse \begin{equation}\dlabel{eq-dynamics3} \dot{p}( t )=-(I_2 \otimes L ) p(t)+ (I_2 \otimes B) u \end{equation} where \begin{itemize} \item $\otimes$ is the Kronecker product. The Kronecker product of a $p \times q$ matrix $P$ with an $r \times s$ matrix $Q$ is defined as \begin{equation}\label{def-kron} P \otimes Q = \left[ \begin{matrix} P_{11}Q & P_{12}Q & \cdots & P_{1q}Q \\ P_{21}Q & P_{22}Q & \cdots & P_{2q}Q \\ \vdots & \vdots & \cdots & \vdots \\ P_{p1}Q & P_{p2}Q & \cdots & P_{pq}Q\\ \end{matrix} \right] \end{equation} \item $I_2$ is the $2 \times 2$ identity matrix \item $L$ is the Laplacian of the interactions graph, $L=L^U$ for uniform interactions or $L=L^S$ for scaled interactions. \item $B$ is a vector indicating the leaders, i.e. $B_i=1$ if agent $i$ is a leader and $0$ if it is a follower \item $\mathbf{u}$ is a two dimensional exogenous control, $\mathbf{u}=[u_x \text{ } u_y]^T$ \end{itemize} Equivalently, \fi \begin{eqnarray*} \dot{x} ( t ) &=& -L\cdot x(t)+ Bu_x \\ \dot{y}( t ) &=& -L\cdot y(t)+ Bu_y \end{eqnarray*} Applying the results derived in section \dref{Gen-1D}, for one time interval, we can write: \begin{equation}\label{p-decoupled} p(t)= p^{(h)}(t)+p^{(a)}(t)+p^{(b)}(t)= \left [\begin{matrix} x^{(h)}(t)+x^{(a)}(t)+x^{(b)}(t)\\ y^{(h)}(t)+y^{(a)}(t)+y^{(b)}(t) \end{matrix} \right ] \\ \end{equation} where $t$ is the time from the beginning of the interval. Thus, for the $x$ axis, we have the following expressions and for the $y$ axis we have the same with $y$ replacing $x$. \begin{itemize} \item for the \textbf{uniform case} \begin{eqnarray*} x^{(h)}(t) & = &\frac{1}{n}{\mathbf{1}_n^{T}}x(0)\mathbf{1}_n + \sum_{k=2}^n e^{-\lambda^U_k t} ((V^U_k)^T x(0)) V^U_k\\ x^{(a)}(t) & = &\frac{n_l}{n} u_x t \mathbf{1}_n\\ x^{(b)}(t) &= &\left[ \sum_{k=2}^n \frac{1}{\lambda^U_k}(1-e^{-\lambda^U_k t})V^U_k (V^U_k)^T \right ] B u_x \end{eqnarray*} where \begin{itemize} \item $\lambda^U_i; \quad i=1, ...,n$ are the eigenvalues of the Laplacian $L^U$ and $V^U_i; \quad i=1, ...,n $ are the corresponding right eigenvectors, selected such that the eigenvector corresponding to $\lambda^U_1=0$ is $V^U_1=\displaystyle \frac{1}{\sqrt{n}}\mathbf{1}_n$ and $V^U$, the matrix with columns $V^U_i$, is orthonormal. \item we assumed in the expression for $ x^{(a)}(t)$ that there are $n_l$ agents receiving the exogenous input \end{itemize} \item for the \textbf{scaled case} \begin{eqnarray*} x^{(h)}(t) & = & \frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i} \mathbf{1}_n+ \sum_{i=2}^n e^{-\lambda^S_i t}(W^S_i)^{T} x(0) V^S_i\\ x^{(a)}(t) & = & \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i} u_x t \mathbf{1}_n\\ x^{(b)}(t) &= &\left[ \sum_{k=2}^n \frac{1}{\lambda^S_k}(1-e^{-\lambda^S_k t})V^S_k (W^S_k)^T \right ] B u_x \end{eqnarray*} \end{itemize} where \begin{itemize} \item $\lambda^S_i; \quad i=1... n$ are the eigenvalues of the Laplacian $L^S$, s.t. $\lambda^S_1=0$ \item $V^S$ is a matrix whose columns, $V^S_i$, are the \emph{normalized} right eigenvectors of $L^S$. In particular, the normalized right eigenvector corresponding to $\lambda^S_1=0$, is $\displaystyle V^S_1=\frac{1}{\sqrt{n}} \mathbf{1}_n$. \item $(W^S)^T$ is the matrix of left eigenvectors, selected s.t. $(W^S)^T = (V^S)^{-1}$ . The first row of $(W^S)^T$, denoted by $(W^S_1)^T$, is a left eigenvector of $L^S$ corresponding to $\lambda^S_1=0$ and satisfies Theorem \dref{T-WsT1}: \begin{equation*} (W^S_1)^T = \frac{\sqrt{n} \mathbf{d}^T}{\sum_{i=1}^n d_i} \end{equation*} \item $d_i$ are the number of neighbors of node $i$ and $ \mathbf{d}$ is a vector with $d_i$ as its $i'th$ element \end{itemize} \msubsection{Interpretation of the asymptotic deviations in the Euclidean space}{Dev-mean} \subsubsection{Asymptotic position of agent $i$}\label{AsympPos2D} The asymptotic positions of agent $i$, in the two-dimensional space, when an external control $\mathbf{u}=(u_x \quad u_y)^T$ is detected by $n_l$ agents, will be \begin{equation}\label{pAsymp} p_i(t \rightarrow \infty)= \left [\begin{matrix} \alpha_x +\beta u_x t + \gamma_i u_x\\ \alpha_y +\beta u_y t + \gamma_i u_y \end{matrix} \right ] \\ \end{equation} where \begin{itemize} \item $\alpha=(\alpha_x\text{ }\alpha_y)^T$ is the agreement, or gathering, point when there is no external input \item $\beta (u_x\text{ } u_y)^T $ is the collective velocity. \item $\gamma_i (u_x \text{ } u_y)^T $ are the $x$ and $y$ components of the asymptotic deviation of agent $i$ \end{itemize} The values of $\alpha_x, \text{ }\alpha_y, \text{ }\beta$ and $\gamma$, the coefficients of the asymptotic position, are a function of the assumed influence model, as shown in Table \dref{AsymptoticPosCoef} for a general visibility graph. Note that $\gamma_i$ is the deviation factor, i.e. $\gamma_i u_x$ is the deviation of agent $i$ in the $x$ direction and similarly for $y$. \begin{table} \centering \caption{Coefficients of asymptotic position} \label{AsymptoticPosCoef} \begin{tabular}{||c|| c| c||} \hline & \textbf{Uniform influence} & \textbf{Scaled influence} \\ [0.5ex] \hline\hline $\alpha_x$ & $\displaystyle \frac{1}{n} \mathbf{1}_n^T x(0)$ & $\displaystyle \frac{ \mathbf{d}^T x(0)}{\sum_{i=1}^n d_i}$ \\[2ex] \hline $\alpha_y$ & $\displaystyle \frac{1}{n} \mathbf{1}_n^T y(0)$ &$ \displaystyle \frac{ \mathbf{d}^T y(0)}{\sum_{i=1}^n d_i}$ \\[2ex] \hline $\beta$ & $\displaystyle \frac{n_l}{n}$ & $ \displaystyle \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i}$ \\[2ex] \hline $\gamma$ & $\displaystyle \sum_{k=2}^n \left[ \frac{1}{\lambda^U_k}V^U_k (V^U_k)^T \right ] B$ & $\displaystyle \sum_{k=2}^n \left[ \frac{1}{\lambda^S_k}V^S_k (W^S_k)^T \right ] B$ \\[2ex] \hline \end{tabular} \end{table} \subsubsection{Asymptotic deviations} The vector of asymptotic deviations, s.t. $\varrho_i = \gamma_i \mathbf{u}$ is the deviation of agent $i$, in the $(x, y)$ space, from the (moving) consensus $\mathbf{\alpha}+\beta \mathbf{u} t$, where $\mathbf{\alpha} = (\alpha_x, \alpha_y)$. The agents align along a line in the direction of $\mathbf{u}$. The line is anchored at the zero-input gathering, or consensus, point $\mathbf{\alpha}$. Since $\gamma$ is time independent, the asymptotic dispersion of agents along this line is time independent, as illustrated in Fig. \dref{fig-dev}. The swarm moves with velocity $\beta \mathbf{u}$. \begin{figure} \begin{center} \includegraphics[scale=0.7]{Deviations.jpg} \caption{Asymptotic dispersion of agents along the direction of $u$}\label{fig-dev} \end{center} \end{figure} \msubsection{Example of simulation results - Single time interval }{SimEx} A single time interval of a piecewise constant system is equivalent to a time-independent configuration with constant exogenous control and leaders. We consider a network of 5 agents and illustrate the group behaviour, for a constant $u$, both in case of incomplete visibility graph and of complete visibility graph. In these examples, the exogenous control is $\mathbf{u} = (10, 2)$. The initial positions $x(0), y(0)$ were once randomly selected in $[-50,50]$, and kept common for all runs. \subsubsection{Incomplete visibility graph}\label{SimExIncomplete} In the examples in this section, we illustrate the impact of the influence model applied by the agents and of the leader selection on the agents dynamics when the interaction graphs, $G^U$ and $G^S$, are as illustrated in Fig. \dref{GraphExSim}. \begin{figure} \begin{center} \includegraphics[scale=0.6]{UniScaledGraph5n6m.jpg} \caption{Simulated pairwise interaction graph}\label{GraphExSim} \end{center} \end{figure} Fig. \dref{ExSimDyn} shows the emergent dynamics of the agents when agent 5 detects the constant exogenous control, thus is the leader. This example will be named Ex1. In Fig. \dref{ExSimDyn} the leader is colored red and the followers are blue. \begin{figure} \begin{center} \includegraphics[scale=0.55]{UniScaledDyn_Ex1.jpg} \caption{Emergent dynamics in Ex1 with uniform and scaled influences}\label{ExSimDyn} \end{center} \end{figure} The agents are seen to asymptotically align, in both cases, along a line in the direction of $\mathbf{u}$, in this case a line with slope 0.2, as expected. The dots indicate the position of the units at consecutive times, t=1,2,3,... We can also see that, in this example, the collective speed of the agents with scaled influence is considerably lower than that of the agents with scaled influence. While the collective speed in the uniform case is $0.2 |u| $, corresponding to $\displaystyle \beta=\frac{n_l}{n}$ with $n_l=1, n=5$, for the scaled case it is only $0.0833 |u|$, corresponding to $ \displaystyle \beta = \frac{\sum_{i \in N^l} d_i}{\sum_{i=1}^n d_i}$. \\ Fig. \dref{Ex1Dyn} shows a comparative view of the agents' dynamics in Ex1. Here again we see the difference in the collective speed with uniform vs scaled influence, but we also see that the agents' alignment lines are parallel and each is anchored at the corresponding zero-input gathering point. \begin{figure} \begin{center} \includegraphics[scale=0.7]{UniScaledDynEx1.jpg} \caption{Comparative view of emergent dynamics with uniform and scaled influences - Ex1}\label{Ex1Dyn} \end{center} \end{figure} Note however that while for the uniform case the coefficient of the collective speed is a function of only the number of leaders, for the scaled case it is also a function of the topology itself, i.e. of the number and distribution of links. Thus, by selecting now agent 4 instead of 5 as leader, we do not change the speed in the uniform case but increase it three times in the scaled case. This brings the velocities of the agents with scaled influence to be larger than the ones with uniform influence, as illustrated in Fig. \dref{Ex2Dyn} \begin{figure} \begin{center} \includegraphics[scale=0.7]{UniScaledDynEx2.jpg} \caption{Comparative view of emergent dynamics with uniform and scaled influences - Leader is agent 4 - Ex2}\label{Ex2Dyn} \end{center} \end{figure} Fig. \dref{ExSimDev} shows the asymptotic deviation of the agents relative to the moving gathering point, in both leader cases, agent 4 or agent 5. In both cases, agents 1 and 3 were equivalent, therefore had the same deviation, but this is not always the case. For example, if agent 1 is selected as leader, Ex3, there will be no equivalent agents, as shown in Fig \dref{Ex3SimDev}. Therefore, \textbf{equivalence is not preserved under change of leader}. \begin{figure} \begin{center} \includegraphics[scale=0.5]{DevLeader4vsLeader5.jpg} \caption{Impact of leader selection on agents' asymptotic derivations relative to the moving consensus}\label{ExSimDev} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[scale=0.6]{UniScaledDevEx3.jpg} \caption{Asymptotic derivations relative to the moving consensus when agent 1 is leader}\label{Ex3SimDev} \end{center} \end{figure} Another issue to be considered is that of the impact of the influence model on the time of convergence. Fig. \dref{ZeroInpDyn} shows the convergence to consensus for uniform and scaled dynamics, when no exogenous control is applied. We see clearly that the convergence time with scaled influence is longer than that with uniform influence. The convergence time is a function of the first non-zero eigenvalue of the Laplacian, $\lambda_2$. In our examples one has $\lambda^U_2=0.8299$ and $\lambda^S_2=0.5657$. Since in these examples we only change leaders, $G^U$ and $G^S$ and the corresponding $\lambda_2$ do not change. Therefore the time of convergence, is identical for all 3 examples. However, we can say that the time of convergence with scaled influence is always at least the time of convergence with uniform influence, since \begin{itemize} \item the eigenvalues of the Laplacian of $G^S$ are the eigenvalues of the \textbf{normalized Laplacian} of the corresponding graph with uniform influence, $G^U$. If we denote the eigenvalues of the normalized Laplacian by $\phi^U_i; i=1,.,n$ Then $\lambda^S_i=\phi^U_i; \quad i=1,...n$ \item As shown by Butler in \dcite{ButlerPhD}, Theorem 4 \begin{equation*} \frac{1}{d_{max}} \lambda^U_i \leq \phi^U_i \leq \frac{1}{d_{min}} \lambda^U_i \end{equation*} where $d_{max}$ is the maximum degree and $d_{min}$ is the minimum degree of a vertex in $G_U$. \end{itemize} Thus, $\lambda^S_2 \leq \lambda^U_2$ for any graph $G^U$ and corresponding $G^S$. \iffalse If $G^U$ is a \textbf{complete graph} then \begin{itemize} \item $\lambda^U_i = n; \quad i=2,...,n$ \item $d_{max}=d_{min} = n-1$ \end{itemize} and we obtain $\displaystyle \lambda^S_i = \frac{n}{n-1}; \quad i=2,...,n$, as expected. \fi \begin{figure} \begin{center} \includegraphics[scale=0.6]{ZeroInputConvEx2.jpg} \caption{Zero Input dynamics as a function of time}\label{ZeroInpDyn} \end{center} \end{figure} \subsubsection{Complete visibility graph}\label{SimExComplete} In this example we assume a network of 5 agents with complete visibility graphs. Fig. \dref{Ex5nComplete} illustrates the emergent behavior in case of scaled influence vs uniform influence and shows that \begin{itemize} \item the zero input gathering point coincides \item the position of the moving gathering point coincides, thus the collective velocity coincides \item the dispersion of the agents around the moving gathering point is larger when the influence is scaled, in fact exactly 4 times larger, as expected \item the time for convergence to (moving) consensus is larger in case of scaled influence, as expected \end{itemize} \begin{figure} \begin{center} \includegraphics[scale=0.7]{DynUniScaledComp-5nComplete.jpg} \caption{Comparative view of emergent dynamics with uniform and scaled influences when the visibility graph is complete }\label{Ex5nComplete} \end{center} \end{figure} \newpage \msec{Multiple time intervals}{MultiIntEx} In the previous sections we considered a single time interval were the system is time independent, i.e. the visibility graph and the corresponding Laplacian $L$, the leaders and the exogenous control $u$ are constant along the interval. We showed the dependence of the emergent behavior on the influence model, scaled or uniform, in case of complete and incomplete visibility graphs. We now consider a sequence of time intervals, $[t_k, t_{k+1})$, where a new time interval is triggered by changes in one of the system parameters, the broadcast control $u$, the agents detecting the broadcast control, i.e. the leaders, or the visibility graph and the corresponding Laplacian. In order for the group behavior along multiple intervals to be a concatenation of dynamics along single intervals, with the end states of one interval becoming the start states of the next interval, we need to ensure that the visibility graph remains strongly connected. In this section we derive sufficient conditions, which are conditions for never losing friends, i.e. for initially adjacent pairs of agents to remain adjacent. Thus, we require $G_k \subseteq G_{k+1}$ and consider two cases of visibility graphs, each for uniform and scaled influences: \begin{enumerate} \item $G_k$ is complete \item $G_k$ is incomplete \end{enumerate} Changes in the visibility graph are state dependent, i.e. a link $(i,j)$ exists at time $t$ iff $|p_i(t)-p_j(t)| \leq R$, where $R$ is the visibility, or sensing, range. In the next sections we derive \textbf{conditions for \emph{never losing neighbors}} and illustrate their effect by simulations. We show that \begin{enumerate} \item if the initial interactions graph is complete and the sensing range is $R$, then \begin{itemize} \item Follower to follower distance and leader to leader distance are monotonically decreasing, therefore initial links are preserved, independently of the value of $u$, for both uniform influence ( Lemma \dref{L-L2Lpreserve}) and scaled influence (Theorem \dref{T-ScaledComplete}a) \item Leader to follower links can be proven to be maintained only if the exogenous control is limited to \begin{enumerate} \item $|u| \leq nR$, for the \textbf{uniform case}, Theorem \dref{T-L2Fpreserve} \item $\displaystyle |u| \leq \frac{n}{n-1} R$, for the \textbf{scaled case}, Theorem \dref{T-ScaledComplete}b \end{enumerate} \end{itemize} Recalling that in case of complete graphs, all leaders asymptotically move together (one moving gathering point) and all followers move together (at another point gathering point) we note that the distance between any leader to any follower tends to $d_{lf} = (\gamma_l-\gamma_f)|u|$ (cf. section \dref{Dev-mean}) and thus preserving the link requires $d_{lf} \leq R$. Since for the complete visibility case $\gamma_l^S=(n-1)\gamma_l^U$ and $\gamma_f^S=(n-1)\gamma_f^U$ the ratio between the bounds on $u$ shown above becomes evident.\\ In section \dref{LimitedComplete-Ex} we show an example of emergent dynamics when $|u|$ is within bounds and another example where $|u|$ exceeds the derived limit \item if the initial graph is incomplete then \begin{itemize} \item conditions for never losing friends are tightly related to the graph topology.\\ Since an external controller does not know the time-dependent topology these are not useful in practice. \item for a general form of incomplete graph, bounds cannot be derived or are too loose to be useful. \end{itemize} Although useful bounds could not be derived, many simulations show that if the interactions graph starts as an incomplete graph, when inputting $u$ such that $|u| < R$, the agents fast converge to a complete graph, as shown by some examples in section \dref{IncompleteEx}. \end{enumerate} \msubsection{Conditions for maintaining complete graphs }{never-losing2} Denote by $\delta_{ij}$ the distance between two adjacent agents $i$ and $j$ \begin{equation}\label{dist2} \delta_{ij}=\delta_{ji}= |p_i-p_j|= \sqrt{(p_i-p_j)^T(p_i-p_j)} \end{equation} Since the movement of the agents is smooth, a necessary and sufficient condition for the link to be always preserved is $d{\delta}_{ij}/dt \leq 0$ when $\delta_{ij}=R$, or equivalently $d(\delta_{ij}^2)/dt \leq 0$, when $\delta_{ij}=R$. Note that $d(\delta_{ij}^2)/dt$ has the same sign as $d{\delta}_{ij}/dt$ and is defined on all of $\textbf{R}^n$ while $d\delta_{ij}/dt$ is not defined when $p_i=p_j$. We have \begin{equation}\label{deriv-dist} \frac{d(\delta_{ij}^2)}{dt} =2 \delta_{ij} \dot{\delta_{ij}} = 2(p_i-p_j)^T(\dot{p_i}-\dot{p_j}) \end{equation} \subsubsection{Uniform influence}\mbox{}\\ If the visibility graph is a complete graph with uniform influences then $\sigma_{ji}=1; \quad \forall i,j$ and equations (\dref{gen-SelfDyn}), (\dref{gen-LeadDyn}) can be combined and reformulated as \begin{equation}\label{gen-dyn} \dot{p_i}=-n_i p_i+\sum_{k \in N_i} p_k+b_i u \end{equation} where $p_i = (x_i \quad y_i)^T$ , $u =(u_x \quad u_y)^T$, $N_i$ is the neighborhood of agent $i$, $n_i=|N_i|$ and $b_i= 1$ if $i$ is a leader and 0 if $i$ is a follower. Since for a complete graph $n_i=n-1; \quad \forall i$, Eq. (\dref{gen-dyn})can be rewritten as \begin{equation*} \dot{p_i}=-(n-1) p_i+\sum_{k =1; k \neq i}^n p_k + b_i u \end{equation*} and similarly for $\dot{p_j}$, where $n$ is the total number of agents. Thus \begin{equation*} \dot{p}_i-\dot{p}_j=-n(p_i-p_j)+(b_i-b_j) u \end{equation*} Denoting $b_{ij}=(b_i-b_j)$, we have \begin{equation}\dlabel{eq-bij} b_{ij} = \begin{cases}0 ; \text{ if both } i,j \text{ are followers or both are leaders}\\ 1 ; \text{ if } i \text{ is leader and } j \text{ is follower} \end{cases} \end{equation} Substituting in eq. (\dref{deriv-dist}) one obtains \begin{equation}\label{Uni-dist2} \frac{d\delta_{ij}^2}{dt} = -2n \delta_{ij}^2 +2(p_i-p_j)^T b_{ij} u \end{equation} \LB{L-L2Lpreserve} If the visibility graph of $n$ agents with uniform influences is a complete graph, then all leader to leader and follower to follower links are preserved, independently of the externally applied control $u$. \LE {\em Proof:\/\ }\cmd{ PB} If $i$ and $j$ in (\dref{Uni-dist2})are both leaders or both followers, then $b_{ij}=0$ and thus eq. (\dref{Uni-dist2}) becomes \begin{equation*} \frac{d\delta_{ij}^2}{dt} = -2n \delta_{ij}^2 \end{equation*} with the solution \begin{equation*} \delta_{ij}^2 (t) = e^{-2nt} \delta_{ij}^2 (0) \end{equation*} Thus, $\delta_{ij}(t)$ decreases monotonically from the initial condition. {\hfill $\Box$}{\cmd{PE}} We shall consider now the \textbf{case when $i$ is a leader and $j$ is a follower}. \TB{T-L2Fpreserve} If the visibility graph of $n$ agents with uniform influences is a complete graph and the magnitude of the exogenous control is limited to $|u| \leq nR$, then the connection of a leader $i$ and a follower $j$ is never lost. \TE {\em Proof:\/\ }\cmd{ PB} When $i$ is a leader and $j$ is a follower Eq. (\dref{Uni-dist2}) becomes \begin{equation}\label{Uni-distL2F} \frac{d\delta_{ij}^2(t)}{dt} = -2n \delta_{ij}(t)^2 +2(p_i(t)-p_j(t))^T u \end{equation} But $(p_i-p_j)^T u = \langle (p_i-p_j), u \rangle \leq |p_i-p_j||u| =\delta_{ij}|u|$. Consider a time $t_1$ when for the first time $\delta_{ij}(t_1)=R$ holds. Since $|u| \leq n R$ for all $t$, we obtain $\displaystyle \frac{d\delta_{ij}^2}{dt}(t_1) \leq 0$. Therefore, when $i$ is a leader and $j$ is a follower, if $|u| \leq nR$, then $\delta_{ij}(t_1) \leq R$ and by induction this result holds for all $t$. {\hfill $\Box$}{\cmd{PE}} \subsubsection{Scaled influence}\mbox{}\\ If the visibility graph is a complete graph with scaled influences then $\sigma_{ji}=\frac{1}{n-1}; \quad \forall i,j$ and equations (\dref{gen-SelfDyn}), (\dref{gen-LeadDyn}) can be combined and reformulated as \begin{eqnarray}\dlabel{Scaled-dyn} \dot{p_i} & = &\sum_{k \in N_i} \frac{1}{n-1}( p_k-p_i)+b_i u\\ & = & -p_i+ \frac{1}{n-1} \sum_{k=1,k \neq i}^n p_k +b_i u\\ & = & -\frac{n}{n-1} p_i + \sum_{k=1}^n p_k+b_i u \end{eqnarray} and similarly for $\dot{p_j}$. Thus, we have \begin{equation}\dlabel{Deriv-Diff} \dot{p_i}-\dot{p_j}=-\frac{n}{n-1}( p_i-p_j) +b_{ij}u \end{equation} where $b_{ij}$ as in (\dref{eq-bij}). Therefore, multiplying (\dref{Deriv-Diff}) from the left by $2(p_i-p_j)^T$ we obtain: \begin{equation}\dlabel{Scaled-deriv} \frac{d (\delta_{ij}^2)}{dt}=-2 \frac{n}{n-1} \delta_{ij}^2 + 2 b_{ij} (p_i-p_j)^T u \end{equation} The dynamics of $\delta_{ij}(t)$, as expressed by (\dref{Scaled-deriv}), for the case when $i$ and $j$ are both followers or both leaders and for the case when $i$ is a leader and $j$ is a follower are summarized by the following theorem: \TB{T-ScaledComplete} If the visibility graph of $n$ agents with scaled influences is a complete graph, then \begin{alphlist} \titem{a} All follower-to-follower links and all leader-to-leader links are monotonically decreasing from the initial conditions, thus these links are preserved, independently of the exogenous control $u$ \titem{b} If the exogenous control satisfies $|u| \leq \frac{n}{n-1} R$ then all Leader-to-Follower links are preserved \end{alphlist} \TE {\em Proof:\/\ }\cmd{ PB} \tref{a} If $i$ and $j$ are both followers or leaders then $b_{ij}=0$. By substituting in (\dref{Scaled-deriv})and solving the resulting homogenous equation, one obtains \begin{equation*} \delta_{ij}^2(t)=e^{-2 \frac{n}{n-1}t} \delta_{ij}^2(0) \end{equation*} Thus, $\delta_{ij}$ monotonically decreases for any two followers or any two leaders and therefore the link is preserved.\\ \tref{b} If $i$ is a leader and $j$ is a follower then $b_{ij}=1$. Substituting this in eq. (\dref{Scaled-deriv}) and letting again $t_1$ be the first time when $\delta_{ij}(t_1)=R$ we obtain \begin{equation*} \frac{d( \delta_{ij}^2)}{dt}(t_1) \leq -2 \frac{n}{n-1} R^2 + 2R|u| \end{equation*} where we used the inequality for inner products $(p_i-p_j)^T u \leq \delta_{ij} |u|$. If $|u| \leq \frac{n}{n-1} R$ for all $t$, then $\displaystyle \frac{d( \delta_{ij}^2)}{dt}(t_1) \leq 0$, and thus, by induction, the leader-to-follower link is preserved for all $t$. {\hfill $\Box$}{\cmd{PE}} \subsubsection{Simulation examples - Effect of $|u|$ on complete graph preservation}\label{LimitedComplete-Ex} We illustrate the emergent behavior of a group of 6 agents with initially complete visibility graph, $R=50$, $u$ and leaders randomly selected, as shown. In the first example, Ex1, where $u_x, u_y$ have random values in the range $[100, 100]$, at $t=5sec$, the restriction on $|u|$ for the scaled case is not satisfied while for the uniform case it is satisfied. Thus, when the scaled influence is applied, the graph splits in two parts (after ~ 5 sec), leaders forming one component and followers forming the other component. When the split occurs, the agents dynamics simulation is stopped. Thus, in Fig. \dref{fig-LimitedVisDyn-Ex1}, the dynamics with scaled influence (cyan and magenta) stopped soon after the beginning of the run (at t=5 sec) while the dynamics with uniform influence (blue and red) evolved for the whole requested period (40 sec). \begin{figure} \begin{center} \includegraphics[scale=0.6]{GroupDynamics-Ex2.jpg} \caption{Group dynamics with limited visibility and high $|u|$ -Ex1}\label{fig-LimitedVisDyn-Ex1} \end{center} \end{figure} When the range of $u_x, u_y$ is reduced to within the limits, all links are preserved, as illustrated in Ex2, where $u_x \in [-20,20]$, $u_y \in [-10,10]$. The leaders were again randomly selected. \begin{figure} \begin{center} \includegraphics[scale=0.6]{GroupDynamics-Ex3.jpg} \caption{Group dynamics with limited visibility and $|u|$ within limits - Ex2}\label{fig-LimitedVisDyn-Ex2} \end{center} \end{figure} In this case the initial complete graph is preserved for both the scaled and the uniform influence and the agents complete the run in both cases. \msubsection{Conditions for never losing friends when visibility graph is incomplete}{LimitedIncomplete} In this section we show that for a general case of incomplete graphs, the "never losing friends" requirement imposes stringent conditions on the topology. Moreover, we show, by examples, that for specific topologies these conditions are too stringent and the property can be proven under relaxed restrictions.\\ We employ the following notations: \begin{itemize} \item $\textbf{N}^f$ denotes the set of followers \item $\textbf{N}^l$ denotes the set of leaders \item $n_f = |\textbf{N}^f|$ is the number of followers \item $n_l= |\textbf{N}^l|$ is the number of leaders \item $n = n_f+n_l$ is the total number of agents \item the set of followers adjacent to an agent $i$, leader or follower, is denoted by $\textbf{N}_i^f $\\ $\textbf{N}_i^f \subseteq \textbf{N}^f$ \item the set of leaders adjacent to an agent $i$, leader or follower, is denoted by $\textbf{N}_i^l $\\ $\textbf{N}_i^l \subseteq \textbf{N}^l$ \item $n_{il}$ is the number of leaders agent $i$ is connected to, $n_{il}=|\textbf{N}_i^l |$ \item $n_{if}$ is the number of followers agent $i$ is connected to, $n_{if}=|\textbf{N}_i^f |$ \item $\textbf{N}_i$ is the neighborhood of $i$, $\textbf{N}_i = \textbf{N}_i^f \bigcup \textbf{N}_i^l$ \item $n_i$ is the size of the neighborhood of $i$, $n_i = n_{il} + n_{if}$ \item $\textbf{N}_{(ij)}^l$ denotes the set of leaders adjacent to both $i$ and $j$ \item $n_{(ij)l}$ is the number of leaders that have a link to both $i$ and $j$ \item $\textbf{N}_{(ij)}^f$ denotes the set of followers adjacent to both $i$ and $j$ \item $n_{(ij)f}$ is the number of followers that have a link to both $i$ and $j$ \end{itemize} Can we find conditions on the topology and on $|u|$ s.t. any two nodes, $i$ and $j$, initially connected, i.e. satisfying $\delta_{ij}(0) = |p_i(0)-p_j(0)| \leq R$ will remain connected, i.e. will satisfy $\frac{d}{dt}\delta_{ij}^2 \leq 0$ when $\delta_{ij}(t)=R$ ? We consider $t_1$, the first time when for one or more links holds $\delta_{ij}(t_1)=R$ and derive conditions for $\frac{d}{dt}\delta_{ij}^2(t_1) \leq 0$ for each link type and each influence type. \subsubsection{General incomplete topology - Uniform case}\label{IncompleteGeneral} If each agent applies the movement equation with uniform influence, then we have \begin{itemize} \item for followers \begin{equation}\dlabel{ML-FolDynU} \dot{p}_i = -\sum_{k \in \textbf{N}_i^f} (p_i-p_k) - \sum_{k \in \textbf{N}_i^l} (p_i-p_k); \quad i \in \textbf{N}^f \end{equation} \item for leaders \begin{equation}\dlabel{ML-LeadDynU} \dot{p}_i=-\sum_{k \in \textbf{N}_i^l} (p_i-p_k) - \sum_{k \in \textbf{N}_i^f} (p_i-p_k) + u; \quad i \in \textbf{N}^l \end{equation} \end{itemize} \begin{enumerate} \item If $i$ and $j$ are \textbf{\emph{both followers}}, then applying (\dref{ML-FolDynU}) to $i$ and $j$ we obtain \begin{equation}\dlabel{eq-F2F-U} \begin{aligned} \dot{p}_i - \dot{p}_j = & -\sum_{k \in \textbf{N}_i^f} (p_i-p_k) - \sum_{k \in \textbf{N}_i^l} (p_i-p_k) \\ & +\sum_{k \in \textbf{N}_j^f} (p_j-p_k) + \sum_{k \in \textbf{N}_j^l} (p_j-p_k) \end{aligned} \end{equation} Separating now the set of neighbors common to $i$ and $j$ from the set of private neighbors to $i$ or $j$ and using \begin{eqnarray*} \textbf{N}_i^f & = & \textbf{N}_{ij}^f + \textbf{N}_i^f \backslash \textbf{N}_{ij}^f\\ \textbf{N}_i^l & = & \textbf{N}_{ij}^l + \textbf{N}_i^l \backslash \textbf{N}_{ij}^l \end{eqnarray*} and similarly for $j$, we obtain \begin{equation*} \begin{aligned} \dot{p}_i - \dot{p}_j = & -n_{(ij)f} p_i + \sum_{k \in \textbf{N}_{ij}^f} p_k -\sum_{k \in \textbf{N}_i^f \backslash \textbf{N}_{ij}^f} (p_i-p_k) \\ & -n_{(ij)l} p_i + \sum_{k \in \textbf{N}_{ij}^l} p_k -\sum_{k \in \textbf{N}_i^l \backslash \textbf{N}_{ij}^l} (p_i-p_k) \\ & +n_{(ij)f} p_j - \sum_{k \in \textbf{N}_{ij}^f} p_k +\sum_{k \in \textbf{N}_j^f \backslash \textbf{N}_{ij}^f} (p_j-p_k) \\ & + n_{(ij)l} p_j -\sum_{k \in \textbf{N}_{ij}^l} p_k +\sum_{k \in \textbf{N}_j^l \backslash \textbf{N}_{ij}^l} (p_j-p_k)\\ = & -(n_{(ij)f}+n_{(ij)l})(p_i-p_j) - \sum_{k \in \textbf{N}_i^f \backslash \textbf{N}_{ij}^f} (p_i-p_k) -\sum_{k \in \textbf{N}_i^l \backslash \textbf{N}_{ij}^l} (p_i-p_k)\\ & +\sum_{k \in \textbf{N}_j^f \backslash \textbf{N}_{ij}^f} (p_j-p_k)+\sum_{k \in \textbf{N}_j^l \backslash \textbf{N}_{ij}^l} (p_j-p_k) \end{aligned} \end{equation*} Consider now the time $t_1$, the first time when one or more links satisfy $\delta_{ij}(t_1)=R$ and let the considered follower to follower link be among them. Then we have $\delta_{ij}(t_1)=R$ and $\delta_{ik}(t_1) = |p_i(t_1)-p_k(t_1)| \leq R; \quad \forall k \in \textbf{N}_i^f \quad \text{and } k \in \textbf{N}_i^l $ and similarly for $j$. Recalling that $\displaystyle \frac{d}{dt}\delta_{ij}^2 = 2(p_i-p_j)^T ( \dot{p}_i - \dot{p}_j)$ we obtain at \begin{equation}\dlabel{F2Fderiv} \begin{aligned} \frac{1}{2} \frac{d}{dt}\delta_{ij}^2(t_1) = & -(n_{(ij)f}+n_{(ij)l})R^2 +\sum_{k \in \textbf{N}_i^f \backslash \textbf{N}_{ij}^f} (p_i-p_j)^T (p_k-p_i) +\sum_{k \in \textbf{N}_i^l \backslash \textbf{N}_{ij}^l} (p_i-p_j)^T (p_k-p_i)\\ & +\sum_{k \in \textbf{N}_j^f \backslash \textbf{N}_{ij}^f} (p_i-p_j)^T (p_j-p_k)+\sum_{k \in \textbf{N}_j^l \backslash \textbf{N}_{ij}^l} (p_i-p_j)^T (p_j-p_k) \end{aligned} \end{equation} Using now $V_1^T V_2 \leq |V_1| |V_2|$ we can write eq. (\dref{F2Fderiv}) as \begin{equation}\dlabel{F2Fderiv2} \begin{aligned} \frac{1}{2} \frac{d}{dt}\delta_{ij}^2 \leq & -(n_{(ij)f}+n_{(ij)l})R^2 +(n_{if}-n_{(ij)f}) R^2+ (n_{il}-n_{(ij)l}) R^2\\ & + (n_{jf}-n_{(ij)f}) R^2+ (n_{jl}-n_{(ij)l}) R^2\\ =& [ n_i+n_j-3 n_{(ij)f} - 3 n_{(ij)l}]R^2 \end{aligned} \end{equation} where we used \begin{equation*} n_i =n_{if} + n_{il} \end{equation*} and similarly for $j$.\\ Thus, if $n_i+n_j \leq 3 n_{(ij)f} + 3 n_{(ij)l}$ is satisfied then $ \frac{d}{dt}\delta_{ij}^2(t_1) \leq 0$ when $i, j \in \textbf{N}^f$. \item If $i$ and $j$ are \textbf{\emph{both leaders}}, then applying (\dref{ML-LeadDynU}) to $i$ and $j$ we obtain \begin{equation}\dlabel{eq-L2L-U} \begin{aligned} \dot{p}_i - \dot{p}_j = & -\sum_{k \in \textbf{N}_i^f} (p_i-p_k) - \sum_{k \in \textbf{N}_i^l} (p_i-p_k) + u\\ & +\sum_{k \in \textbf{N}_j^f} (p_j-p_k) + \sum_{k \in \textbf{N}_j^l} (p_j-p_k) -u \end{aligned} \end{equation} Since $u$ is common to $i$ and $j$ eq. (\dref{eq-L2L-U}) reduces to eq. (\dref{eq-F2F-U}) and therefore we obtain the same condition for never losing neighbors: if $n_i+n_j \leq 3 n_{(ij)f} + 3 n_{(ij)l}$ is satisfied then $ \frac{d}{dt}\delta_{ij}^2(t_1) \leq 0$ when $i, j \in \textbf{N}^l$ and at $t_1$ $\delta_{ij}=R$ and $\delta_{ik} \leq R$, $\delta_{jk} \leq R \quad \forall k \neq i, k \neq j$ \item If \textbf{\emph{$i$ is leader and $j$ is follower}}, then from (\dref{ML-LeadDynU}) for $i$ and (\dref{ML-FolDynU}) for $j$ we obtain \begin{equation*} \begin{aligned} \dot{p}_i - \dot{p}_j = & -\sum_{k \in \textbf{N}_i^f} (p_i-p_k) - \sum_{k \in \textbf{N}_i^l} (p_i-p_k) + u \\ & +\sum_{k \in \textbf{N}_j^f} (p_j-p_k) + \sum_{k \in \textbf{N}_j^l} (p_j-p_k) \end{aligned} \end{equation*} which, following the same technique as above, reduces to \begin{equation*} \begin{aligned} \dot{p}_i - \dot{p}_j = & -(n_{(ij)f}+n_{(ij)l})(p_i-p_j) +\sum_{k \in \textbf{N}_i^f \backslash \textbf{N}_{ij}^f} (p_i-p_k) -\sum_{k \in \textbf{N}_i^l \backslash \textbf{N}_{ij}^l} (p_i-p_k) + u\\ & +\sum_{k \in \textbf{N}_j^f \backslash \textbf{N}_{ij}^f} (p_j-p_k)+\sum_{k \in \textbf{N}_j^l \backslash \textbf{N}_{ij}^l} (p_j-p_k) \end{aligned} \end{equation*} \end{enumerate} \begin{equation}\dlabel{L2Fderiv} \begin{aligned} \frac{1}{2} \frac{d}{dt}\delta_{ij}^2 \leq & -(n_{(ij)f}+n_{(ij)l})R^2 +(n_{if}-n_{(ij)f}) R^2+ (n_{il}-n_{(ij)l}) R^2 + |u|R\\ & + (n_{jf}-n_{(ij)f}) R^2+ (n_{jl}-n_{(ij)l}) R^2\\ =& [ n_i+n_j-3 n_{(ij)f} - 3 n_{(ij)l}]R^2 +|u|R \end{aligned} \end{equation} Thus, if $3 ( n_{(ij)f} + n_{(ij)l} ) \geq (n_i+n_j) $ and $|u| \leq [ 3 ( n_{(ij)f} + n_{(ij)l} ) - (n_i+n_j)]R$ then for $i \in \textbf{N}^l$ and $j \in \textbf{N}^f$, $\delta_{ij}^2 \leq 0$ when $\delta_{ij}=R$.\\ All of the above results can be summarized by the following theorem: \TB{IncompleteGen} Given a group of agents with connected visibility graph and uniform influence, any link $(i,j)$ satisfying the following conditions will be preserved \begin{enumerate} \item ( $i \in N^f$ and $j \in N^f$) or ($i \in N^l$ and $j \in N^l$) and $n_i+n_j \leq 3 n_{(ij)f} + 3 n_{(ij)l}$, independently of the exogenous control \item if $i \in N^l$ and $j \in N^f$ and $n_i+n_j \leq 3 n_{(ij)f} + 3 n_{(ij)l}$ then an input $u$ that satisfies \begin{equation*} |u| \leq[ 3 (n_{(ij)f}+ n_{(ij)l})-(n_i+n_j)] R \end{equation*} will ensure the link preservation \end{enumerate} where \begin{itemize} \item $n_{(ij)l}$ is the number of leaders that have a link to both $i$ and $j$ \item $n_{(ij)f}$ is the number of followers that have a link to both $i$ and $j$ \item $n_i$ is the number of nodes adjacent to $i$ \item $n_j$ is the number of nodes adjacent to $j$ \end{itemize} \TE Note that these conditions are not useful to us since the controller is unaware of the time varying and random values needed in the quantity limiting the control speed $|u|$, in order to ensure the preservation of all initial visibility links. \paragraph{Effect of assuming a Specific topology on conditions for never losing neighbors}\mbox{}\\ In this section we show that if a specific topology is assumed, then the bounds derived in section \dref{IncompleteGeneral}, for a general incomplete graph with uniform influences, can be tightened. We illustrate the effect on the example shown in Fig. \dref{fig-Ex1-Incomplete}. A more general example, although with some specific features, is shown in Appendix \dref{incomplete}, where all leaders form a complete graph and all followers form a complete graph. \begin{figure} \begin{center} \includegraphics[scale=0.6]{Ex1-Incomplete.jpg} \caption{Ex1 - specific case of incomplete graph}\label{fig-Ex1-Incomplete} \end{center} \end{figure} If we consider link $(4,3)$ and apply theorem \dref{IncompleteGen}, we have: \begin{itemize} \item $n_4=2$ \item $n_3=3$ \item $n_{(34)f}=1$ \item $n_{(34)l}=0$ \end{itemize} Thus the condition $n_3+n_4 \leq 3 n_{(34)f} + 3 n_{(34)l}$ does not hold and there is no $u$ that will ensure that link $(4,3)$ is preserved. However, if we consider the particular structure of the graph we obtain: \begin{equation*} \begin{aligned} \dot{p}_4-\dot{p}_3 = & -(p_4-p_3)-(p_4-p_2)+u + (p_3-p_4)+(p_3-p_2)+ (p_3-p_1)\\ = & -3 (p_4-p_3) + (p_3-p_1) + u \end{aligned} \end{equation*} Using the same technique as above, we obtain \begin{equation*} \frac{1}{2} \frac{d}{dt}\delta_{34}^2 \leq - 2 R^2 + |u| R \end{equation*} Thus, for this particular, incomplete, topology, $|u| \leq 2 R$ ensures the preservation of link $(4,3)$. \subsubsection{ General incomplete topology - Scaled case}\mbox{} If each agent applies the movement equation with scale influence, then we have \begin{itemize} \item for followers \begin{equation}\dlabel{ML-FolDynS} \dot{p}_i = -\frac{1}{n_i}\sum_{k \in \textbf{N}_i^f} (p_i-p_k) -\frac{1}{n_i} \sum_{k \in \textbf{N}_i^l} (p_i-p_k); \quad i \in \textbf{N}^f \end{equation} \item for leaders \begin{equation}\dlabel{ML-LeadDynS} \dot{p}_i=-\frac{1}{n_i}\sum_{k \in \textbf{N}_i^l} (p_i-p_k) -\frac{1}{n_i} \sum_{k \in \textbf{N}_i^f} (p_i-p_k) + u; \quad i \in \textbf{N}^l \end{equation} \end{itemize} \begin{enumerate} \item $i$ and $j$ are \textbf{followers}\\ Applying (\dref{ML-FolDynS}) to $i$ and $j$ one can write: \begin{equation}\label{F2Fs} \dot{p}_i - \dot{p}_j = \frac{1}{n_i}\sum_{k \in \textbf{N}_i^f} (p_k-p_i) +\frac{1}{n_i} \sum_{k \in \textbf{N}_i^l} (p_k-p_i)+\frac{1}{n_j}\sum_{k \in \textbf{N}_i^f} (p_j-p_k) +\frac{1}{n_j} \sum_{k \in \textbf{N}_i^l} (p_j-p_k) \end{equation} \begin{eqnarray*} \frac{d}{dt}\delta_{ij}^2 & = & 2(p_i-p_j)^T ( \dot{p}_i - \dot{p}_j)\\ & = & 2 \left[ \frac{1}{n_i}\sum_{k \in \textbf{N}_i^f} (p_i-p_j)^T (p_k-p_i) +\frac{1}{n_i} \sum_{k \in \textbf{N}_i^l} (p_i-p_j)^T (p_k-p_i) \right ]\\ & + & 2 \left[ \frac{1}{n_j}\sum_{k \in \textbf{N}_j^f} (p_i-p_j)^T (p_j-p_k) +\frac{1}{n_j} \sum_{k \in \textbf{N}_i^l}(p_i-p_j)^T (p_j-p_k) \right ] \end{eqnarray*} Using now \begin{itemize} \item $V_1^T V_2 \leq |V_1| |V_2|$ \item $\delta_{ij} = R$ \item $\delta_{ik} = |p_i-p_k| \leq R; \quad \forall k \in \textbf{N}_i^f \quad \text{and } k \in \textbf{N}_i^f $ and similarly for $j$ \end{itemize} we obtain \begin{equation}\label{dF2Fs} \frac{d}{dt}\delta_{ij}^2 \leq 2 \left[ \frac{1}{n_i} \left ( n_{if} R^2 + n_{il} R^2 \right ) + \frac{1}{n_j} \left ( n_{jf} R^2 + n_{jl} R^2 \right ) \right] \end{equation} Since $n_i = n_{if} + n_{il}$, and similarly for $j$, eq. (\dref{dF2Fs}) becomes \begin{equation}\label{dF2Fs2} \frac{d}{dt}\delta_{ij}^2 \leq 4 R^2 \end{equation} Note that equation \dref{dF2Fs2} does not ensure that $\frac{d}{dt}\delta_{ij}^2 \leq 0$ when the distance between two followers approaches the visibility range $R$. As such it is not useful, since it does not ensure that this distance does not increase beyond $R$. \item $i$ and $j$ are \textbf{leaders}\\ Since $u$ is common to all leaders we obtain the same bound on leader to leader link as on follower to follower link, (\dref{dF2Fs2}) \item $i$ is a \textbf{follower and} $j$ is a \textbf{leader}\\ Following the same procedure as above, we obtain \begin{equation}\label{dL2Fs2} \frac{d}{dt}\delta_{ij}^2 \leq 4 R^2 + 2|u|R \end{equation} Thus, without any assumptions on the topology of the graph, the property of never losing friends when applying the protocol with scaled influence, cannot be proven. \end{enumerate} \paragraph{Specific cases of topology - Uniform vs Scaled influence}\mbox{} Although it seems from the above that scaled influence is weaker than uniform influence in never losing neighbors, we will show here that there are specific cases where visibility link preservation with uniform influence can be proven only under the assumption of certain initial configurations, i.e. under "conditional topology" conditions, while scaled influence relaxes these conditions. We consider the case of a single leader with a single link to a complete sub-graph of followers, as illustrated in Fig. \dref{fig-case1}. \begin{figure} \begin{center} \includegraphics[scale=0.6]{OneLeaderEx.jpg} \caption{Illustration of incomplete graph-case 1}\label{fig-case1} \end{center} \end{figure} and show that \begin{itemize} \item if the uniform protocol is applied, then some very strict constraints on the initial states are required in order to ensure the property of never lose neighbors for $n_f > 2$ \item if the scaled protocol is applied, then for $ |u|\leq \frac{n}{n-1} R$ all initial visibility links are preserved for any $n>2 \quad \text{or} \quad n_f>1$. \end{itemize} \subparagraph{Uniform influence with a-single-leader-to-a-single-follower connection} The topology considered here belongs to the class of incomplete graphs with the followers forming a complete subgraph and the leaders forming a complete subgraph, discussed in Appendix \dref{incomplete}. Assuming the leader to be agent $n$ and its adjacent follower to be agent $n-1$ we have: \begin{itemize} \item $n_{il}=0$ for $i=1,...,n-2$ \item $n_{il}=1$ for $i = n-1$ \item $n_{if}=1$ for $i=n$ \item $n_{if}=n-2$ for $i=1,...,n-1$ \item $n_{(ij)l}=0$, for all $i,j$. \end{itemize} Thus, conditions (\dref{F2F-connect})-(\dref{L2F-preserve}) in Appendix \dref{incomplete} become: \begin{itemize} \item The leader to leader condition (\dref{L2L-connect}) is not applicable \item The condition for follower to follower connection preservation (\dref{F2F-connect}): since $n_{il} + n_{jl} \leq 1$, we obtain : $n_f\geq 1$, obvious. \item The conditions for the leader to follower connection preservation (\dref{L2F-preserve}) yields: $n<4$ (namely $n \leq 3$ or equivalently $n_f\leq 2$) and thus (\dref{u-connect}) yields $|u| \leq R (4-n)$ \end{itemize} Therefore, a single leader with a single connection to followers cannot be proven to drive the followers without losing the connection \textbf{unless the number of followers $n_f \leq 2$} and the exogenous control satisfies $|u| \leq R (4-n)$. \subparagraph{Conditional initial links preservation without limiting the number of followers} By conditional initial links preservation we mean that the initial links can be proven to be maintained only when the \textbf{initial states are limited to certain configurations}. We have shown above that for the single leader with single leader-follower connection, the initial links can be proven to be preserved only when the number of followers is limited to two. Here we show that with certain initial configurations, the restriction on the number of followers is removed. In particular, we show that there exists an exogenous control $u$ such that for certain initial configurations, the link between the leader and the leading-follower is preserved for any number of followers. In the following two lemmas we look at a graph where agent $n$ is \underline{a single leader with a single link} to a follower labelled $n-1$, which will be called "leading follower". We assume that the followers subgraph is initially complete and denote the visibility range by $R$. \LB{lemma1} Suppose that the following initial condition holds: \begin{equation*} \delta_{n-1,i}(0) < \frac{R}{n-1}; \text{ } i=1,...,n-2 \end{equation*} Then for all times $t$ we have that: \begin{equation*} \delta_{ij}(t) < \frac{2R}{n-1}; \text{ } i,j=1,...,n-2 \end{equation*} \LE {\em Proof:\/\ }\cmd{ PB} By the triangle inequality the following holds: \begin{equation*} \delta_{ij}(0) < \frac{2R}{n-1}; \text{ } i,j=1,...,n-2 \end{equation*} The Lemma follows from the fact that \emph{the distance between non-leading followers is monotonically decreasing}. This is seen from the fact that, given that the followers subgraph is complete, we have for $i,j=1,...n-2$: \begin{eqnarray*} \dot{p}_i &=& -(n-1)p_i+\sum_{k=1}^{n-1}p_k \\ \dot{p}_j &=& -(n-1)p_j+\sum_{k=1}^{n-1}p_k \end{eqnarray*} and thus \begin{eqnarray*} \frac{d}{dt} \delta_{ij}^2 (t)&=& 2(p_i(t)-p_j(t))^T(\dot{p}_i(t)-\dot{p}_j(t))\\ &=& -2(n-1) \delta_{ij}^2(t) \end{eqnarray*} with the solution \begin{equation}\dlabel{Conditional-F2F} \delta_{ij}^2 (t)=e^{-2(n-1)t}\delta_{ij}^2(0) \end{equation} {\hfill $\Box$}{\cmd{PE}} \LB{lemma2} Suppose that $|u|\leq \frac{n}{n-1} R$ and that the following initial conditions hold: \begin{eqnarray*} \delta_{n,n-1}(0) & < & R \\ \delta_{n-1,i}(0) & < & \frac{R}{n-1}; \text{ } i=1,...,n-2 \end{eqnarray*} Then for all times $t$ hold: \begin{alphlist} \titem{a} $\delta_{n,n-1}(t) \leq R$ \titem{b} $\displaystyle \delta_{n-1,i}(t) \leq \frac{R}{n-1}; \text{ } i=1,...,n-2$ \end{alphlist} \LE {\em Proof:\/\ }\cmd{ PB} We shall prove the Lemma by contradiction. Suppose \tref{a}, \tref{b} do not hold and let $t_1$ be \textbf{the first time when \tref{a} and/or \tref{b} is contradicted by one or more links}, namely that \begin{eqnarray*} \delta_{n,n-1}(t_1^+) &>& R \\ \text{and/or } \delta_{n-1,i}(t_1^+) & > & \frac{R}{n-1} \end{eqnarray*} Note that until time $t_1$ both \tref{a} and \tref{b} hold for all links and since all $\delta$'s are continuous functions, at time $t_1$ holds $\delta_{n-1,i}(t_1) \leq \frac{R}{n-1}$ and $\delta_{n,n-1}(t_1) \leq R; \text{ } i=1,...,n-2$. Consider \textbf{any one} of the links that contradicts \tref{a} or \tref{b} at time $t_1$. If link $(n,n-1)$ contradicts \tref{a}, then \begin{eqnarray} \delta_{n-1,n}(t_1) &=&R\label{n1n-t1} \\ \delta_{n-1,i}(t_1) & \leq & \frac{R}{n-1}; \quad i= (1, ...., n-2)\label{n1i-t1} \\ \frac{d }{dt}\delta_{n-1,n}^2(t_1)& > & 0\label{not-a-t1} \end{eqnarray} Starting from \begin{eqnarray*} \dot{p}_n &=& p_{n-1}-p_n+u \\ \dot{p}_{n-1} &=& \sum_{i=1}^{n-2} (p_i-p_{n-1})+p_n-p_{n-1} \end{eqnarray*} we obtain \begin{eqnarray*} \frac{d }{dt}\delta_{n-1,n}^2(t_1)& =& 2(p_{n-1}(t_1)-p_n(t_1))^T \left [ \sum_{i=1}^{n-2} \left ( p_i(t_1)-p_{n-1}(t_1) \right ) + 2\left ( p_n(t_1)-p_{n-1}(t_1) \right )-u \right ]\\ & = & -4 \delta_{n,n-1}^2(t_1)+2 \sum_{i=1}^{n-2} \left( (p_n(t_1)-p_{n-1}(t_1))^T (p_{n-1}(t_1)-p_i(t_1)) \right ) +2 (p_n(t_1)-p_{n-1}(t_1))^T u \\ & \leq & -4 R^2+2 (n-2)\frac{R^2}{n-1} + 2 R |u| \end{eqnarray*} where we used (\dref{n1n-t1}), (\dref{n1i-t1}) and the property of inner products $V_1^T V_2 \leq |V_1||V_2|$. Since $ |u|\leq \frac{n}{n-1} R$ , we have \begin{equation*} \frac{d }{dt}\delta_{n-1,n}^2(t_1) \leq -2 R^2 < 0 \end{equation*} contradicting (\dref{not-a-t1}), i.e. the assumption that \tref{a} does not hold. Now suppose that the considered link is $(n-1,i)$ for some follower $i \in 1,...,n-2$. At time $t_1$ holds \begin{eqnarray} \delta_{n,n-1}(t_1) & \leq & R\label{b1-t1} \\ \delta_{n-1,i}(t_1)& = & \frac{R}{n-1}; \quad i \in 1,...,n-2\label{b2-t1} \\ \delta_{n-1,j}(t_1) & \leq & \frac{R}{n-1}; \quad j= 1,...,n-2 ; \quad j \neq i\label{b3-t1} \end{eqnarray} If \tref{b} is contradicted by link $(n-1,i)$ for the first time at $t_1$ then $\displaystyle \frac{d}{dt}\delta_{n-1,i}^2(t_1) > 0 $ will hold. \begin{eqnarray*} \dot{p}_{n-1} &=& \sum_{j=1}^{n-2} (p_j-p_{n-1})+p_n-p_{n-1}\\ & = & -(n-1)p_{n-1}+\sum_{j=1}^{n-1} p_j-p_{n-1} +p_n\\ \dot{p}_i & = & \sum_{j=1}^{n-1} (p_j-p_i)\\ & = & - (n-1) p_i + \sum_{j=1}^{n-1} p_j \end{eqnarray*} \begin{eqnarray*} \frac{d }{dt}\delta_{n-1,i}^2(t) &= & 2(p_{n-1}(t)-p_i(t))^T [-2((n-1)(p_{n-1}(t)-p_i(t))+(p_n(t)-p_{n-1}(t))\\ & \leq & -2(n-1) \delta_{n-1,i}^2+2\delta_{n-1,i} \delta_{n,n-1}(t) \end{eqnarray*} where we used again the inequality for inner products. Considering now the above at time $t_1$ and using(\dref{b1-t1}), (\dref{b2-t1}), we obtain \begin{eqnarray*} \frac{d }{dt}\delta_{n-1,i}^2(t_1) & \leq & 2 \frac{R^2}{n-1}-2 \frac{(n-1)R^2}{(n-1)^2}\\ & \leq & 0 \end{eqnarray*} contradicting the assumption that \tref{b} does not hold at time $t_1$ for link $(n-1,i)$. {\hfill $\Box$}{\cmd{PE}} From the previous two Lemmas it follows that: \TB{cond-preserve} Let agent $n$ be a single leader with a single link to a follower labelled $n-1$. Assume that followers subgraph is initially complete and denote the visibility range by $R$. Suppose that $|u|\leq \frac{n}{n-1} R$ and that the following initial conditions hold: \begin{eqnarray*} \delta_{n,n-1}(0) & < & R\\ \delta_{n-1,i}(0) & < & \frac{R}{n-1}; \text{ } i=1,...,n-2 \end{eqnarray*} Then neighbors are never lost, i.e. all initial links are preserved. \TE \subparagraph{Scaled influence with a-single-leader-to-a-single-follower connection} We assume as before that the followers form a complete graph. The agents are labeled s.t. agent $n$ is the leader and $n-1$ is the leading follower. There are no constraints on the initial conditions, i.e. $\delta_{ij}(0) \leq R; \quad \forall i \sim j$. Recall that all agents apply the scaled protocol \begin{equation*} \dot{p}_i = \frac{1}{n_i}\sum_{j \in N_i} (p_j-p_i)+b_i u \end{equation*} where \begin{itemize} \item $p_i = (x_i \quad y_i)^T$ \item $N_i$ is the neighborhood of $i$ and $n_i=|N_i|$ \item $b_i$ is 1 if $i=n$, i.e. $i$ is the leader, and 0 otherwise \end{itemize} \TB{T-ScaledSingle} Let $n$ agents with scaled influence and with visibility range $R$ have a-single-leader-to-a-single-follower connection and complete followers subgraph. If we label the leader by $n$ and the leading follower by $n-1$, then \begin{alphlist} \titem{a} \underline{for $ i,j=1,...n-2$}, $\delta_{ij}(t)$ is monotonically decreasing, thus if $\delta_{ij}(0) \leq 0 $ then $\delta_{ij}(t) <0$ for all $t$, independently of the external control, $u$ \titem{b} if $ |u|\leq \frac{n}{n-1} R$ and \begin{eqnarray*} \delta_{n,n-1}(0) & < & R \\ \delta_{n-1,i}(0) & < & R; \quad i=1,....n-2 \end{eqnarray*} then \begin{eqnarray*} \delta_{n,n-1}(t) & \leq & R \\ \delta_{n-1,i}(t) & \leq & R; \quad i=1,....n-2 \end{eqnarray*} for all $t$ \end{alphlist} \TE {\em Proof:\/\ }\cmd{ PB} \underline{Property \tref{a}} - As before, we consider $\displaystyle \frac{d}{dt}\delta_{ij}^2(t); \quad i,j=1,...,n-2$ and use \begin{eqnarray*} \frac{d}{dt}\delta_{ij}^2(t) &= & 2(p_i(t)-p_j(t))^T(\dot{p}_i(t)-\dot{p}_j(t)) \\ \dot{p}_i(t) & = & -\frac{n-1}{n-2}p_i(t)+\frac{1}{n-2}\sum_{k=1}^{n-1}p_k(t) \\ \dot{p}_j (t)& = & -\frac{n-1}{n-2}p_j(t)+\frac{1}{n-2}\sum_{k=1}^{n-1}p_k(t) \end{eqnarray*} to obtain \begin{equation*} \frac{d}{dt}\delta_{ij}^2(t)=-2 \frac{n-1}{n-2}\delta_{ij}^2(t) \end{equation*} with the solution \begin{equation}\dlabel{Scaled-F2F} \delta_{ij}^2 (t)=e^{- 2 \frac{n-1}{n-2} t}\delta_{ij}^2(0) \end{equation} which is monotonically decreasing\\ \underline{Property \tref{b}} - We shall prove this property again by contradiction. Suppose that $t_1$ is the first time that this property is contradicted by one or more links, namely an external control $ |u|\leq \frac{n}{n-1} R$ is applied and \begin{eqnarray*} \delta_{n,n-1}(t_1^+) & > & R \\ \text{and/or } \delta_{n-1,i}(t_1^+) & > & R; \quad i=1,....n-2 \end{eqnarray*} Since $t_1$ is the first time that the above holds and all links sizes are continuous functions, for all $t \leq t_1$ property \tref{b} holds, thus \begin{eqnarray*} \delta_{n,n-1}(t_1) & \leq & R \\ \delta_{n-1,i}(t_1) & \leq & R; \quad i=1,....n-2 \end{eqnarray*} Consider any one of the links that contradicts \tref{b} at time $t_1$. \begin{enumerate} \item Assume that the link that contradicts \tref{b} at time $t_1$ is $(n,n-1)$, then let \begin{eqnarray} \delta_{n,n-1}(t_1) & = & R\label{dist-t1-nn1}\\ \delta_{n-1,i}(t_1) & \leq & R; \quad i=1,....n-2\label{dist-t1-n1i} \end{eqnarray} and show that $ \frac{d }{dt}\delta_{n,n-1}^2(t_1) > 0$ \textbf{does not hold}. \begin{eqnarray*} \dot{p}_n &=& p_{n-1}-p_n+u \\ \dot{p}_{n-1} &=& \frac{1}{n-1} \left [ \sum_{j=1}^{n-2}(p_j-p_{n-1})+(p_n-p_{n-1}) \right ] \end{eqnarray*} \begin{equation*} \dot{p}_n- \dot{p}_{n-1}=-2 (p_n-p_{n-1})+\frac{1}{n-1} \sum_{j=1}^{n-2}(p_{n-1}-p_j)+u \end{equation*} Multiplying from left by $2(p_n-p_{n-1})^T$ and using the inequality for inner products, we obtain \begin{equation*} \frac{d }{dt}\delta_{n,n-1}^2(t) \leq -4\delta_{n,n-1}^2(t)+ \frac{2}{n-1}\sum_{j=1}^{n-2} \delta_{n,n-1}(t) \delta_{n-1,j}(t)+2 \delta_{n,n-1}(t)|u| \end{equation*} Since at $t_1$ eqs. (\dref{dist-t1-nn1}), (\dref{dist-t1-n1i}) hold, we have \begin{equation*} \frac{d }{dt}\delta_{n,n-1}^2(t_1) \leq -4R^2+2\frac{n-2}{n-1}R^2+2 R |u| \end{equation*} Thus, if $|u| \leq \frac{n}{n-1}R$, then $ \frac{d }{dt}\delta_{n,n-1}^2(t_1) \leq 0$, contradicting the assumption that statement \tref{b} of the Theorem does not hold for link $(n,n-1)$. \item Consider now a link $(n-1,i)$, for any $i=1,...,n-2$ and show that it cannot be the one that first contradicts \tref{b} at time $t_1$. As for \tref{a}, we have at $t_1$ \begin{eqnarray*} \delta_{n,n-1}(t_1) & \leq & R \\ \delta_{n-1,i}(t_1) & \leq & R; \quad i=1,....n-2 \end{eqnarray*} For any link $j; \quad j \in 1,....,n-2$ assumed to contradict statement \tref{b} of the Theorem, we have to show that when $ \delta_{n-1,j}(t_1) = R$, while \begin{eqnarray} \delta_{n,n-1}(t_1) & \leq & R\label{scaled1-t1} \\ \delta_{n-1,i}(t_1) & \leq & R; \quad i \neq j\label{scaled2-t1} \end{eqnarray} if $|u| \leq \frac{n}{n-1}R$, then $ \frac{d }{dt}\delta_{n-1,j}^2(t_1) > 0$ does not hold, contradicting the assumption that statement \tref{b} of the Theorem does not hold for link $(n-1,j)$. We have \begin{eqnarray*} \dot{p}_{n-1} &=& \frac{1}{n-1} \left [ \sum_{j=1}^{n-2}(p_j-p_{n-1}) +(p_n-p_{n-1}) \right ]\\ \dot{p}_{j} &=& \frac{1}{n-2} \left [ \sum_{k=1}^{n-2}(p_k-p_j) +(p_{n-1}-p_j) \right ] \end{eqnarray*} \begin{eqnarray*} \frac{d}{dt} \delta_{n-1,j}^2(t) &=& 2(p_{n-1}(t)-p_j(t))^T (\dot{p}_{n-1}(t)-\dot{p}_j(t)) \\ &=& \frac{2}{n-1} \left [ \sum_{k=1}^{n-2}(p_{n-1}(t)-p_j(t))^T(p_k(t)-p_{n-1}(t)) + (p_{n-1}(t)-p_j(t))^T(p_n(t)-p_{n-1}(t)) \right ]\\ & & - \frac{2}{n-2}\left [ \sum_{k=1}^{n-2}(p_{n-1}(t)-p_j(t))^T(p_k(t)-p_j(t)) + \delta_{n-1,k}^2(t)\right ] \end{eqnarray*} Using now in the above equation, at $t=t_1$, $ \delta_{n-1,j}(t_1) = R$, (\dref{scaled1-t1}), (\dref{scaled2-t1}) and the inequality for inner products $V_1^T V_2 \leq |V_1||V_2|$, we obtain \begin{eqnarray*} \frac{d}{dt} \delta_{n-1,i}^2(t_1) & \leq & -\frac{2}{n-1}R^2 \\ & < & 0 \end{eqnarray*} again contradicting the assumption that statement \tref{b} of the Theorem does not hold for this link. \end{enumerate} {\hfill $\Box$}{\cmd{PE}} \subsubsection{Some simulation results with incomplete initial interaction graph}\label{IncompleteEx} In this section two examples are shown where the agents initial interaction topology is incomplete. Although we could not find analytic limits on $|u|$ such that the property of never lose neighbors is ensured we ran both cases, and many others, with $|u|< R$ and in both cases the agents converged to a complete graph which was afterwards preserved. \paragraph{Incomplete initial interaction graph - Ex1} \begin{itemize} \item n=6 \item initial number of links = 10 \item $u$ and leaders as shown in Fig. \dref{fig-IncompleteEx1} \end{itemize} \begin{figure} \begin{center} \includegraphics[scale=0.5]{UniScaledDyn-IncompleteEx1.jpg} \caption{Dynamics with Incomplete initial graph - Ex1}\label{fig-IncompleteEx1} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[scale=0.7]{InteractionsGraph-IncompleteEx1.jpg} \caption{Convergence to complete graph - Ex1}\label{fig-GraphsIncompleteEx1} \end{center} \end{figure} \paragraph{Incomplete initial interaction graph - Ex2} \begin{itemize} \item n=8 \item initial number of links = 12 \item $u$ and leaders randomly selected, as shown in Fig. \dref{fig-IncompleteEx2} \end{itemize} \begin{figure} \begin{center} \includegraphics[scale=0.4]{StateDependentTopo-Ex2Incomplete.jpg} \caption{Convergence to complete graph - Ex2}\label{fig-GraphsIncompleteEx2} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[scale=0.4]{Ex2IncompleteDyn.jpg} \caption{Dynamics with initially incomplete graph - Ex2}\label{fig-IncompleteEx2} \end{center} \end{figure} \newpage \msec{Summary and directions for future research}{future} In this report we introduced a model for controlling swarms of identical, simple, oblivious, myopic agents by broadcast velocity control that is received by a random set of agents in the swarm. The agents detecting the broadcast control are the ad-hoc leaders of the swarm, while they detect the exogenous control. All the agents, modeled as single integrators, apply a local linear gathering control, based on the weighted relative position to all neighbors. The weights are the neighbors' influence on the agent. The leaders superimpose the received exogenous control, a desired velocity $u$. We considered two models of neighbors influence, uniform and scaled by the size of agent's neighborhood. We have shown that if the the system is piecewise constant, where in each time interval the system evolves as a time-independent dynamic linear system with a connected visibility graph, then in each such interval, $[t_k, t_{k+1})$, the swarm tends to asymptotically align on a line in the direction of $u(t_k)$, anchored at the zero-input gathering point, $\alpha(t_k)$ and moves with a collective velocity that is a fraction of the desired velocity. We denote this fraction by $\beta(t_k)$. If the visibility graph in the interval is complete, then $\alpha(t_k)$ for both influence models is the same, the average of all agents' positions at the beginning of the interval, and $\displaystyle \beta(t_k)=\frac{n_l(t_k)}{n} $. However, if the visibility graph in the interval is incomplete then $\alpha(t_k)$ and $\beta(t_k)$ are not the same for the two models. Moreover, in the scaled case they are a function of the topology of the graph, the in-degree of its nodes and of the selected leaders. Since we assumed that in each interval the visibility graph is connected we need conditions to ensure that a connected graph remains connected. We showed that if the graph is complete then restrictions on $|u|$, pending the influence model, will ensure that it remains complete. However, when the graph is incomplete, conditions for never losing neighbors are tightly related to the graph topology and therefore not useful in practice. Never losing neighbors might be too stringent a requirement. We note that although conditions, independent of specific topologies, for never losing friends in an incomplete graph were not found, in practice all simulations that we ran showed convergence to complete graphs which were afterwards preserved if $|u|$ was within bounds.\\ In future research, we intend to extend the dynamic model to double integrators, i.e. acceleration controlled agents. Also, we are currently considering the same paradigm of stochastic broadcast control in conjunction with non-linear gathering processes, as for example \dcite{Bellaiche}, and connectedness preserving gathering processes, as for example \dcite{DJ2010}. \newpage
9ad3588cfe0e00cbf343a3bf61d2140e69b9d402
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Measurements on separated subsystems in a joint entangled state may display correlations that cannot be mimicked by local hidden variable (LHV) models. These correlations are termed nonlocal and are detected by violating Bell inequalities \cite{bell1964,review}. In recent years it has become clear that non-locality is interesting not only for fundamental reasons, but also as a resource for device-independent (DI) quantum information tasks \cite{review} such as quantum key distribution \cite{mayersyao,acin07} or random number generation \cite{colbeck pironionature}. Thus, violations of Bell inequalities are not only indicators of non-locality, but can also be used to make qualitative and quantitative statements about operationally relevant quantum properties. Traditionally, the problem of constructing Bell inequalities has been addressed from the point of view of deriving constraints satisfied by LHV models. Following this standard approach, the inequalities are derived using well-known techniques in convex geometry. Indeed, the set of correlations admitting LHV models defines a polytope \cite{review}, i.e., a bounded convex set with a finite number of vertices. These vertices correspond to local deterministic assignments, while the facets are the desired Bell inequalities. Facet (or tight) Bell inequalities provide necessary and sufficient criteria to detect the non-locality of given correlations. Clauser-Horne-Shimony-Holt (CHSH)~\cite{chsh1969} and Collins-Gisin-Linden-Massar-Popescu (CGLMP)~\cite{cglmp} Bell inequalities are examples thereof. Although such facet Bell inequalities are optimal detectors of non-locality, they are not necessarily optimal for inferring specific quantum properties in the DI setting. For instance, in a scenario where two binary measurements are performed on two entangled subsystems, it is well known that the violation of the CHSH inequality \cite{chsh1969} is a necessary and sufficient condition for non-locality. But certain ``non-facet" Bell inequalities are better certificates of randomness than the CHSH one when the two quantum systems are partially entangled \cite{acin2012}. The main aim of this work is to introduce Bell inequalities valid for an arbitrary number of measurements and outcomes whose maximal quantum violation, usually referred to as the \textit{Tsirelson bound} \cite{tsirelson1980}, is attained by maximally entangled states \begin{equation} |\psi^+_d \rangle=(1/\sqrt{d})\sum_{i=0}^{d-1}|ii\rangle. \end{equation} This is a desirable property since these states have particular features such as perfect correlations between outcomes of local measurements in the same bases, and therefore many quantum information protocols rely on them. In the particular case of two measurements CHSH is the simplest example of a Bell inequality with the above property, but others are known \cite{wonminson,lee2007,devicente2015} (see also results for many settings \cite{Ji,Liang,Lim}). Our construction works, however, for arbitrary numbers of measurements and outcomes, and, crucially, the Tsirelson bound of the resulting Bell inequalities can be computed \textit{analytically}. In the case where only two measurements are made on each subsystem, all facet Bell inequalities are known for a small number of outputs and they are of the CGLMP form \cite{cglmp}. However, they are not maximally violated by the maximally entangled states of two qudits (except in the case $d=2$ corresponding to the CHSH inequality) \cite{acindurt,zohrengill,npa2007}. Thus, we should not expect our Bell inequalities to be tight, and indeed they are not. This implies that we cannot use standard tools from convex geometry to construct them. In fact, no quantum property is used for the construction of tight Bell inequalities like the CGLMP one and, in this sense, it is not surprising that their maximal violation does not require maximal entanglement. Our approach is completely different: it starts from quantum theory and exploits the symmetries and perfect correlations of maximally entangled states to derive a Bell inequality (cf. Ref. \cite{Lim} for a similar method). It exploits sum of squares decompositions of Bell operators, which is used to determine their Tsirelson bound. Thus, contrary to any previous derivation of Bell inequalities, quantum theory becomes a key ingredient of our method. Our results provide new insight into the structure of the boundary of the set of quantum correlations (see discussion in Appendix \ref{appendixstructure}). In addition, our Bell inequalities have the potential to be used in DI quantum information protocols such as random number generation, quantum key distribution or to self-testing \cite{mckague2012} (see more detailed discussion in Section \ref{secappli}). \section{Preliminaries} We consider a Bell scenario with two distant parties $A$ and $B$ performing one of $m$ measurements $A_x$ and $B_y$ with $d$ outcomes on their share of some physical system. We label the measurements and outcomes as $x,y \in \{1,\ldots, m\}$ and $a,b \in \{0,\ldots, d-1\}$. The correlations obtained in this experiment are described by $(md)^2$ joint probabilities $P(A_x=a,B_y=b)$ that $A$ and $B$ obtain $a$ and $b$ upon performing the $x$th and $y$th measurement, respectively. These probabilities are ordered into a vector $\vec{p}:=\{P(A_x=a,B_y=b)\}_{a,b,x,y} \in \mathbbm{R}^{(md)^2}.$ Importantly, the set of allowed vectors $\vec{p}$ varies depending on the physical principle they obey. If the measurements define spacelike separated events, the observed correlations should obey the \textit{no-signalling principle}, which prevents any faster-than-light communication among the parties. These correlations form a convex polytope denoted $\mathcal{N}$. Contained in this set is the set of quantum correlations $\mathcal{Q}$ which is formed by those $\vec{p}$ whose components can be written as $P(A_x=a,B_y=b)=\langle \psi|P_a^{(x)}\otimes P_b^{(y)}|\psi\rangle$, where $|\psi\rangle$ is some state in a product Hilbert space $H_A\otimes H_B$ of unconstrained dimension, and $\{P_a^{(x)}\}$ and $\{P_b^{(y)}\}$ are projection operators defining, respectively, Alice's and Bob's measurements. Finally, the set of correlations admitting LHV models, denoted $\mathcal{L}$, contains those $\vec{p}$ that can be written as a convex sum of product deterministic correlations $P(A_x=a,B_y=b)=P(A_x=a)P(B_y=b)$ with $P(A_x=a),P(B_y=b)=0,1$ for all $x,y$ \cite{fine1982}. Bell was the first to prove that not all quantum correlations admit an LHV model \cite{bell1964}. To this end, he used the concept of a Bell inequality $I\leq C_b$, where $I$ is the so-called Bell expression that is a linear combination of the $(md)^2$ joint probabilities of the form \begin{equation}\label{BE} I:=\sum_{abxy} I_{abxy} P(A_x=a,B_y=b), \end{equation} and $C_b=\max_{\vec{p} \in \mathcal{L}}I$ is its classical bound. The quantum or Tsirelson bound of $I$ is the maximum value $Q_b=\max_{\vec{p} \in \mathcal{Q}}I$ that it can achieve for quantum correlations. A Bell expression $I$ gives rise to a proper Bell inequality---one that is violated by quantum theory---if $C_b < Q_b$. If $\vec{p}$ violates a Bell inequality, the correlations described by $\vec{p}$ are termed nonlocal. Finally, one defines $NS_b=\max_{\vec{p} \in \mathcal{N}}I$ as the maximum value of $I$ over no-signalling correlations. For most of the known Bell inequalities $NS_b>Q_b>C_b$ \cite{bell1964,PR,ravi}. Let us stress that although $\mathcal{Q}$ is convex, it is not a polytope. More importantly, the boundary of $\mathcal{Q}$ remains unknown despite several attempts to characterize it analytically \cite{pawlowski09,wunderlich,lorthogonality,navascues2015} (see, nevertheless, \cite{Masanes}). This clearly makes the derivation of Tsirelson bounds a hard task. Given a Bell inequality, there is no procedure that guarantees finding its quantum bound, and it was achieved analytically only in a handful of cases. There is, however, a practical approximation scheme based on semidefinite programming, which consists in a hierarchy of sets $\mathcal{Q}_{1} \supseteq \mathcal{Q}_{2} \supseteq \dots \supseteq \mathcal{Q}_{k} \supseteq \dots$ converging to $\mathcal{Q}$ as $k \to \infty$, and allows one to bound $Q_b$ from above \cite{npa2007} (see also \cite{devicente2015}). Although for small Bell scenarios this method yields good numerical bounds (often tight), it becomes computationally expensive for scenarios involving a large number of measurements or outcomes. \section{Class of Bell expressions}\label{secclass} Our aim now is to introduce a family of Bell expressions, whose maximal quantum value is attained by the \textit{two-qudit} maximally entangled state $|\psi^+_d \rangle$. To derive them, we start from the premise that their maximal quantum values are obtained when Alice and Bob perform the optimal CGLMP measurements introduced in \cite{kaszlikowski,cglmp,barrett2006} (cf. Appendix \ref{appendixmeasurements}). This choice stems from the fact that these measurements generalize the CHSH measurements ($d=2$) to arbitrary dimensions and they lead to non-local correlations that are most robust to noise~\cite{kaszlikowski} or for $m=2$ give a stronger statistical test~\cite{gill}. The probabilities $P(A_x=a,B_y=b)$ obtained when using the optimal CGLMP measurements on $|\psi^+_d \rangle$ have several symmetries. For instance, they only depend on the difference $a-b=k\mod d$. If we impose that our Bell expressions respect this symmetry, the probabilities $P(A_{x} = j + k \text{ mod }d,\text{ } B_{y} = j)$ should be treated equally for all $j$, i.e., the Bell expressions should be linear combinations of $P(A_{x} = B_{y} + k) := \sum_{j = 0}^{d-1} P(A_{x} = j + k \text{ mod }d,\text{ } B_{y} = j)$. Taking into account all symmetries, a generic form for our Bell expressions is \begin{equation}\label{Bellproba} I_{d,m}:=\sum_{k=0}^{\left \lfloor d/2\right \rfloor-1}\left( \alpha_k \mathbbm{P}_k -\beta_k \mathbbm{Q}_k \right ), \end{equation} where $\mathbbm{P}_k := \sum_{i=1}^{m}[P(A_i=B_i+k)+P(B_i=A_{i+1}+k)]$, $\mathbbm{Q}_k := \sum_{i=1}^{m}[P(A_i=B_i-k-1)+P(B_i=A_{i+1}-k-1)]$ with $A_{m+1}:=A_1+1$. The parameters $\alpha_k$ and $\beta_k$ are our degrees of freedom. Taking, e.g., $\alpha_k = \beta_k = 1 - 2k/(d - 1)$ for $m=2$, one recovers the CGLMP Bell inequalities. To exploit the symmetries inherent in Bell inequalities, we often write them in terms of correlators instead of probabilities. As we consider an arbitrary number of outcomes, we appeal to the notion of generalized correlators (see, e.g., Ref. \cite{Liang} and Ref. \cite{bancal12} for other options). These are complex numbers that are defined through the two-dimensional Fourier transform of the probabilities $P(A_x=a,B_y=b)$: \begin{equation} \langle A_x^kB_y^l\rangle=\sum_{a,b=0}^{d-1}\omega^{ak+bl}P(A_x=a,B_y=b), \label{correlators} \end{equation} where $\omega = \text{exp}(2\pi i /d)$, $k,l\in \{0,\ldots,d-1\}$, and $\{A_x^k\}_k$ and $\{B_y^l\}_l$ can be thought of as measurements with outcomes labelled by roots of unity $\omega^i$ $(i=0,\ldots,d-1)$. For quantum correlations $\vec{p}$, the correlators $\langle A_x^kB_y^l\rangle$ are average values of the tensor product of the operators $A_x^{k}=\sum_{a=0}^{d-1}\omega^{ak}P_a^{(x)}\quad \mathrm{and} \quad B_y^{l}=\sum_{b=0}^{d-1}\omega^{bl}P_b^{(y)}$ in the state $\ket{\psi}$. Note that they are unitary, their eigenvalues are the roots of unity, and they satisfy $(A_x^{k})^{\dagger}=A_x^{d-k}$ and $(B_y^{l})^{\dagger} = B_{y}^{d-l}$ for any $k,l$. Now, exploiting (\ref{correlators}), expression (\ref{Bellproba}) can be rewritten as \begin{equation}\label{Bellcorr} \widetilde{I}_{d,m}= \sum_{i=1}^m\sum_{l=1}^{d-1}\langle A_i^l\bar{B}_i^l \rangle, \end{equation} where, for clarity, the change of variables $\bar{B}_i^l = a_l B_i^{d-l}+a_l^{*}B_{i-1}^{d-l}$ with $ a_l=\sum_{k=0}^{\lfloor d/2\rfloor-1}(\alpha_k\omega^{-kl}-\beta_k\omega^{(k+1)l})$ was introduced on Bob's side. Due to the convention $A_{m+1}=A_1+1$, the term $\bar{B}_1^l$ is defined as $\bar{B}_1^l=a_lB_{1}^{d-l}+a_l^* \omega^lB_{m}^{d-l}$. For simplicity, in (\ref{Bellcorr}) we ignored the irrelevant scalar term corresponding to $l=0$ and rescaled the expression. Below we denote the classical, quantum and no-signaling bound of $\widetilde{I}_{d,m}$ by $\widetilde{C}_b$, $\widetilde{Q}_b$ and $\widetilde{NS}_b$, respectively. Our aim now is to fix the free parameters $\alpha_k$ and $\beta_k$ according to the quantum property we need: maximal violation by the maximally entangled state $| \psi^+_d \rangle$. At this point, it is instructive to look at the specific example of the CHSH Bell expression ($m=2$, $d=2$). In the notation (\ref{Bellcorr}) the CHSH Bell expression $\langle A_1B_1\rangle + \langle A_1 B_2\rangle + \langle A_2 B_1\rangle- \langle A_2 B_2\rangle$ reads $\widetilde{I}_{2,2}=\langle A_1 \bar{B}_1\rangle +\langle A_2 \bar{B}_2\rangle,$ where $\bar{B}_1=(B_1 + B_2)/\sqrt{2}$, $\bar{B}_2=(B_1 - B_2)/\sqrt{2}$. Then, for the optimal measurements leading to the Tsirelson bound of $\widetilde{I}_{2,2}$, we have $\bar{B}_1=A_1^*$ and $\bar{B}_2=A_2^*$. This reflects the property that for the maximally entangled state \begin{equation}\label{Raimat} M\otimes N\ket{\psi_d^+}=\mathbbm{1}\otimes NM^{T}\ket{\psi_d^+},\quad \forall M,N. \end{equation} This condition implies that a measurement by Alice is perfectly correlated with its complex conjugate by Bob. Our intuition to derive Bell inequalities detecting maximal entanglement is to impose this property for any $m$ and $d$: we choose the parameters $\alpha_k$ and $\beta_k$ such that \begin{equation}\label{conditions} \bar{B}_{i}^{l} = (A_{i}^{l})^{*} \end{equation} hold for $l = 1,\ldots, d -1$ and $i = 1, \ldots, m$ with the initial operators $\{P_a^{(x)}\}$ and $\{P_b^{(y)}\}$ being the optimal CGLMP operators. Conditions (\ref{conditions}) give rise to a set of linear equations for $\alpha_k$ and $\beta_k$ which yields (see Appendix \ref{appendixcoefficients} for details) \begin{equation}\label{alpha} \alpha_k= \frac{1}{2d}\tan\left(\frac{\pi}{2m}\right) \left[g(k) - g\left(\left\lfloor \frac{d}{2} \right\rfloor\right)\right], \end{equation} \begin{equation}\label{beta} \beta_k= \frac{1}{2d}\tan\left(\frac{\pi}{2m}\right) \left[g\left(k + 1 - \frac{1}{m}\right) + g\left(\left\lfloor \frac{d}{2} \right\rfloor\right)\right] \end{equation} with $g(x):=\cot(\pi(x+ 1/2m)/d)$. To sum up, our class of Bell expressions is given by $I_{d,m}$ (\ref{Bellproba}) or equivalently by $\widetilde{I}_{d,m}$ (\ref{Bellcorr}), with coefficients (\ref{alpha}) and (\ref{beta}). We arrived at it by writing the most general Bell expression satisfying the symmetry of CGLMP correlations, re-writing these Bell expressions in the simple form (\ref{Bellcorr}) through a change of variable on Bob's side, and then imposing the conditions (\ref{conditions}) that take into account the symmetries of the maximally entangled state, as CHSH does for two binary measurements. \section{Properties of the novel Bell expressions}\label{secproperties} We now analyze the main properties of our Bell expressions: we compute all the relevant bounds $\widetilde{C}_b$, $\widetilde{Q}_b$, $\widetilde{NS}_b$, and show that $\widetilde{C}_b<\widetilde{Q}_b<\widetilde{NS}_b$ for any $d$ and $m$. For clarity we only include sketches of proofs (see Appendices \ref{appendixclassical}, \ref{appendixtsirelson} and \ref{appendixns} for details). Let us begin with the classical bound. \begin{thm}\label{theoclass} The classical bound of $\widetilde{I}_{d,m}$ is given by $\widetilde{C}_{b} = (1/2) \tan \left(\pi/2m\right) \left\{ (2m - 1) g(0) - g(1 - 1/m)\right\} - m.$ \end{thm} \begin{proof} We start with the expression $I_{d,m}$. Since we can restrict the problem to local deterministic strategies, finding $\widetilde{C}_b$ becomes a question of distributing $0$s and $1$s over all the terms $P(A_x = B_y + z)$. It turns out that the optimal strategy is to set $2m - 1$ of the terms multiplied by $\alpha_0$ and a single term multiplied by $\beta_0$ to one, and the remaining terms to zero. \end{proof} Importantly, the resulting Bell inequality $\widetilde{I}_{d,m}\leq \widetilde{C}_b$ is violated by quantum theory; one can reach the value $\widetilde{I}_{d,m}=m(d-1)$ by applying the CGLMP measurements on $\ket{\psi^+_d}$. This is seen by using Eq. (\ref{conditions}), the unitarity of $A_{i}^{k}$, and the symmetries of the maximally entangled states (\ref{Raimat}). Then, all the correlators in (\ref{Bellcorr}) equal one, yielding the quantum violation of $m(d-1)$. This violation is optimal and defines the tight Tsirelson bound of $\widetilde{I}_{d,m}$. \begin{thm}\label{theoquantum} The Tsirelson bound of $\widetilde{I}_{d,m}$ is given by $\widetilde{Q}_{b} = m(d-1)$. \end{thm} \begin{proof} We construct a sum-of-squares (SOS) decomposition of the shifted Bell operator $\widetilde{\mathcal{B}} := \widetilde{Q}_{b}\mathbbm{1} - \mathcal{B}$, where $\mathbbm{1}$ is the identity operator and $\mathcal{B}$ the Bell operator corresponding to expression (\ref{Bellcorr}) (see, e.g., \cite{bamps2015,doherty}). For any positive semidefinite operator $\mathcal{P}$, an SOS decomposition is a collection of operators ${P_{\lambda}}$ such that $\mathcal{P} = \sum_{\lambda} P_{\lambda}^{\dagger} P_{\lambda}.$ If $\widetilde{\mathcal{B}}$ admits the latter form it must be positive semidefinite, implying that $\widetilde{Q}_{b}$ upper bounds our Bell expression, i.e., $\langle \psi | \mathcal{B} | \psi \rangle \leq \widetilde{Q}_{b}$ for any $\ket{\psi}$. To show that $\widetilde{Q}_{b} = m(d-1)$ is indeed the Tsirelson bound of $\widetilde{I}_{d,m}$, we prove that $\widetilde{Q}_{b}\mathbbm{1}-\mathcal{B}$ decomposes as \begin{equation}\label{sos} \widetilde{Q}_{b}\mathbbm{1}-\mathcal{B} = \frac{1}{2}\sum_{i=1}^{m}\sum_{k=1}^{d-1} P_{ik}^{\dagger} P_{ik} + \frac{1}{2}\sum_{i=1}^{m - 2}\sum_{k=1}^{d-1} T_{ik}^{\dagger} T_{ik}, \end{equation} where $P_{ik} = \mathbbm{1} \otimes \bar{B}_i^k - (A_i^{k})^{\dagger} \otimes \mathbbm{1}$, and $T_{ik} = ( \mu_{i,k} B_{2}^{d - k} + \nu_{i,k} B_{i + 2}^{d - k} + \tau_{i,k} B_{i + 3}^{d - k})$ with $\mu_{i,k}$, $\nu_{i,k}$, $\tau_{i,k} \in \mathbb{R}$. The Bell operator reads $\mathcal{B} = \sum_{i=1}^m\sum_{l=1}^{d-1} A_i^k\otimes\bar{B}_i^k$, and the decomposition is independent of the choice of $A_i^k$ and $B_i^k$. The exact values of the coefficients along with details on the SOS decomposition can be found in Appendix \ref{appendixtsirelson}. \end{proof} A few remarks are in order. First, it is not difficult to see that $\widetilde{Q}_b>\widetilde{C}_b$ for any $m,d\geq 2$, meaning that all our Bell inequalities are nontrivial (cf. Appendix \ref{appendixscalings}). Second, let us elaborate on how the SOS works in the case of two measurements, $m=2$, which justifies the choice of conditions (\ref{conditions}). For $m=2$, the second part of the SOS decomposition \eqref{sos} vanishes. For the optimal CGLMP measurements both sides of (\ref{sos}) must yield zero when applied to $\ket{\psi^+_d}$, which stems from conditions (\ref{Raimat}) and (\ref{conditions}). This allows one to grasp the intuition behind conditions (\ref{conditions}), i.e., they allow one to construct in a quite direct way an SOS decomposition (\ref{sos}), in which all operators $P_{ik}$ are polynomials of the measurement operators $A_{i}^{k}$ and $B_{i}^{k}$ of order one, significantly facilitating the computation of the Tsirelson bound. For the CHSH Bell inequality, one observes the same effect, as these same properties of the optimal state and measurements allow the Bell operator $\mathcal{B}_{\text{CHSH}} = A_1 \otimes B_1 + A_1 \otimes B_2 + A_2 \otimes B_1 - A_2 \otimes B_2$ to have the decomposition: $2\sqrt{2}\mathbbm{1} - \mathcal{B}_{\text{CHSH}} = ( P_{1}^{\dagger}P_1 + P_{2}^{\dagger}P_2)/\sqrt{2},$ with $P_1 = (1/\sqrt{2})\mathbbm{1}\otimes (B_1 +B_2) - A_1 \otimes \mathbbm{1}$, and $P_2 = (1/\sqrt{2})\mathbbm{1} \otimes(B_1 - B_2) - A_2 \otimes \mathbbm{1} $. Thus, our construction generalizes this quantum aspect of the CHSH Bell operator. For larger number of measurements, $m>2$, the first part of the SOS decomposition is not enough and one has to add ``by hand'' the extra term in which all $ T_{ik}$'s are also of order one in $B_i^k$. Note that for two measurements, our Bell expressions coincide with those introduced in \cite{wonminson} and then rederived in \cite{devicente2015} using a different approach. Moreover, the Tsirelson bounds of these Bell inequalities was computed in Refs. \cite{lee2007,devicente2015} exploiting other techniques, and it was proven in \cite{lee2007} that they are not tight. On the other hand, for $d=2$ and any $m$, our class recovers the well-known chained Bell inequalities \cite{chained}. We finally notice that the alternative generalization of the CHSH Bell inequality to three measurements and outcomes given in \cite{BuhrmanMassar} was also found to be maximally violated by $\ket{\psi^+_3}$ \cite{Liang}. Let us eventually compute the no-signalling bound of our Bell expressions. \begin{thm}\label{theons} The no-signalling bound of $\widetilde{I}_{d,m}$ is given by $\widetilde{NS}_{b} = m \tan \left(\pi/2m\right) g(0) - m$. \end{thm} \begin{proof} We provide no-signalling correlations $\vec{p}$ and show that they attain the algebraic bound of $I_{d,m}$. They correspond to having all the probabilities which are multiplied by $\alpha_0$ in $I_{d,m}$ equal to one, and all the others equal to zero (see Appendix \ref{appendixns}). \end{proof} Again, it is not difficult to see that $\widetilde{NS}_b>\widetilde{Q}_b$ for any $m,d\geq 2$ (see Appendix \ref{appendixscalings} for the proof and scalings of $\widetilde{C}_b$, $\widetilde{Q}_b$ and $\widetilde{NS}_b$ with $m$ and $d$). \section{Applications to device-independent protocols}\label{secappli} A natural application for our Bell inequalities is self-testing---a DI protocol in which a state and measurements performed on it are certified up to local isometries, based on the nonlocal correlations they produce. To perform self-testing, the correlations $\vec{p}$ maximally violating the given Bell inequality must be unique, i.e., attained, up to local isometries, by certain state and measurements. This is generally hard to prove. There exists, however, a numerical method for self-testing \cite{swaptest1}. We applied it to the simplest case $m=2$ and $d = 3$, and the results are plotted in Figure \ref{swap3}. It shows that one can self-test the maximally entangled state of two qutrits $\ket{\psi^+_3} =(\ket{00} + \ket{11} + \ket{22})/\sqrt{3}$ with our inequalities. \begin{figure} \includegraphics[width=0.4\textwidth]{swap_v9.pdf} \caption{Minimum fidelity of the state in the black box to the maximally entangled state of two qutrits, as a function of the violation of $\widetilde{I}_{3,2}$. At the maximal violation $4$, the fidelity is equal to $1$, meaning that the quantum state used in the Bell experiment must be maximally entangled. The numerical method that we used does not yield a positive lower bound on the fidelity below $ \widetilde{I}_{3,2} \approx 3.79$ (for comparison, the classical bound is $\widetilde{I}_{3,2} \leq (1 + 3\sqrt{3})/2 \approx 3.01$).\label{swap3}} \end{figure} An open question is whether one can generalize this result to any dimension. Our inequalities could then be applied in DI random number generation protocols \cite{colbeck pironionature,pironiomassar}. Indeed, if $\vec{p}$ maximally violating $\widetilde{I}_{d,m}$ is unique, one can apply the method of \cite{dhara2013} and use the symmetries of the Bell expressions to guarantee a dit of perfect randomness. This, by increasing the dimension $d$, would result in unbounded randomness expansion. Our inequalities could also find applications in DI quantum key distribution. An advantage that our inequalities have over CGLMP in that scenario \cite{marcusmarcin} is that, as said before, the maximal violation is obtained for the maximally entangled state. This state can produce perfect correlations between the users, which reduces the error-correcting phase of the protocol and can lead to better key generation rates. We study this question in Appendix \ref{diqkd}. Numerically, we find that for $m=2$ and $d=3$, our inequalities lead to higher key rates than CGLMP for levels of white noise up to $4.2$ percent. While this advantage is not very large, we believe it grows with the dimension of the systems, at least in the noiseless case \cite{zohrengill}. Moreover, it is known that maximally entangled states are much simpler to prepare experimentally than fine-tuned partially entangled states. It would be interesting to confirm these conjectures in a future work focused on DIQKD. \section{Conclusions} In this work, we introduced a new technique allowing to construct Bell inequalities with arbitrary numbers of measurements and outcomes that are maximally violated by the maximally entangled states. It exploits the SOS decompositions of Bell operators and, crucially, allows one to compute analytically their Tsirelson bounds. Our results are general as, unlike previous works, we do not consider a particular Bell scenario, but allow for arbitrary number of measurements $m$ and outcomes $d$. Our inequalities can be seen as the ``quantum'' or the DI-oriented generalization of CHSH Bell inequality, in the same spirit as the CGLMP inequality generalizes the CHSH one classically. \begin{acknowledgments} We wish to thank Y.-C. Liang, M. Navascu\'{e}s, T. V\'{e}rtesi and J. Kaniewski for fruitful discussions, and especially J.-D. Bancal for sharing with us his code. This work was supported ERC CoG QITBOX and AdG OSYRIS, the AXA Chair in Quantum Information Science, Spanish MINECO (FOQUS FIS2013-46768-P, SEV-2015-0522, QIBEQI FIS2016-80773-P and FISICATEAMO FIS2016-79508-P), Fundaci\'o Privada Cellex, the Generalitat de Catalunya (SGR 874, SGR 875 and the CERCA programme), the EU projects QALGO and SIQS, and the John Templeton Foundation. This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreements No 705109 and No 748549. We acknowledge financial support from the Fondation Wiener-Anspach and the Interuniversity Attraction Poles program of the Belgian Science Policy Office under the grant IAP P7-35 photonics@be. J. T. acknowledges the CELLEX-ICFO-MPQ programme. S. P. is a Research Associate of the Fonds de la Recherche Scientifique F.R.S.-FNRS (Belgium). \end{acknowledgments}
693a2d99c73b0b020794763c08ad7c514e6530c6
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction \label{sec1}} The on going search for new physics (NP) is mostly inspired by the shortcomings of the SM in addressing some of the fundamental questions in modern particle physics, such as the hierarchy problem, the flavor patterns in the fermionic sector and dark matter. Some of these unresolved issues may be closely related and may have TeV-scale origins, thus inspiring the search for TeV-scale NP, both theoretically and experimentally. Indeed, two seemingly unrelated interesting measurements of both the ATLAS \cite{ATLAS1,ATLAS2} and the CMS \cite{CMS1,CMS2} collaborations at CERN, have been recently reported: \begin{enumerate} \item A possible $(2-4) \sigma$ (local) excess in the diphoton invariant mass distribution around 750 GeV, corresponding to a signal cross-section roughly in the range $\sigma(pp \to \gamma \gamma) \sim 3-13$ fb $(1\sigma)$, see e.g., \cite{1512.05777,1601.04751,1605.09401}. The interpretation of this excess signal has a slight preference to a spin 0 resonance, produced via gluon-fusion and having a total width ranging from sub-GeV to 45 GeV, with a more significant signal obtained in the ATLAS analysis for a scalar with a total width $\Gamma \sim 45~{\rm GeV}$ \cite{ATLAS1}. \item A possible $(1-2.5) \sigma$ excess in the measurement of the LFV decay $h \to \tau \mu$ of the 125 GeV light Higgs. In particular, the CMS collaboration finds $BR(h \to \tau \mu) = 0.84\%^{+0.39\%}_{-0.37\%}$ \cite{CMS2}, while the ATLAS collaboration finds $BR(h \to \tau \mu) = (0.53 \pm 0.51)\%$ \cite{ATLAS2}. \end{enumerate} Whether or not these two measurements are confirmed, it emphasizes the importance of the current efforts in the search for NP, since it provides an interesting manifestation/example of the exciting possibility that the building blocks of new TeV-scale physics may have rather non-conventional properties, potentially with important repercussions for both flavor and the hierarchy problems. For example, the new heavy scalar particle, $S$, responsible for the 750 GeV $\gamma \gamma$ excess, should have a rather narrow width and suppressed decay rates into ``conventional" channels such as $S \to WW,~ZZ,~t \bar t$, for which no excess signals has been observed within the currently available sensitivity of the corresponding LHC searches. In addition, such a heavy scalar $S$ is most likely related to the light 125 GeV scalar state and, therefore, might also be involved in flavor changing (FC) transitions in the fermionic sector. Such properties of the would be new 750 GeV resonating particle are, therefore, very challenging to accommodate in models beyond the SM, in particular, in supersymmetric models or in models that involve extra space-time dimensions, which seem to have a more fundamental origin and, therefore, likely linked to physics at higher energy scales. Nonetheless, we will show in this paper that a certain class of low-energy effective 2HDM frameworks with a 4th generation of heavy chiral fermions may be interesting candidates for such ``exotic" TeV-scale NP. In particular, since no evidence for such fundamentally structured theories has yet been seen, a frequently adopted phenomenological approach in studies of NP, is to construct TeV-scale models which require a UV completion and may, thus, be viewed as low energy effective frameworks for the underlying dynamics. Such models are useful as a guide for the exploration and model building of more fundamental theories and they often include new heavy fermionic and scalar states with sub-TeV masses. One of the simplest variants of an effective low-energy NP candidate, which dates back to the 1980's \cite{SM4proc}, is the SM with an additional 4th generation of fermions; the so called SM4 (for useful reviews see e.g., \cite{sher0}). Indeed, since three generations of chiral fermions have been observed in nature, it is natural to ask why not four generations of chiral fermions? It is quite interesting that this simple extension of the SM may address some of the theoretical challenges in particle physics, such as: electroweak symmetry breaking (EWSB) and the hierarchy problem \cite{DEWSB}, the CP-violation and the strength of the first-order phase transition needed to explain the origin of matter - anti matter asymmetry in the universe \cite{baryo-ref,fok}, and flavor physics \cite{flavor}. As is well known, the SM4 (i.e., with four generations of fermions and one Higgs doublet) is now excluded, since it cannot accommodate the measured SM-like properties of the 125 GeV scalar, see e.g., \cite{SM4-Higgs-bound,Lenz}, primarily due to an ${\cal O}(10)$ enhancement in the gluon-fusion light Higgs production mechanism from diagrams with $t^\prime$ and $b^\prime$ in the loops \cite{Kribs_EWPT}; see, however, W.-S. Hou in \cite{1606.03732}. This fact, along with the rather stringent direct limits on the masses of such heavy quarks (to be discussed later), has led to a common belief that generic extensions to the SM with heavy chiral 4th generation fermions $t^\prime,b^\prime,\nu^\prime,\tau^\prime$ are excluded. However, as was suggested by us a few years ago \cite{ourpaper1} and will be demonstrated again here, this is not the case when the heavy 4th generation chiral sector is embedded in frameworks with an extended Higgs sector (see also \cite{1312.4249}). Indeed, an extended Higgs sector in the context of 4th generation heavy fermions may come in handy for further addressing flavor problems \cite{ourpaper1} and the strength of the EW phase transition required for baryogenesis \cite{fok,1401.1827}. In particular, we will consider in this paper a version of a 2HDM introduced by us in \cite{ourpaper1} - the so called 4G2HDM of type I, where a chiral 4th generation doublet of heavy fermions (quark and lepton) is added and is coupled {\it only} to one of the scalar doublets (the ``heavy" doublet), while the SM 1st-3rd generations fermions are coupled only to the other doublet (the ``light" doublet). We will show in this paper that this 4G2HDM is a well motivated and valid low-energy model, which is compatible with the 125 GeV signals (see also \cite{our125}), with PEWD and with the existing direct bounds on the heavy fermions, and at the same time can also accommodate the recent indications for a new 750 GeV scalar resonance in the $\gamma \gamma$ channel. As was shown in \cite{ourpaper1}, the price to pay when adding another heavy SM-like chiral fermion doublet is that such constructions posses a nearby threshold/cutoff at the several TeV scale, which is manifest (as Landau poles) in the evolution of the Yukawa and Higgs potential couplings \cite{ourpaper1,ourhybrid}. Indeed, the large Yukawa couplings of the heavy chiral fermions can be thought of as a reflection of an underlying TeV-scale strong dynamics, so that the 4G2HDM framework should be viewed as a low energy (i.e., sub-TeV) effective model of an underlying strongly interacting sector. In particular, if the new heavy chiral fermions are viewed as the agents of EWSB (and are, therefore, linked to strong dynamics at the nearby TeV-scale, see e.g., \cite{DEWSB-heavyF,luty}), then more Higgs particles, which may be composites of these 4th generation fermions, are expected at the sub-TeV regime.$^{[1]}$\footnotetext[1]{Early attempts in this direction investigated the possibility of using the top-quark as the agent of dynamical EWSB via top-condensation \cite{top-con}. These models, however, fail to reproduce the observed value of the top-quark mass. Moreover, as opposed to the case of condensates of the heavy 4th generation fermions, where the typical cutoff for the new strong interactions is of ${\cal O}(1)$ TeV, the top-condensate models require a corresponding cutoff many orders of magnitudes larger than $m_t$, i.e., of ${\cal O}(10^{17})$ GeV, thus resulting in a severely fine-tuned picture of dynamical EWSB.} In such scenarios the resulting low-energy effective theory may contain more than a single composite Higgs field \cite{ourpaper1,luty,DEWSB-multiH} and may thus resemble a two (or more) Higgs doublet framework (for other related studies of the phenomenology of multi-Higgs 4th generation models see e.g., \cite{4G2HDM-others}). The purpose of this work is to revisit the 4G2HDM of \cite{ourpaper1}, studying its compatibility with the updated measurements of the 125 GeV light Higgs signals and with PEWD. We will also confront our model with the 750 GeV $\gamma \gamma$ excess and study its compatibility with a sub-percent branching ratio of the light Higgs in the FC decay channel $h \to \tau \mu$. Indeed, many interesting and exotic constructions beyond the SM have been suggested as possible explanations of the 750 GeV $\gamma \gamma$ excess (too many to be cited here); in most cases involving new degrees of freedom beyond just the 750 GeV resonating particle. In particular, the relevance of 2HDM frameworks to the 750 GeV $\gamma \gamma$ excess has been intensively studied in the past several months, where it was shown that the simplest 2HDM extension to the SM, in which no additional heavy degrees of freedom are added (i.e., beyond the extended scalar sector), cannot accommodate the necessary enhancement in $\sigma(pp \to H(750) \to \gamma \gamma)$, see e.g., \cite{gilad}. Consequently, extended 2HDM models with TEV scale vector-like (VL) fermions have been suggested for addressing the 750 GeV resonance signal \cite{2HDMVLQ}. The upshot of these studies is that, the needed enhancement in the 1-loop production and decay channels $gg \to H(750)$ and $H(750) \to \gamma \gamma$, requires several copies of VL fermions and/or VL fermions with charges appreciably larger than those of the SM fermions, unless their Yukawa couplings are much larger than one. The 4G2HDM considered in this work is, therefore, conceptually simpler, relying on new heavy fermionic degrees of freedom with properties similar to the SM fermions in a model that already exists in the literature. The paper is organized as follows: in section 2 we describe the type I 4G2HDM and we layout the physical parameters that are used in the numerical analysis. In section 3 we show our results and in section 4 we discuss their phenomenological consequences. In section 5 we discuss our results and summarize. \section{The 4G2HDM: a 2HDM with 4th generation fermions \label{sec2}} Motivated by the idea that TeV-scale scalar degrees of freedom may emerge as composites associated with heavy fermions, we assume that the low-energy (sub-TeV) effective framework is parameterized by a 2HDM with a chiral SM-like 4th generation of heavy fermions. Specifically, the model is constructed following \cite{ourpaper1}, such that one of the Higgs fields ($\phi_h$ - the ``heavier" field) couples only to the new heavy 4th generation fermionic fields, while the second Higgs field ($\phi_\ell$ - the ``lighter" field) is responsible for the mass generation of all other (lighter) fermions (i.e., the 1st-3rd generation SM fermions). In this model, named in \cite{ourpaper1} the 4G2HDM of type I (here we will refer to it simply as the 4G2HDM), the Yukawa interaction Lagrangian can be realized in terms of a $Z_2$-symmetry under which the fields transform as follows: \begin{eqnarray} \Phi_{\ell}\to-\Phi_{\ell},~ \Phi_{h}\to+\Phi_{h},~ F_{L}\to+F_{L}~, f_{R}\to-f_{R}\;(f={\rm SM~fermions}),~f^\prime_{R}\to +f^\prime_{R}\;(f^\prime={\rm 4th~gen.~fermions})~, \label{eq:z2} \end{eqnarray} where $F_L$ and $f_{R},f_R^\prime$ are the SU(2) fermion (quark or lepton) doublets and singlets, respectively, and $\Phi_{\ell,h}$ are the two Higgs doublets $\Phi_i =\left( \phi^{+}_i,\frac{v_i+\phi^{0}_i}{\sqrt{2}} \right)$, $i=\ell,h$. The Yukawa potential that respects the above $Z_2$-symmetry is: \begin{eqnarray} \mathcal{L}_{Y}= -\bar{F}_{L} \left( \Phi_{\ell} Y_d^f \cdot \left( I-{\cal I} \right) + \Phi_{h}Y_d^f \cdot {\cal I} \right) f_{d,R} -\bar{F}_{L} \left( \tilde\Phi_{\ell} Y_u^f \cdot \left( I - {\cal I} \right) + \Phi_{h} Y_u^f \cdot {\cal I} \right) f_{u,R} + h.c.\mbox{ ,} \label{eq:LY4G} \end{eqnarray} where $f_{u,R}$ and $f_{d,R}$ are the up and down-type SU(2) fermion singlets (quark or lepton of all four generations), $I$ is the identity matrix and ${\cal I}$ is the diagonal $4\times4$ matrix ${\cal I} \equiv {\rm diag}\left(0,0,0,1\right)$. The scalar sector contains five massive states: a charged scalar $H^+$, a CP-odd state $A$ and two CP-even scalars $h,H$, so that $h$ is the lighter one, corresponding to the observed 125 GeV Higgs boson. These physical states are related to the components of the two SU(2) scalar doublets via: \begin{eqnarray} H= s_\alpha {\rm Re} \left( \phi_h^0 \right) + c_\alpha {\rm Re}\left( \phi_\ell^0 \right) ~&,&~ A= s_\beta {\rm Im} \left( \phi_\ell^0 \right) - c_\beta {\rm Im}\left( \phi_h^0 \right) ~, \nonumber \\ h= c_\alpha {\rm Re}\left(\phi_h^0\right) - s_\alpha {\rm Re}\left(\phi_\ell^0\right) ~&,&~ H^+= s_\beta \phi_\ell^+ - c_\beta \phi_h^+ ~, \label{Higgsangles} \end{eqnarray} where $s_\alpha(c_\alpha)=\sin\alpha(\cos\alpha)$, $\alpha$ being the Higgs mixing angle in the CP-even sector and $s_\beta(c_\beta)=\sin\beta(\cos\beta)$, where $\tan\beta \equiv v_h/v_\ell$ is the ratio between the VEV's of the heavy and light Higgs fields. The Yukawa Higgs-quark-quark interactions in the 4G2HDM are (similar terms can be written for the leptons) \cite{ourpaper1}: \begin{eqnarray} {\cal L}(h q_i q_j) &=& \frac{g}{m_W \sin 2\beta} \bar q_i \left\{ m_{q_i} s_\alpha s_\beta \delta_{ij} - \cos(\beta -\alpha) \cdot \left[ m_{q_i} \Sigma_{ij}^q R + m_{q_j} \Sigma_{ji}^{q \star} L \right] \right\} q_j h \label{Sff1}~, \\ {\cal L}(H q_i q_j) &=& \frac{g}{m_W \sin 2\beta} \bar q_i \left\{ -m_{q_i} c_\alpha s_\beta \delta_{ij} + \sin(\beta -\alpha)\cdot \left[ m_{q_i} \Sigma_{ij}^q R + m_{q_j} \Sigma_{ji}^{q \star} L \right] \right\} q_j H ~, \\ {\cal L}(A q_i q_j) &=& - i I_q \frac{g}{m_W \sin 2\beta} \bar q_i \left\{ m_{q_i} s_\beta^2 \gamma_5 \delta_{ij} - \left[ m_{q_i} \Sigma_{ij}^q R - m_{q_j} \Sigma_{ji}^{q \star} L \right] \right\} q_j A ~, \end{eqnarray} \begin{eqnarray} {\cal L}(H^+ u_i d_j) = \sqrt{2} \frac{g}{ m_W \sin 2\beta} \bar u_i \left\{ \left[ m_{d_j} s_\beta^2 \cdot V_{u_id_j} - m_{d_k} V_{ik} \Sigma^{d}_{kj} \right] R + \left[ -m_{u_i} s_\beta^2 \cdot V_{u_id_j} + m_{u_k} \Sigma^{u \star}_{ki} V_{kj} \right] L \right\} d_j H^+ \label{Sff2}~, \end{eqnarray} where $V$ is the $4 \times 4$ CKM matrix, $q=d$ or $u$ for down or up-quarks with $I_d=-1$ and $I_u=+1$, respectively, and $R(L)=\frac{1}{2}\left(1+(-)\gamma_5\right)$. Also, $\Sigma^d$ and $\Sigma^u$ are new mixing matrices where all FCNC effects of the 4G2HDM are encoded. They are obtained after diagonalizing the quark mass matrices and, therefore, depend on the rotation (unitary) matrices of the right-handed down and up-quarks $D_R$ and $U_R$, respectively. In particular, for ${\cal I} \equiv {\rm diag}\left(0,0,0,1\right)$ in Eq.~\ref{eq:LY4G}, we have (see \cite{ourpaper1}):$^{[2]}$\footnotetext[2]{Note that this is in contrast to ``standard" frameworks such as the SM and the 2HDM's of types I and II, where the right-handed mixing matrices $U_R$ and $D_R$ are non-physical, being ``rotated away" in the diagonalization procedure of the quark masses.} \begin{eqnarray} \Sigma_{ij}^d = D_{R,4i}^\star D_{R,4j} ~,~ \Sigma_{ij}^u =U_{R,4i}^\star U_{R,4j} ~. \label{sigma} \end{eqnarray} The Yukawa structure and couplings defined by Eqs.~\ref{eq:LY4G}-\ref{sigma} is assumed to be copied to the leptonic sector, see \cite{ourg-2}. In the following sections \ref{sec3} and \ref{sec4}, for illustrative purposes (and without loss of generality), we will set $\Sigma^{d,u} \to {\rm diag}\left(0,0,0,1\right)$ in both the quark and lepton sectors, so that FCNC effects (in particular, between the 4th generation fermions and the SM fermions) are ``turned off". In fact, from the phenomenological point of view, it is sufficient to assume that $\Sigma^{u}_{34,43} \to 0$ (i.e., forbidding the decay $t^\prime \to t h$) and $V_{i4,4i} \to 0$ ($i=1,2,3$, thus forbidding the decays $t^\prime \to d_i W$ and $b^\prime \to u_i W$ with $d_i=d,s,b$ and $u_i=u,c,t$) in order to accommodate relatively light $t^\prime$ and $b^\prime$ with masses as low as 350 GeV, since the existing stringent exclusion limits of $m_{t^\prime},m_{b^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: 700$ GeV, are based on searches that assume 100\% branching ratios of the 4th generation quarks into one of the channels: $t^\prime \to th,tZ,d_i W$ and $b^\prime \to Zb,u_iW$ \cite{pdg,1509.04261}. We will, therefore, assume that the dominant $t^\prime$ and $b^\prime$ decays are into one of the FC channels $t^\prime \to u_ih$ and $b^\prime \to d_i h$ ($u_i=u,c$ and $d_i=d,s,b$), due to small FCNC entries in $\Sigma^{u,d}$ (which have no effect on the results presented in sections \ref{sec3} and \ref{sec4}), in which case small off-diagonal CKM entries $V_{14,41}$ and/or $V_{24,42}$ are also allowed as long as $BR(t^\prime \to d_i W),~BR(b^\prime \to u_iW) \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 0.5$ \cite{1509.04261}. Such flavor structures, may have interesting phenomenological implications, as will be discussed in section \ref{sec5}. The 2HDM scalar sector is parameterized by seven free parameters (after minimization of the potential), which, in the so called ``physical basis", can be chosen as the four physical Higgs masses ($m_h,~m_H,~m_A,~m_{H^+}$), the two angles $\beta$ and $\alpha$ and one parameter from the scalar potential, which is needed in order to specify the scalar couplings, in particular, $hH^+H^-$ (which enters in the 1-loop $h \to \gamma \gamma$ decay), $HH^+H^-$ (which enters the 1-loop $H \to \gamma \gamma$ decay) and $Hhh$ (required for the decay $H \to hh$). In the physical basis, these scalar couplings can be written at tree-level as (see e.g., \cite{0408364}): \begin{eqnarray} \lambda_{Hhh}=-\frac{\cos(\alpha-\beta)}{2 v \sin 2\beta} \left[ \sin 2 \alpha \left(m_h^2+2m_H^2 \right) - \left( 3 \sin 2 \alpha - \sin 2 \beta \right) \frac{m_{\ell h}^2}{s_\beta c_\beta} \right] \label{lam1} ~, \end{eqnarray} \begin{eqnarray} \lambda_{hH^+H^-}=-\frac{1}{2 v \sin 2\beta} \left[ \left( \cos(\alpha-3\beta)+3 \cos(\alpha+\beta) \right) m_h^2 - 4 \sin 2 \beta \sin(\alpha-\beta) m_{H^\pm}^2 - 4 \cos(\alpha+\beta) \frac{m_{\ell h}^2}{s_\beta c_\beta} \right] \label{lam2} ~, \end{eqnarray} \begin{eqnarray} \lambda_{HH^+H^-}=-\frac{1}{2 v \sin 2\beta} \left[ \left( \sin(\alpha-3\beta)+3 \sin(\alpha+\beta) \right) m_H^2 + 4 \sin 2 \beta \cos(\alpha-\beta) m_{H^\pm}^2 - 4 \sin(\alpha+\beta) \frac{m_{\ell h}^2}{s_\beta c_\beta} \right] \label{lam3} ~, \end{eqnarray} where $m_{\ell h}^2$ is a mass-like term, $m_{\ell h}^2 \Phi_\ell^\dagger \Phi_h + h.c.$, which softly breaks the above $Z_2$-symmetry (i.e., $\Phi_{\ell}\to-\Phi_{\ell},~ \Phi_{h}\to+\Phi_{h}$), and which can be used to specify the above tree-level scalar couplings. However, since the working assumption of the 4G2HDM is that the scalar sector may be strongly interacting at the near by few TeV scale, the scalar potential is expected to be subject to significant renormalization and threshold effects. Thus, the above scalar couplings are expected to deviate from their tree-level values, depending on the details of the UV completion and on the masses of the heavy degrees of freedom of this model, see e.g., \cite{0408364,loop-cor}. As an example, consider the 1-loop corrections to the $Hhh$ coupling $\lambda_{Hhh}$, for $|\alpha| \to \pi/2$, in which case there is no mixing between the light and heavy Higgs fields (see Eq.~\ref{Higgsangles}), as required in order to accommodate the 750 GeV $\gamma \gamma$ excess in the 4G2HDM (see section \ref{sec4}). In this limit, the Yukawa couplings of the 4th generation fermions to the light Higgs state $h$ (i.e., $t^\prime t^\prime h$) vanish (see Eq.~\ref{Sff1} and Table \ref{tab4}) and we find that the dominant effect arises from the 1-loop triangle diagram with the charged Higgs exchange in the loop, giving a ``renormalized" $Hhh$ coupling $\bar\lambda_{Hhh} \equiv a_{Hhh} \lambda_{Hhh}$, with: \begin{eqnarray} a_{Hhh} \approx 1+ \frac{m_{\ell h}^4}{m_H^2 v^2} \frac{\left( 1-2c_\beta^2 \frac{m_{H^+}^2}{m_{\ell h}^2} \right) \left(1+c_\beta^2 \frac{m_{H}^2}{m_{\ell h}^2} - 2s_\beta^2 \frac{m_{H^+}^2}{m_{\ell h}^2} \right)}{2 \pi^2 (\sin2\beta)^2} I\left(m_h,m_H,m_{H^+} \right)~, \end{eqnarray} where $I\left(m_h,m_H,m_{H^+} \right)$ is the charged Higgs triangle loop integral, given by: \begin{eqnarray} I\left(m_h,m_H,m_{H^+} \right) = - \int_0^1 dx \int_0^{1-x} dy ~ \frac{1}{ (x+y) (x+y-1) m_h^2 - xy m_H^2 + m_{H^+}^2 } ~. \end{eqnarray} In particular, one roughly finds $|a_{Hhh}| \in \left\{0, 2\right\}$ when $m_{H^+} \in \left\{500~{\rm GeV}, 1~{\rm TeV} \right\}$ and with $m_H =750$ GeV, $m_h =125$ GeV and $m_{\ell h} \sim {\cal O}(1~{\rm TeV})$. For example, $a_{Hhh} \sim -0.15$ for $m_{H^+} = m_H = 750$ GeV and $m_{\ell h} =1.2$ TeV. In what follows we will, therefore, define the ``renormalized" scalar couplings as: $\bar\lambda_i \equiv a_i \lambda_i$, where $\lambda_i$ ($i=Hhh,~ hH^+H^-,~HH^+H^-$) are the corresponding tree-level couplings in Eqs.~\ref{lam1}-\ref{lam3}, and $a_i$ will be treated as free-parameters in the fit that will be varied in the range $|a_i| \in \left\{0, 2\right\}$. \section{The 125 GeV Higgs signals and PEWD \label{sec3}} The measured signals of the 125 GeV Higgs particle, which in the 4G2HDM is the light Higgs $h$, and PEWD impose stringent constraints on the free parameter space of the 4G2HDM. For the 125 GeV Higgs signals we use the measured values of the ``signal strength" parameters, which are defined as the ratio between the measured rates and their SM expectation. In particular, for a specific production and decay channel $i \to h \to f$, the signal strength is defined as: \begin{eqnarray} \mu_{i}^f \equiv \mu_i \cdot \mu^f ~, \end{eqnarray} with \begin{eqnarray} \mu_i = \frac{\sigma(i \to h)}{\sigma(i \to h)_{SM}} = k_i^2~,~ \mu^f = \frac{BR(h \to f)}{BR(h \to f)_{SM}} = \frac{k_f^2}{R^T} ~, \end{eqnarray} where $k_j$ is the 4G2HDM coupling involved in $j \to h$ or $h \to j$ production or decay processes, normalized by its SM value, and $R^T$ is the ratio between the total width of $h$ in the 4G2HDM and the total width of the SM 125 GeV Higgs. In particular, \begin{eqnarray} k_j \equiv \frac{k_j^{4G2HDM}}{k_j^{SM}} ~,~ R^T \equiv \frac{\Gamma_{h_{4G2HDM}}^{Total}}{\Gamma_{h_{SM}}^{Total}}~, \end{eqnarray} so that $\mu_i^f = k_i^2 k_f^2 / R^T$. In Table \ref{tab1} we list the latest combined ATLAS and CMS six parameter fit from RUN1 \cite{ATLAS-CMS-125res}, of the measured values for $\mu_{gg}^{\gamma \gamma},~\mu_{gg}^{WW^\star}, ~\mu_{gg}^{ZZ^\star},~\mu_{gg}^{bb},~\mu_{gg}^{\tau \tau}$ and $\mu_V/\mu_{gg}$, where $\mu_V$ stands for Higgs production via vector-boson fusion (VBF) or in association with a vector-boson (VH).$^{[3]}$\footnotetext[3]{We neglect Higgs production via $pp \to tth$ which, although included in the fit, is 2-3 orders of magnitudes smaller than the gluon-fusion channel}. We also write in Table \ref{tab1} the model predictions for the various signal strengths in terms of the normalized couplings defined above. \begin{table}[htb] \begin{center} \begin{tabular}{c||c|c|} & measured value & model prediction / couplngs \\ \hline \hline $\mu_{gg}^{\gamma \gamma} $ & $1.13^{+0.24}_{- 0.21}$ & $k_g^2 k_\gamma^2 /R^T $ \\ \hline $\mu_{gg}^{ZZ^\star} $ & $1.29^{+0.29}_{- 0.25}$ & $k_g^2 k_V^2 /R^T$ \\ \hline $\mu_{gg}^{WW^\star} $ & $1.08^{+0.22}_{- 0.19}$ & $k_g^2 k_V^2 /R^T$ \\ \hline $\mu_{gg}^{bb} $ & $0.65^{+0.37}_{- 0.28}$ & $k_g^2 k_b^2/R^T$ \\ \hline $\mu_{gg}^{\tau \tau} $ & $1.07^{+0.35}_{- 0.28}$ & $k_g^2 k_\tau^2 /R^T$ \\ \hline $\mu_V/\mu_{gg} = $ & $1.06^{+0.35}_{- 0.27}$ & $k_V^2 / k_g^2$ \\ \hline \end{tabular} \caption{Measured values \cite{ATLAS-CMS-125res} and model predictions in terms of normalized couplings (see text) of the various production and decay channels for the 125 GeV Higgs, using the signal strength prescription. Note that while $k_V,k_b$ and $k_\tau$ are ratios of tree-level couplings, $k_g$ and $k_\gamma$ are the normalized (with respect to the SM) 1-loop 4G2HDM couplings $hgg$ and $h\gamma \gamma$, respectively, calculated using the formula in \cite{HHG}. Also, in our 4G2HDM $k_W=k_Z=k_V$.} \label{tab1} \end{center} \end{table} For the PEWD constraints on the 4G2HDM, we update our study in \cite{ourpaper1}. In particular, the effects of any new physics can be divided into those which do and which do not couple directly to the ordinary SM fermions. For the former, the leading effect in the 4G2HDM comes from the decay $Z \to b \bar b$, which is mainly sensitive to the $H^+ t^\prime b$ and $W^+ t^\prime b$ couplings through one-loop exchanges of $H^+$ and $W^+$, as was analyzed in detail in \cite{ourpaper1}. These contributions to $Z \to b \bar b$ are, however, absent in the currently studied versions of the 4G2HDM, since our working assumption here is that $V_{t^\prime b} \to 0$ and $\Sigma^{d,u} \to {\rm diag}\left(0,0,0,1\right)$, so that the $H^+ t^\prime b$ and $W^+ t^\prime b$ vertices vanish or are negligibly small (see previous section). The effects which do not involve direct couplings to the ordinary fermions, can be analyzed in the formalism of the oblique parameters S,T and U \cite{peskin}. The contribution of a 2HDM with a 4th generation of chiral fermions to the oblique parameters were studied in \cite{ourpaper1}. This includes the pure 1-loop Higgs exchanges to the gauge-bosons 2-point functions and the 1-loop exchanges of $t^\prime$ and $b^\prime$ which shift the T parameter and which involve the new SM4-like diagonal coupling $W t^\prime b^\prime$ (here also the contributions involving the off-diagonal couplings $W t^\prime b$ and $W t b^\prime$ are absent since we assume $V_{t^\prime b},~ V_{t b^\prime} \to 0$, see also \cite{deltaT}). These are calculated with respect to the SM values and are bounded by a global fit to PEWD \cite{gfitter}: \begin{eqnarray} \Delta S &=& S - S_{SM} = 0.06 \pm 0.09 ~, \nonumber \\ \Delta T &=& T - T_{SM} = 0.1 \pm 0.07 \label{SandT}~, \end{eqnarray} with a correlation coefficient of $\rho = +0.91$. These values are obtained for $\Delta U=0$ (the $U$ parameter is often set to zero since it can be neglected in most new physics models and, in particular in our 4G2HDM) and with the SM reference values $M_{H,{\rm ref}}=125$ GeV and $m_{t,{\rm ref}}=173$ GeV. We, thus, consider below the constraints from the 2-dimensional ellipse in the $S-T$ plane which, for a given confidence level (CL), is defined by: \begin{widetext} \begin{eqnarray} \left( \begin{array}{c} S - S_{exp} \\ T - T_{exp} \end{array} \right)^T \left( \begin{array}{cc} \sigma_S^2 & \sigma_S \sigma_T \rho \\ \sigma_S \sigma_T \rho & \sigma_T^2 \end{array} \right) \left( \begin{array}{c} S - S_{exp} \\ T - T_{exp} \end{array} \right) = - 2 {\rm ln}\left( 1 - CL \right) \label{ST2}~, \end{eqnarray} \end{widetext} where $S_{exp} = 0.06$ and $T_{exp} = 0.1$ are the best fitted (central) values, $\sigma_S = 0.09, \sigma_T = 0.07$ are the corresponding standard deviations and $\rho=0.91$ is the (strong) correlation factor between S and T. We thus perform a random (``blind") scan of the relevant parameter space, imposing compatibility at 95\% CL of the 4G2HDM with the measured 125 GeV Higgs signals listed above and with the best fitted values of $S$ and $T$ using Eqs.~\ref{SandT} and \ref{ST2}. In particular, we fix $m_H=750$ GeV (for compatibility with the recent 750 GeV $\gamma \gamma$ signal, see next section) and scan the rest of the parameters over the following ranges: \begin{eqnarray} \alpha \in \left[ - \frac{\pi}{2},\frac{\pi}{2} \right] ~,~ \tan\beta \in \left[ 0.4 , 10 \right]~,~ a_i\in \left[ -2,2 \right] ~ (i=hH^+H^-,HH^+H^-,Hhh) ~, \nonumber \end{eqnarray} \begin{eqnarray} m_{\ell h}^2 \in \left[ - \left(2~ {\rm TeV} \right)^2,\left(2~ {\rm TeV} \right)^2 \right] ~,~ m_{A,H^+} \in \left[ 300~ {\rm GeV} ,1.5~ {\rm TeV} \right] ~, \nonumber \end{eqnarray} \begin{eqnarray} m_{t^\prime,b^\prime} \in \left[ 350~{\rm GeV} , 500~{\rm GeV} \right] ~,~ m_{\nu^\prime,\tau^\prime} \in \left[ 200~{\rm GeV} , 1200~{\rm GeV} \right] ~. \end{eqnarray} \begin{figure} \begin{center} \includegraphics[scale=0.27]{SandT_case1.eps} \includegraphics[scale=0.27]{alpha_vs_tb_case1.eps} \includegraphics[scale=0.27]{deltam_case1.eps} \includegraphics[scale=0.27]{signal_strength_case1.eps} \includegraphics[scale=0.27]{SandT_case3.eps} \includegraphics[scale=0.27]{alpha_vs_tb_case3.eps} \includegraphics[scale=0.27]{deltam_case3.eps} \includegraphics[scale=0.27]{signal_strength_case3.eps} \end{center} \caption{The distribution of the 4G2HDM parameter space that is compatible with the 125 GeV signals and PEWD. From left to right: in the $S -T$ plane (yellow, pink and green ellipses correspond to the 68\%, 95\% and 99\% CL allowed contours, respectively), in the $\tan\beta - \sin\alpha$ plane, in the $\Delta m_{\ell^\prime} - \Delta m_{q^\prime}$ plane, where $\Delta m_{\ell^\prime} \equiv m_{\nu^\prime} - m_{\tau^\prime}$ and $\Delta m_{q^\prime} \equiv m_{t^\prime} - m_{b^\prime}$ and the corresponding 125 GeV signal strengths (on right). Case 1 in the upper row, case 2 in the middle row and case 3 in the lower row, see text. \label{fig1}} \end{figure} We find two types of possible 4G2HDM ``solutions": \begin{description} \item{case 1:} $\tan\beta \leq 0.5$, $\sin\alpha \to -1$ and $m_{\ell h}^2 > 0$. \item{case 2:} $\tan\beta \geq 2$, $\sin\alpha \sim 0.1 - 0.45$ and any $m_{\ell h}^2$ in the entire range scanned. \end{description} In both cases above, $m_A,~m_{H^+}$ and the 4th generation fermion masses can have values spanning over the entire scan ranges. In Fig.~\ref{fig1} we plot the resulting distributions of the relevant parameter space in the $S -T$, $\tan\beta - \sin\alpha$ and $\Delta m_{\ell^\prime} - \Delta m_{q^\prime}$ planes, where $\Delta m_{\ell^\prime} \equiv m_{\nu^\prime} - m_{\tau^\prime}$ and $\Delta m_{q^\prime} \equiv m_{t^\prime} - m_{b^\prime}$. We also show in Fig.~\ref{fig1} the resulting predicted 125 GeV Higgs signal strengths for the two cases above, which, as can seen, have different characteristics. We next discuss the compatibility of the above two 4G2HDM solutions with the recently observed 750 GeV $\gamma \gamma$ excess. \section{The 4G2HDM and the 750 GeV $\gamma \gamma$ resonance \label{sec4}} We search here for the portion of parameter space of the two 4G2HDM cases found in the previous section, that survive once the 4G2HDM is also required to accommodate the 750 GeV $\gamma \gamma$ excess, which is being interpreted here as the decay of one or both of the heavy neutral Higgs (i.e., assumed to have masses $\sim 750$ GeV) $H \to \gamma \gamma$ and/or $A \to \gamma \gamma$. Given the exploratory nature of our study, we will simplify our analysis at this point, assuming that the scalar spectrum have the characteristics of the so-called decoupling limit (see e.g., \cite{decouplinglimit}). In particular, we assume that it is split into 2 typical scales: $m_{light} \sim 125$ GeV, corresponding to the observed light Higgs and $m_{heavy} \sim 750$ GeV around which the three heavy Higgs masses lie, i.e., $m_H,m_A,m_{H^+} \sim 750$ GeV. Even though we find a wider range of allowed masses for the non-resonant heavy scalar states (i.e., for $m_A$ and $m_{H^+}$, see below) that can accommodate the 750 GeV signal, the choice $m_H,~m_A,m_{H^+} \sim 750$ GeV will suffice for conveying our point: that the 750 GeV resonance in the $\gamma\gamma$ channel can be accommodated by one of the heavy scalars of the 4G2HDM without any conflict with other existing relevant data. Indeed, if this measurement will be eventually confirmed, then it will be instructive to study the 4G2HDM within a wider range of the relevant parameter space. We, thus, re-scan the 4G2HDM parameter space corresponding to two 4G2HDM cases found in the previous section, where now $m_H$, $m_A$ and $m_{H^+}$ are varied within a 30 GeV mass range around 750 GeV, i.e., $m_{H,A,H^+} \in 750 \pm 30$ GeV. The scan is performed with the following additional ``filters"/requirements (i.e., in addition to the requirement of compatibility with PEWD and with the measured 125 GeV Higgs signals, as outlined in the previous section): \begin{itemize} \item Reproducing the 750 GeV $\gamma \gamma$ excess within the range $3 ~{\rm fb} < \sigma(pp \to H/A \to \gamma \gamma) < 13~{\rm fb}$. We find that the (by far) dominant $H$ and/or $A$ production mechanism is the gluon-fusion one $gg \to H/A$, so that all the relevant cross-sections $\sigma(pp \to H/A \to f)$ are calculated in the narrow width approximation via: \begin{eqnarray} \sigma(pp \to H/A \to f) = \frac{C_{gg}}{s m_{H/A}} \Gamma(H/A \to gg) BR(H/A\to f) ~, \end{eqnarray} where $\sqrt{s} =8$ or $13$ TeV and $C_{gg}$ is the gluon luminosity: \begin{eqnarray} C_{gg} = \frac{\pi^2}{8} \int_{m_{H/A}^2/s}^1 \frac{dx}{x} g(x) g\left( \frac{m_{H/A}^2}{sx} \right) ~, \end{eqnarray} giving $C_{gg} \sim 2140(175)$ at $\sqrt{s} =13(8)$ TeV, see \cite{gluonPDF}. \item The resonating scalar which produces the 750 GeV $\gamma \gamma$ excess is required to have a width smaller than 45 GeV, i.e., $\Gamma_{H/A} < 45$ GeV. \item We impose the existing experimental bounds on the production and decays of the heavy neutral scalars $H$ and $A$, as obtained at the 8 and 13 TeV LHC runs (in particular when applied to $m_H,m_A \sim 750$ GeV) in all other channels which are relevant to our study: $pp \to W^+W^-,~ZZ,~t \bar t,~\tau \tau,~b \bar b,~hh,~hZ$. In particular, we use the 95\% CL bounds in Table \ref{tab2} quoted in \cite{1605.09401}. \end{itemize} \begin{table}[htb] \begin{center} \begin{tabular}{c||c|c|} final state & $\sigma$ at $\sqrt{s}=8$ TeV & $\sigma$ at $\sqrt{s}=13$ TeV \\ \hline \hline $pp \to H \to W^+ W^- $ & $< 40 $ fb & $< 300$ fb\\ \hline $pp \to H \to ZZ $ & $< 12 $ fb & $< 200$ fb\\ \hline $pp \to H \to hh $ & $< 39 $ fb & $< 120$ fb\\ \hline $pp \to A \to hZ $ & $< 19 $ fb & $< 116$ fb\\ \hline $pp \to H/A \to t \bar t $ & $< 450 $ fb & \\ \hline $pp \to H/A \to b \bar b $ & $< 1 $ pb & \\ \hline $pp \to H/A \to j j $ & $< 2.5 $ pb & \\ \hline $pp \to H/A \to \tau \tau $ & $< 12 $ fb & $< 60$ fb \\ \hline \end{tabular} \caption{Upper bounds at 95\% CL on $\sigma(pp \to S \to f)$ for various final states $f$, produced through a narrow resonance with $m_S \sim 750$ GeV and $\Gamma_S/m_S \sim {\cal O}(10^{-2})$, as applied to our scan with $S = H,A$. The bound on $\sigma( pp \to H/A \to jj)$ is relevant for $j = {\rm gluon}$. Table taken from \cite{1605.09401}.} \label{tab2} \end{center} \end{table} Applying the above filters, we find that: \begin{enumerate} \item Only the CP-even scalar state $H$ (with $m_H =750$ GeV), can accommodate the 750 GeV $\gamma \gamma$ resonance, since $\sigma(pp \to A \to \gamma \gamma) \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: {\cal O}(0.01)$ fb, which is 2-3 orders of magnitudes smaller than the measured $\gamma \gamma$ excess, see also Table \ref{tab4}. \item Only a ``shrinked" version of the 4G2HDM case 1 survives out of the two cases that were found to be compatible with PEWD and the 125 GeV light Higgs signals. In particular, the surviving 4G2HDM models have (see Fig.~\ref{fig2}): $\tan\beta \leq 0.5$, $\alpha \to -\pi/2$ and $m_{\ell h} > 600$ GeV, having some correlation with the renormalization factors of the scalar couplings $a_i = \bar\lambda_i/\lambda_i$, $i=Hhh,~ hH^+H^-,~HH^+H^-$. \item The resulting heavy fermions mass ranges are narrowed to: $350 ~ {\rm GeV} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: m_{t^\prime},m_{b^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 390 ~ {\rm GeV}$, where the lower limit is from direct searches (see section \ref{sec2}), and $900 ~ {\rm GeV} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: m_{\nu^\prime},m_{\tau^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 1200 ~ {\rm GeV}$, where the upper limit is a rough estimate of the perturbativity bound on heavy chiral leptons. \end{enumerate} In Fig.~\ref{fig2} we show three scatter plots of the resulting 4G2HDM parameter space, corresponding to the mass spectrum of the heavy fermions, the correlation between the soft breaking mass parameter $m_{\ell h}$ and the renormalization factor of the scalar couplings $a_i = \bar\lambda_i/\lambda_i$, $i=Hhh,~ hH^+H^-,~HH^+H^-$, and the resulting allowed ranges of the 125 GeV light Higgs signal strengths in all the measured channels. We see that, while $|m_{t^\prime} - m_{b^\prime}| \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 30$ GeV, the mass splitting of the heavy leptons is typically $|m_{\nu^\prime} - m_{\tau^\prime}| \:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: m_W$. We also see that smaller values of $m_{\ell h}$ typically require larger values of the renormalization factors of the scalar vertices $a_i$, e.g., $a_{Hhh} \sim 1$ for $m_{\ell h} \sim 700$ GeV. \begin{figure} \begin{center} \includegraphics[scale=0.37]{Fig2_deltam.eps} \includegraphics[scale=0.37]{fig2_ai_vs_mlh.eps} \includegraphics[scale=0.37]{fig2_125_signals.eps} \end{center} \caption{Scatter plots of the 4G2HDM parameter space that is compatible with the 125 GeV signals, with PEWD, with $\sigma(pp \to H\to \gamma \gamma) = 3-13$ fb, with $\Gamma_H \leq 45$ GeV and with all 8 and 13 TeV LHC bounds on the cross-section $\sigma(pp \to H/A \to f)$ in all final states $f$ relevant to the $H$ and $A$ decays, see Table \ref{tab2}. The scatter plots are given for the mass splitting spectrum of the heavy fermions (left), the correlation between the soft breaking mass parameter $m_{\ell h}$ and the renormalization factor of the scalar couplings $a_i = \bar\lambda_i/\lambda_i$, $i=Hhh,~ hH^+H^-,~HH^+H^-$ (middle), and the resulting allowed ranges of the 125 GeV light Higgs signal strengths in all the measured channels (right). \label{fig2}} \end{figure} It is interesting to note that the resulting mass spectrum of the heavy chiral quarks, which is required to accommodate the 750 GeV $\gamma \gamma$ resonance, is rather narrow and roughly centered around $m_H /2$, i.e., $m_{t^\prime},m_{b^\prime} \sim 350 - 390 ~ {\rm GeV}$. This may hint back to the possibility that the the heavy scalars are composites primarily of the heavy chiral quarks, in which case the 4G2HDM might indeed be interpreted as a low energy effective framework for some TeV-scale strongly interacting theory. Such an effective low energy 2HDM, with features similar to the 4G2HDM discussed here, was introduced in \cite{ourhybrid}, where it was shown that, using the Nambu-Jona-Lasinio (NJL) mechanism \cite{Nambu}, it is possible to construct an effective sub-TeV 2HDM hybrid framework, in which the 125 GeV light Higgs is mostly a fundamental scalar, while the heavy Higgs states are components of a composite field of the form $\Phi_h \sim g_{t^\prime}^\star < \bar Q_L^{\prime c} (i \tau_2) t^{\prime c}_R > + g_{b^\prime} < \bar Q_L^\prime b^\prime_R >$, which is responsible for EW symmetry breaking and for the dynamical mass generation of the heavy quarks.$^{[4]}$\footnotetext[4]{Another interesting framework which entertains the idea that heavy chiral quarks may form the 750 GeV composite was recently suggested in \cite{1602.05539}.} \section{Phenomenology of the 4G2HDM \label{sec5}} Inspired by the indications of the 750 GeV $\gamma \gamma$ resonance and following the analysis of the previous section, we briefly consider here some of the distinct phenomenological consequences of the 4G2HDM with characteristics similar to those required to accommodate such a heavy scalar resonance. In particular, we will assume below that $\tan\beta \sim 0.5$ and $\sin\alpha \sim -1$, in which case the light 125 GeV Higgs of the 4G2HDM, $h$, does not couple to $f^\prime f^\prime$, while the heavy CP-even Higgs, $H$, does not couple to a pair of SM fermions (see Eqs.~\ref{Sff1}-\ref{Sff2} and Table \ref{tab3}). Also, the 4th generation heavy fermions are assumed to have masses in the ranges $350 ~ {\rm GeV} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: m_{t^\prime},m_{b^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 400 ~ {\rm GeV}$ and $900 ~ {\rm GeV} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: m_{\nu^\prime},m_{\tau^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 1200 ~ {\rm GeV}$, and the dominant decay channels of the heavy quarks are $t^\prime \to uh$ ($u=u,c$) and $b^\prime \to dh$ ($d=d,s,b$), with corresponding branching ratios $ \:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: 0.5$, due to small off diagonal-entries $\Sigma^u_{4i}$ ($i=1,2$) and/or $\Sigma^d_{4i}$ ($i=1,2,3$) (see Table \ref{tab3} and discussion in section \ref{sec2}). \begin{table}[htb] \begin{center} \begin{tabular}{|c||c|c|c|} & \multicolumn{3}{c|}{Yukawa couplings in the 4G2HDM with $\sin\alpha \sim -1$} \\ \hline & $v \cdot y(\bar f f)$ & $v \cdot y(\bar f^\prime f^\prime)$ & $v \cdot y(\bar f_i f_{j})$ ($i, j = 1-4, i\neq j$) \\ \hline \hline $h$ & $-\frac{m_f}{\cos\beta}$ & 0 & $\frac{\Sigma^f_{ij}}{\cos\beta} (m_{f_i} R + m_{f_j} L)$ \\ \hline $H$ & 0 & $\frac{m_f}{\sin\beta}$ & $ \frac{\Sigma^f_{ij} }{\sin\beta} (m_{f_i} R + m_{f_j} L)$ \\ \hline $A$ & $-i I_f m_f \tan\beta$ & $i I_f m_{f^\prime} \cot\beta$ & $i I_f \frac{\Sigma^f_{ij}}{\sin\beta \cos\beta} (m_{f_i} R - m_{f_j} L)$ \\ \hline \end{tabular} \caption{Yukawa couplings of the neutral Higgs particles in the 4G2HDM with $\sin\alpha \to -1$ and assuming $\Sigma^f_{ij} \ll \Sigma^f_{44} =1$ for $ij \neq 44$, see section \ref{sec2}. In the first column $f$ is a SM fermion of the 1st-3rd generations, while in the second column $f^\prime$ stands for a 4th generation fermion. In the 3rd column $f_i$ correspond to any fermion of the $i$th generation. Also, $I_f = 1(-1)$ for up(down) type fermions.} \label{tab3} \end{center} \end{table} In Table \ref{tab4} we list three benchmark points (BMP1,BMP2,BMP3) which have some distinct characteristics and which are compatible with PEWD, with the 125 GeV Higgs signals, with the 750 GeV $\gamma \gamma$ signal and with the LHC bounds on all relevant 750 GeV Higgs resonance channels $pp \to H/A \to f$ given in Table \ref{tab2}. For definiteness, we have generated the benchmark points for the case of $m_H = 750 $ GeV and $m_A,m_{H^+} \sim m_H \pm 50$ GeV, but the discussion below has a more general scope, i.e., with regard to some of the possible phenomenological signatures of the 4G2HDM associated with the TeV-scale heavy scalars of the model and independent of whether the 750 GeV $\gamma \gamma$ resonance is confirmed or not. The three benchmark points include cases where the 750 GeV Higgs total width ranges from a few GeV to $\sim 45$ GeV, having a resonance cross-section to $\gamma \gamma$ between 4-12 fb. They also correspond to cases where $BR(H/A \to \bar q^\prime q^\prime) \sim 1$ and $BR(H^+ \to \bar q^\prime q^\prime) \sim 1$. \begin{table}[htb] \begin{center} \begin{tabular}{|c||c|c|c|} \hline & BMP1 & BMP2 & BMP3 \\ \hline \hline $m_{t^\prime},m_{b^\prime}, m_{H^+}, m_A$ [GeV] & $352,382,709,780$ & $384,373,795,778$ & $368,369,691,731$ \\ \hline $H$ total width, $\Gamma_H$ [GeV] & 43 & 4 & 17 \\ \hline $\sigma(pp \to H/A \to \gamma \gamma)$ [fb] & $4/0.004$ & $12/0.006$ & $8/0.005$ \\ \hline $BR(H \to \bar t^\prime t^\prime, \bar b^\prime b^\prime, \bar t t, hh,gg)$ & $0.94,0,{\cal O}(10^{-6}),{\cal O}(10^{-4}),0.06$ & $0,0.36,0.04,0.007,0.64$& $0.47,0.38,{\cal O}(10^{-6}),{\cal O}(10^{-4}),0.14$ \\ \hline $BR(A \to \bar t^\prime t^\prime, \bar b^\prime b^\prime, \bar t t, hZ,gg)$ & $0.61,0.33,0.015,0.04,{\cal O}(10^{-4})$ & $0.33,0.61,0.016,0.05,0.004$ & $0,0,0.29,0.71,{\cal O}(10^{-4})$ \\ \hline $\sigma(pp \to H \to \bar t^\prime t^\prime, \bar b^\prime b^\prime, \bar t t, hh)$ [fb] & $15000,0,0.015,5$ & $0,4000,0.65,109$ & $7500,6000,0.02,9$ \\ \hline $\sigma(pp \to A \to \bar t^\prime t^\prime, \bar b^\prime b^\prime, \bar t t, hZ)$ [fb] & $160,87,19,52$ & $90,164,21,57$ & $0,0,38,91$ \\ \hline $BR(H^+ \to t^\prime \bar b^\prime, t \bar b ,W^+h)$ & $0,0.31,0.69$ & $0.9,0.03,0.07$ & $0,0.32,0.68$ \\ \hline \end{tabular} \caption{Benchmark points with some distinct characteristics, which are consistent with PEWD, with the 125 GeV Higgs signals, with the 750 GeV $\gamma \gamma$ signal and with the LHC bounds on all relevant 750 GeV Higgs resonant channels $pp \to H/A \to f$ given in Table \ref{tab2}.} \label{tab4} \end{center} \end{table} In particular, if $m_H,m_A > m_{q^\prime}/2$, then $H/A \to \bar q^\prime q^\prime$ is open and typically dominates, having a branching ratio of ${\cal O}(1)$ (see Table \ref{tab4}). In that case, we find that within the 4G2HDM parameter space discussed here, the corresponding resonance cross-sections for $\bar q^\prime q^\prime$ production at the 13 TeV LHC are typically $\sigma(pp \to H \to q^\prime q^\prime) \sim {\cal O}(10)$ [pb] and $\sigma(pp \to A \to q^\prime q^\prime) \sim {\cal O}(0.1)$ [pb], (both $H$ and $A$ produced through gluon-fusion $gg \to H/A$), so that in the case of $H \to q^\prime q^\prime$ (see Table \ref{tab4}), this is about an order of magnitude larger than the QCD (continuum) $\bar q^\prime q^\prime$ production rate. Therefore, if the 750 GeV $\gamma \gamma$ resonance persists, one should also expect an observable resonance signal at least in the $H \to \bar q^\prime q^\prime$ channel. Let us, therefore, briefly investigate the signal $H \to \bar q^\prime q^\prime$ under more general grounds, i.e., when $m_H > m_{q^\prime}/2$ but not necessarily $m_H \sim 750$ GeV. For example, in the case of $H \to \bar t^\prime t^\prime$, the $t^\prime$ will further decay either via the FC channels $t^\prime \to uh$ ($u=u$ or $c$) or via the 3-body decay $t^\prime \to b^\prime W \to dhW$ ($d=d,s,b$), where $b^\prime W$ are either off-shell or on-shell (i.e., when $m_{t^\prime} > m_{b^\prime}+m_W$, see Fig.~\ref{fig1}). If the former case (i.e., $t^\prime \to uh$) dominates, then the resulting resonance signal should be searched for in $pp \to \bar t^\prime t^\prime \to (jh)_{t^\prime}(jh)_{t^\prime}$ ($j$ is a light jet), while if the 3-body $t^\prime$ decay dominates then $pp \to \bar t^\prime t^\prime \to (jhW^+)_{t^\prime}(jhW^-)_{t^\prime}$. In either case, the SM-like light Higgs ($h$) further decays into $b \bar b$ or $WW$ with SM rates, giving rise to resonance signatures of the form $pp \to (nj+mb+\ell W)_H$, with $(n,m,\ell)=(2,4,0),(2,0,4),(2,2,2),(2,4,2),(2,2,4),(2,0,6),(0,2,6),(0,4,4),(0,6,2)$ and with unique kinematic features that distinguishes them from more conventional signatures. Similar signals are also expected for $H \to \bar b^\prime b^\prime$. We recognize that these type of signals are very challenging and may require new strategies, in particular, for reconstructing the parent $q^\prime$'s in such a high jet-multiplicity environment. The decay pattern of the charged Higgs may also change in the 4G2HDM, in particular for the case when $m_{H^+} > m_{t^\prime} + m_{b^\prime}$, for which the decay of $H^+$ into a pair of heavy 4th generation fermions can dominate (see BMP1 in Table \ref{tab4}). In particular, taking $m_{t^\prime} \sim m_{b^\prime} \equiv m_{q^\prime}$ and assuming that $H^+$ is sufficiently heavier than $2m_{q^\prime}$, so that we can ignore corrections of ${\cal O}(4m_{q^\prime}^2/m_{H^+}^2)$ in the phase-space factors, we have in the 4G2HDM: \begin{eqnarray} R_{t^\prime b^\prime/tb} &\equiv& \frac{\Gamma(H^+ \to t^\prime b^\prime)}{\Gamma(H^+ \to t b)} \sim 2 \frac{m_{q^\prime}^2}{m_t^2} \cot^4\beta ~, \\ R_{t^\prime b^\prime/Wh} &\equiv& \frac{\Gamma(H^+ \to t^\prime b^\prime)}{\Gamma(H^+ \to Wh)} \sim 12 \frac{m_{q^\prime}^2}{m_{H^+}^2} \left(\frac{\cot\beta}{\cos(\beta -\alpha)}\right)^2 ~. \end{eqnarray} Thus, for $\alpha \sim - \pi/2$, $\tan\beta \sim 0.5$ (i.e., $\cos(\beta - \alpha) \sim -0.45$), $m_{q^\prime} \sim 350$ GeV (i.e., values of the 4G2HDM parameter space that can accommodate the 750 GeV $\gamma \gamma$ signal) and taking $m_{H^+} \sim {\cal O}(1)$ TeV, we obtain: $R_{t^\prime b^\prime/tb} \sim {\cal O}(100)$ and $R_{t^\prime b^\prime/Wh} \sim {\cal O}(10)$, in which case $BR(H^+ \to t^\prime b^\prime) \sim 1$ (e.g., as in the case of BMP2), leading to some interesting signatures of the heavy charged Higgs at the LHC. In particular, the dominant production channels of $H^+$ at the LHC are $gg/gb \to H^+ b \bar t, H^+ W^-/ H^+ \bar t$, with a typical cross-section of $\sim 100$ fb when $\tan\beta \sim 1$ \cite{Hplusprod}. The subsequent $H^+$ decay to a pair of 4th generation heavy fermions with $BR(H^+ \to t^\prime \bar b^\prime) \sim 1$ will, thus, lead to new $H^+$ signals, e.g., $pp \to t (t^\prime b^\prime)_{H^+} \to (b W)_t (jh)_{t^\prime} (jh)_{b^\prime}$, again with the typical 4G2HDM heavy fermion high jet-multiplicity signatures of the form $pp \to nj+mb+\ell W$. This is in contrast to ``standard" 2HDM frameworks where the heavy charged Higgs will dominantly decay to $Wh$ and/or $tb$ (see BMP1 and BMP3), leading to a lower multiplicity of jets in the final state. As noted earlier, a wider range of solutions exist (which are not being discussed here) to all data and filters mentioned above (i.e., including the 750 GeV $\gamma \gamma$ resonance), in which lighter pseudoscalar $A$ and charged Higgs $H^{+}$ are allowed, with masses as low as $300$ GeV. In such 4G2HDM scenarios, the heavy 4th generation quarks (and leptons) can have substantial decay rates in channels involving also the heavy Higgs species, i.e., $t^\prime \to H^+ d, Au$ ($d=d,s,b$ and $u=u,c)$ and $b^\prime \to H^+ u, Ad$ ($d=d,s,b$ and $u=u,c)$, followed by $H^+ \to W^+ h, t \bar b$ and $A \to hZ, t \bar t$. Indeed, such decay patterns can also lead to some un-explored collider signatures of the 4G2HDM. We leave the discussion of the phenomenology of such wider range of 4G2HDM scenarios to a later work. Finally, we wish to comment on the flavor violating structure of the 4G2HDM and its compatibility with the recently reported indications of the LFV decay of the 125 GeV light Higgs $h \to \tau \mu$ \cite{ATLAS2,CMS2}. Writing the LFV couplings of $h$ in a general form: \begin{eqnarray} {\cal L}(h f_i f_j) = {\cal S}_{ij} + {\cal P}_{ij} \gamma_5 ~, \end{eqnarray} one obtains: \begin{eqnarray} \Gamma(h \to \bar f_i f_j + \bar f_j f_i) = \frac{m_h}{4 \pi} \left( |{\cal S}_{ij}|^2 + |{\cal P}_{ij}|^2 \right) ~. \end{eqnarray} In our 4G2HDM we have for the case of the LFV decay $h \to \tau \mu$ (neglecting terms of ${\cal O}(m_\mu/m_\tau)$, see Eq.~\ref{Sff1}): \begin{eqnarray} |{\cal S}_{\tau \mu}| = |{\cal P}_{\tau \mu}| \sim \frac{g}{4} \frac{m_\tau}{m_W} f(\beta,\alpha) \xi_{\tau \mu}~, \end{eqnarray} where we have defined $\Sigma^\ell_{32} = \Sigma^\ell_{23} \equiv \xi_{\tau \mu}$ (see Eq.~\ref{sigma}) and: \begin{eqnarray} f(\beta,\alpha) = \frac{\cos(\beta - \alpha)}{s_\beta c_\beta} ~. \end{eqnarray} Requiring now that $BR(h \to \tau \mu) \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 1\%$ we find: \begin{eqnarray} | f(\beta,\alpha) \xi_{\tau \mu}| \sim {\cal O}(0.1)~. \end{eqnarray} Thus, since for the values of $\tan\beta$ and $\alpha$ that were found to be compatible with all data considered in the previous sections, we find $|f(\beta,\alpha)| \sim 1-5$, and specifically $f(\beta,\alpha) \sim 1$ for $\alpha \to -\pi/2$ and $\tan\beta \sim 0.5$, as required in order to accommodate the 750 GeV $\gamma \gamma$ resonance (see previous section), the 4G2HDM with $|\xi_{\tau \mu}| \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 0.1$ can address the measured $BR(h \to \tau \mu) \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 1\%$ if it persists. \section{Summary \label{sec6}} We have revisited a class of models beyond the SM, suggested by us a few years ago in \cite{ourpaper1}, which put together an additional Higgs doublet with a heavy chiral 4th generation quark and lepton doublet and which have several important and attractive theoretical features. In particular, we focused on the so-called 4G2HDM of type I (in \cite{ourpaper1}), in which a discrete $Z_2$ symmetry couples the ``heavy" scalar doublet only to the heavy 4th generation fermions and the ``light" one to the lighter SM fermions. We have confronted this model with PEWD, with the measured 125 GeV light Higgs signals and also studied its compatibility with the recent indication of a 750 GeV $\gamma \gamma$ resonance and with the current LHC bounds on heavy scalar resonances in other relevant channels. We found that the CP-even heavy Higgs state of the 4G2HDM with a mass $\sim 750$ GeV can accommodate the measured $750$ GeV excess for a rather unique choice of the parameter space: $\tan\beta \sim 0.5$, $\alpha \sim -\pi/2$ (the Higgs mixing angle) and with heavy chiral fermion masses $m_{t^\prime,b^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 400$ GeV and $m_{\nu^\prime,\tau^\prime} \:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: 900$ GeV. We have shown that the heavy chiral quarks (and leptons) of the 4G2HDM may have FCNC decays into the light 125 GeV Higgs plus a light-quark jet, $q^\prime \to j h$, with branching ratios of ${\cal O}(1)$, thus leading to some un-explored signatures of $q^\prime \bar q^\prime$ production at the LHC and, therefore, being consistent with the current direct bounds on the masses of new heavy fermions. Indeed, new and rich phenomenology in $q^\prime$ - heavy Higgs systems is expected, including possible resonance production of $q^\prime q^\prime$ pairs via either the heavy neutral or heavy charged Higgs particles of the 4G2HDM, which leads to high jet-multiplicity signatures, with or without charged leptons, of the form $\bar q^\prime q^\prime \to nj + mb + \ell W$, with $n+m+\ell=6-8$ and unique kinematic features which are related to the resonating heavy scalar and the decay pattern of the heavy quarks. The reconstruction of the $q^\prime q^\prime$ pairs in such high jet-multiplicity signals is very challenging and require more thought and possibly new search strategies. We also show that the recent indication of a percent-level branching ratio in the LFV decay of the 125 GeV Higgs $h \to \tau \mu$, if it persists, can be readily addressed within the distinct flavor structure of the 4G2HDM. \bigskip \bigskip \bigskip \bigskip \bigskip \bigskip {\bf Acknowledgments:} We thank Pier Paolo Giardino for useful conversations. The work of AS was supported in part by the US DOE contract \#DE-SC0012704. \pagebreak
a0fab30e57b1936f5e7e1addda8de8aea5a298ba
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction}\label{sec:intro} \setlength{\abovedisplayskip}{3pt} \setlength{\abovedisplayshortskip}{3pt} \setlength{\belowdisplayskip}{3pt} \setlength{\belowdisplayshortskip}{3pt} \setlength{\jot}{2pt} \setlength{\floatsep}{2ex} \setlength{\textfloatsep}{2ex} We address the problem of \emph{learning from conditional distributions} where the goal is to learn a function that links conditional distributions to target variables. Specifically, we are provided input samples $\{x_i\}_{i=1}^N\in\Xcal^N$ and their corresponding responses $\{y_i\}_{i=1}^N\in \Ycal^N$. For each $x\in\Xcal$, there is an associated conditional distribution $p(z|x): \Zcal\times \Xcal\rightarrow \RR$. However, we cannot access the entire conditional distributions $\{p(z|x_i)\}_{i=1}^N$ directly; rather, we only observe a limited number of samples or in the extreme case only \emph{one sample} from each conditional distribution $p(z|x)$. The task is to learn a function $f$ which links the conditional distribution $p(z|x)$ to target $y \in \Ycal$ by minimizing the expected loss: \begin{equation}\label{eq:target} \min_{f\in\Fcal}~L(f)=\EE_{x,y}\sbr{\ell\rbr{y, \EE_{z|x}\sbr{f(z,x)}}} \end{equation} where $\ell:\Ycal \times\Ycal \to\RR$ is a convex loss function. The function space $\Fcal$ can be very general, but we focus on the case when $\Fcal$ is a reproducing kernel Hilbert space~(RKHS) in main text, namely, $\Fcal = \{f:\Zcal\times \Xcal\rightarrow \RR \,|\, f(z, x) = \inner{f}{\psi(z, x)}\}$ where $\psi(z, x)$ is a suitably chosen (nonlinear) feature map. Please refer to Appendix~\ref{appendix:extend_dual_embedding} for the extension to arbitrary function approximators, \eg, random features and neural networks. The problem of learning from conditional distributions appears in many different tasks. For example: \begin{itemize}[leftmargin=*,nosep,nolistsep] \item {\bf Learning with invariance.} Incorporating priors on invariance into the learning procedure is crucial for computer vision~\cite{NiyGirPog98}, speech recognition~\cite{AnsLeiRosMut13} and many other applications. The goal of invariance learning is to estimate a function which minimizes the expected risk while at the same time preserving consistency over a group of operations $g=\{g_j\}_{j=1}^\infty$. \citet{MroVoiPog15} shows that this can be accomplished by solving the following optimization problem \begin{equation}\label{eq:invariant} \min_{f\in \tilde\Hcal} \EE_{x, y} [\ell(y, \EE_{z|x\sim \mu(g(x))}[\langle f, \psi(z)\rangle_{\tilde\Hcal}] )] +(\nu/2)\|f\|_{\tilde\Hcal}^2 \end{equation} where $\tilde\Hcal$ is the RKHS corresponding to kernel $\tilde k$ with implicit feature map $\psi(\cdot)$, $\nu>0$ is the regularization parameter. Obviously, the above optimization~(\ref{eq:invariant}) is a special case of (\ref{eq:target}). In this case, $z$ stands for possible variation of data $x$ through conditional probability given by some normalized Haar measure $\mu(g(x))$. Due to computation and memory constraints, one can only afford to generate a few virtual samples from each data point $x$.\\ \item {\bf Policy evaluation in reinforcement learning.} Policy evaluation is a fundamental task in reinforcement learning. Given a policy $\pi(a|s)$ which is a distribution over action space condition on current state $s$, the goal is to estimate the value function $V^\pi(\cdot)$ over the state space. $V^\pi(s)$ is the fixed point of the Bellman equation $$ V^\pi(s) = \EE_{s'|a, s}[R(s,a)+\gamma V^\pi(s')], $$ where $R(s,a):\Scal\times \Acal\to \RR$ is a reward function and $\gamma\in(0, 1)$ is the discount factor. Therefore, the value function can be estimated from data by minimizing the mean-square Bellman error~\cite{Baird95,SutMaeSze08}: \begin{equation}\label{eq:RL_obj} \min_{~V^\pi}~\EE_{s, a}\sbr{\rbr{R(s,a) - \EE_{s'|a, s}\sbr{V^\pi(s) - \gamma V^\pi(s')}}^2}. \end{equation} Restrict the policy to lie in some RKHS, this optimization is clearly a special case of~\eq{eq:target} by viewing $\rbr{(s, a), R(s, a), s'}$ as $(x, y, z)$ in~\eq{eq:target}. Here, given state $s$ and the the action $a\sim \pi(a|s)$, the successor state $s'$ comes from the transition probability $p(s'|a,s)$. Due to the online nature of MDPs, we usually observe only one successor state $s'$ for each action $a$ given $s$,~\ie, only one sample from the conditional distribution given $s, a$.\\ \item{\bf Optimal control in linearly-solvable MDP.} The optimal control in a certain class of MDP, \ie, linearly-solvable MDP, can be achieved by solving the linear Bellman equation~\cite{Todorov06,Todorov09} \begin{equation}\label{eq:linear_optimal_bellman} z(s) = \exp\rbr{-R(s)}\EE_{s'|s\sim p(s'|s)}\sbr{z(s')}, \end{equation} where $R(s)$ denotes the immediate cost and $p(s'|s)$ denotes the passive dynamics without control. With $z(s)$, the trajectory of the optimal control $\pi^*$ can be calculated by $p^{\pi^*}(s'|s) = \frac{p(s'|s)z(s)}{\EE_{s'|s\sim p(s'|s)}\sbr{z(s')}}$. Therefore, $z(\cdot)$ can be estimated from data by optimizing \begin{equation}\label{eq:OC_obj} \min_{~z}~\EE_{s}\sbr{\rbr{z(s) - \EE_{s'|s}\sbr{\exp\rbr{-R(s)}z(s')}}^2}. \end{equation} Restricting function $z(\cdot)$ to lie in some RKHS, this optimization is a special case of~\eq{eq:target} by viewing $\rbr{s, 0, s'}$ as $(x, y, z)$ in~\eq{eq:target}. Here given a state $s$, the successor state $s'$ comes from the passive dynamics. Similar as policy evaluation, we usually observe only one successor state $s'$ given $s$,~\ie, only one sample from the conditional distribution given $s$.\\ \item{\bf Hitting time and stationary distribution in stochastic process.} Estimating the hitting time and stationary distribution of stochastic process are both important problems in social network application and MCMC sampling technique. Denote the transition probability as $p(s'|s)$. The hitting time of $s_j$ starting from $s_i$ is defined as $H(s_i, s_j) = \inf\cbr{n\ge 0; S_n = s_j, S_0 = s_i}$. Hence, the expected hitting time $h(\cdot, \cdot)$ satisfies \begin{eqnarray} h(s_i, s_j) = \begin{cases} 1 + \EE_{s_k\sim p(s|s_i), s_k\neq s_j}\sbr{h(s_k, s_j)} & \quad \text{if } i\neq j\\ 1 + \EE_{s_k\sim p(s|s_i)}\sbr{h(s_k, s_i)} & \quad \text{if } i = j\\ \end{cases}. \end{eqnarray} Based on the property of stochastic process, we can obtain the stationary distribution with $\pi(s_i) = \frac{1}{h(s_i, s_i)}$. The hitting time $h(\cdot, \cdot)$ can be learned by minimizing: \begin{eqnarray}\label{eq:HT_obj} \min_{h}\EE_{s, t}\sbr{\rbr{1 - \EE_{s'\sim \tilde p(s'|s, t)}\sbr{h(s, t) - h(s', t)}}^2}, \end{eqnarray} where $\tilde p(s'|s, t) = p(s'|s)$ if $s=t$, otherwise $\tilde p(s'|s, t) \propto \begin{cases} p(s'|s) & \quad \text{if } s'\neq t\\ 0 & \quad \text{if } s' = t\\ \end{cases}$. Similarly, when restricting the expected hitting time to lie in some RKHS, this optimization is a special case of~\eq{eq:target} by viewing $\rbr{(s, t), 1, s'}$ as $(x, y, z)$ in~\eq{eq:target}. Due to the stochasticity of the process, we only obtain one successor state $s'$ from current state $(s, t)$,~\ie, only one sample from the conditional distribution given $(s, t)$. \end{itemize} \paragraph{Challenges.} Despite the prevalence of learning problems in the form of \eq{eq:target}, solving such problem remains very challenging for two reasons: (\emph{i}) we often have limited samples or in the extreme case only one sample from each conditional distribution $p(z|x)$, making it difficult to accurately estimate the conditional expectation. (\emph{ii}) the conditional expectation is nested inside the loss function, making the problem quite different from the traditional stochastic optimization setting. This type of problem is called {\sl compositional stochastic programming}, and very few results have been established in this domain. \paragraph{Related work.} A simple option to address (\ref{eq:target}) is using sample average approximation (SAA), and thus, instead solve $$ \min_{f\in\Fcal} \frac{1}{N} \sum_{i=1}^N \sbr{\ell\rbr{y_i, \frac{1}{M}\sum_{j=1}^M f(z_{ij}, x_i)}}, $$ where $\{(x_i,y_i)\}_{i=1}^N\sim p(x,y)$, and $\{z_{ij}\}_{j=1}^M\sim p(z|x_i)$ for each $x_i$. To ensure an excess risk of $\epsilon$, both $N$ and $M$ need be at least as large as $\Ocal(1/\epsilon^2)$, making the overall sample required to be $\Ocal(1/\epsilon^4)$; see~\cite{NemJudLanSha09,WanFanLiu14} and references therein. Hence, when $M$ is small, SAA would provide poor results. A second option is to resort to stochastic gradient methods (SGD). One can construct a \emph{biased} stochastic estimate of the gradient using $ {\nabla}_f L = \nabla \ell(y, \langle f,\tilde{\psi}(x)\rangle) \tilde{\psi}(x), $ where $\tilde{\psi}(x)$ is an estimate of $\EE_{z|x}[\psi(z,x)]$ for any $x$. To ensure convergence, the bias of the stochastic gradient must be small,~\ie, a large amount of samples from the conditional distribution is needed. Another commonly used approach is to first represent the conditional distributions as the so-called kernel conditional embedding, and then perform a supervised learning step on the embedded conditional distributions~\cite{SonFukGre13, GruLevBalPatetal12}. This two-step procedure suffers from poor statistical sample complexity and computational cost. The kernel conditional embedding estimation costs $O(N^3)$, where $N$ is number of pair of samples $(x, z)$. To achieve $\epsilon$ error in the conditional kernel embedding estimation, $N$ needs to be $\Ocal(1/\epsilon^4)$\footnote{With appropriate assumptions on joint distribution $p(x, z)$, a better rate can be obtained~\cite{GruLevBalPatetal12}. However, for fair comparison, we did not introduce such extra assumptions.}. Recently,~\citet{WanFanLiu14} solved a related but fundamentally distinct problem of the form, \begin{equation}\label{eq:WanFanLiu} \min_{f\in\Fcal}~L(f)=\EE_{y}\sbr{\ell(y, \EE_{z}[f(z)])} \end{equation} where $f(z)$ is a smooth function parameterized by some finite-dimensional parameter. The authors provide an algorithm that combines stochastic gradient descent with moving average estimation for the inner expectation, and achieves an overall $\Ocal(1/\epsilon^{3.5})$ sample complexity for smooth convex loss functions. The algorithm does not require the loss function to be convex, but it cannot directly handle random variable $z$ with \emph{infinite support}. Hence, such an algorithm does not apply to the more general and difficult situation that we consider in this paper. \paragraph{Our approach and contribution.} To address the above challenges, we propose a novel approach called \emph{dual kernel embedding}. The key idea is to reformulate (\ref{eq:target}) into a min-max or saddle point problem by utilizing the Fenchel duality of the loss function. We observe that with smooth loss function and continuous conditional distributions, the dual variables form a continuous function of $x$ and $y$. Therefore, we can parameterize it as a function in some RKHS induced by any universal kernel, where the information about the marginal distribution $p(x)$ and conditional distribution $p(z|x)$ can be aggregated via a kernel embedding of the joint distribution $p(x,z)$. Furthermore, we propose an efficient algorithm based on stochastic approximation to solve the resulting saddle point problem over RKHS spaces, and establish finite-sample analysis of the generic learning from conditional distributions problems. Compared to previous applicable approaches, an advantage of the proposed method is that it requires only \emph{one sample} from each conditional distribution. Under mild conditions, the overall sample complexity reduces to $\Ocal(1/\epsilon^2)$ in contrast to the $\Ocal(1/\epsilon^4)$ complexity required by SAA or kernel conditional embedding. As a by-product, even in the degenerate case (\ref{eq:WanFanLiu}), this implies an $\Ocal(1/\epsilon^2)$ sample complexity when inner function is linear, which already surpasses the result obtained in \cite{WanFanLiu14} and is known to be unimprovable. Furthermore, our algorithm is generic for the family of problems of learning from conditional distributions, and can be adapted to problems with different loss functions and hypothesis function spaces. Our proposed method also offers some new insights into several related applications. In reinforcement learning settings, our method provides the first algorithm that truly minimizes the mean-square Bellman error~(MSBE) with both theoretical guarantees and sample efficiency. We show that the existing gradient-TD2 algorithm by~\citet{SutMaePreBhaetal09,LiuLiuGhaMah15}, is a special case of our algorithm, and the residual gradient algorithm~\cite{Baird95} is derived by optimizing an upper bound of MSBE. In the invariance learning setting, our method also provides a unified view of several existing methods for encoding invariance. Finally, numerical experiments on both synthetic and real-world datasets show that our method can significantly improve over the previous state-of-the-art performances. \section{Preliminaries}\label{sec:preliminary} We first introduce our notations on Fenchel duality, kernel and kernel embedding. Let $\Xcal\subset \RR^d$ be some input space and $k:\Xcal\times\Xcal\to\RR$ be a positive definite kernel function. For notation simplicity, we denote the feature map of kernel $k$ or $\tilde k$ as $$ \phi(x):= k(x,\cdot),\quad \psi(z) : = \tilde k(z, \cdot), $$ and use $k(x,\cdot)$ and $\phi(x)$, or $\tilde k(z, \cdot)$ and $\psi(z)$ interchangeably. Then $k$ induces a RKHS $\Hcal$, which has the property $h(x) = \langle h,\phi(x)\rangle_{\Hcal}$, $\forall h\in \Hcal$, where $\langle\cdot,\cdot\rangle_{\Hcal}$ is the inner product and $\|h\|_\Hcal^2:=\langle h,h\rangle_\Hcal$ is the norm in $\Hcal$. We denote all continuous functions on $\Xcal$ as $\Ccal(\Xcal)$ and $\|\cdot\|_\infty$ as the maximum norm. We call $k$ a \emph{universal kernel} if $\Hcal$ is dense in $\Ccal(\Omega')$ for any compact set $\Omega'\subseteq\Xcal$,~\ie, for any $\epsilon>0$ and $u\in \Ccal(\Omega')$, there exists $h\in\Hcal$, such that $\|u-h\|_\infty\leq \epsilon$. Examples of universal kernel include the Gaussian kernel, $k(x, x') = \exp\rbr{-\frac{\|x-x'\|_2^2}{\sigma^{-2}}}$, Laplacian kernel, $k(x, x') =\exp\rbr{-\frac{\|x-x'\|_1}{\sigma^{-1}}}$, and so on. \paragraph{Convex conjugate and Fenchel duality.} Let $\ell : \RR^d\rightarrow \RR$, its convex conjugate function is defined as $$ \ell^*(u) = \sup_{v\in \RR^d}\{u^\top v - \ell(v)\}. $$ When $\ell$ is proper, convex and lower semicontinuous for any $u$, its conjugate function is also proper, convex and lower semicontinuous. More improtantly, the $(\ell, \ell^*)$ are dual to each other, \ie, $(\ell^*)^* = \ell$, which is known as Fenchel duality~\cite{HirLem12,RifLip07}. Therefore, we can represent the $\ell$ by its convex conjugate as , $$ \ell(v) = \sup_{u\in \RR^d}\{v^\top u - \ell^*(u)\}. $$ It can be shown that the supremum achieves if $v\in \partial \ell^*(u)$, or equivalently $u\in \partial \ell(v)$. \paragraph{Function approximation using RKHS.} Let $\Hcal^\delta:=\{h\in\Hcal:\|h\|_{\Hcal}^2\leq \delta\}$ be a bounded ball in the RKHS, and we define the approximation error of the RKHS $\Hcal^\delta$ as the error from approximating continuous functions in $\Ccal(\Xcal)$ by a function $h\in \Hcal^\delta$, \ie,~\cite{Bach14,Barron93} \begin{equation} \begin{array}{c} \Ecal(\delta):=\sup_{u\in\CC(\Xcal)}\inf_{h\in \Hcal^\delta}\|u-h\|_\infty. \end{array} \end{equation} One can immediately see that $\Ecal(\delta)$ decreases as $\delta$ increases and vanishes to zero as $\delta$ goes to infinity. If $\Ccal(\Xcal)$ is restricted to the set of uniformly bounded continuous functions, then $\Ecal(\delta)$ is also bounded. The approximation property,~\ie,~dependence on $\delta$ remains an open question for general RKHS, but has been carefully established for special kernels. For example, with the kernel $k(x,x') = 1/(1+\exp(\inner{x}{x'}))$ induced by the sigmoidal activation function, we have $\Ecal(\delta)=O(\delta^{-2/(d+1)}\log(\delta))$ for Lipschitz continuous function space $\Ccal(\Xcal)$~\cite{Bach14}.\footnote{The rate is also known to be unimprovable by~\citet{DeVHowMic89}.} \paragraph{Hilbert space embedding of distributions.} Hilbert space embeddings of distributions~\cite{SmoGreSonSch07} are mappings of distributions into potentially \emph{infinite} dimensional feature spaces, \begin{align} \mu_{x} \, := \, \EE_{x} \sbr{\phi(x)} \, = \, \int_{\Xcal} \phi(x) p(x) dx~:~ \Pcal \mapsto \Hcal \label{eq:embedding} \end{align} where the distribution is mapped to its expected feature map,~\ie,~to a point in a feature space. Kernel embedding of distributions has rich representational power. Some feature map can make the mapping injective~\cite{SriGreFukLanetal08}, meaning that if two distributions are different, they are mapped to two distinct points in the feature space. For instance, when $\Xcal\subseteq\RR^d$, the feature spaces of many commonly used kernels, such as the Gaussian RBF kernel, will generate injective embedding. We can also embed the joint distribution $p(x,y)$ over a pair of variables using two kernels $k(x,x) = \inner{\phi(x)}{\phi(x')}_{\Hcal}$ and $\tilde k(z,z') = \inner{\psi(z)}{\psi(z')}_{\Gcal}$ as \begin{eqnarray*} \Ccal_{zx} \, &:=& \, \EE_{zx} \sbr{\psi(z)\otimes \phi(x)} \\ &=& \int_{\Zcal\times\Xcal} \psi(z)\otimes \phi(x) p(z,x) dzdx : \Pcal \mapsto \Hcal\otimes\Gcal, \end{eqnarray*} where the joint distribution is mapped to a point in a tensor product feature space. Based on embedding of joint distributions, kernel embedding of conditional distributions can be defined as $\Ucal_{z|x}:=\Ccal_{zx}\Ccal_{xx}^{-1}$ as an operator $\Hcal \mapsto \Gcal$~\cite{SonFukGre13}. With $\Ucal_{z|x}$, we can obtain the expectations easily, \ie, \begin{eqnarray} \EE_{z|x}\sbr{g(z)} = \langle g, \langle \Ucal_{z|x}, \phi(x)\rangle_{\Hcal} \rangle_{\Gcal}. \end{eqnarray} Both the joint distribution embedding, $\Ccal_{zx}$, and the conditional distribution embedding, $\Ucal_{z|x}$, can be estimated from \iid~samples $\{(x_i, z_i)\}_{i=1}^N$ from $p(x, z)$ or $p(z|x)$, respectively~\cite{SmoGreSonSch07, SonFukGre13}, as \begin{eqnarray*} \widehat\Ccal_{zx} = \frac{1}{N}\Psi \Upsilon^\top, ~\text{and}~~~ \widehat \Ucal_{z|x} = \Psi(K + \lambda I)^{-1}\Upsilon^\top, \end{eqnarray*} where $\Psi = (\psi(z_1), \ldots, \psi(z_N))$, $\Upsilon = (\phi(x_1), \ldots, \phi(x_N))$, and $K = \Upsilon^\top \Upsilon$. Due to the inverse of $K + \lambda I$, the kernel conditional embedding estimation requires $\Ocal(N^3)$ cost. \section{Dual Embedding Framework} In this section, we propose a novel and sample-efficient framework to solve problem~\eq{eq:target}. Our framework leverages Fenchel duality and feature space embedding technique to bypass the difficulties of nested expectation and the need for overwhelmingly large sample from conditional distributions. We start by introducing the interchangeability principle, which plays a fundamental role in our method. \begin{lemma}[interchangeability principle]\label{lem:switch_correct} Let $\xi$ be a random variable on $\Xi$ and assume for any $\xi\in \Xi$, function $g(\cdot,\xi):\RR\to(-\infty,+\infty)$ is a proper\footnote{We say $g(\cdot, \xi)$ is proper when $\{u\in \RR: g(u, \xi)<\infty\}$ is non-empty and $g(u, \xi)>-\infty$ for $\forall u$.} and upper semicontinuous\footnote{We say $g(\cdot, \xi)$ is upper semicontinuous when $\{u\in \RR: g(u, \xi)<\alpha\}$ is an open set for $\forall \alpha\in\RR$. Similarly, we say $g(\cdot, \xi)$ is lower semicontinuous when $\{u\in \RR: g(u, \xi) >\alpha\}$ is an open set for $\forall\alpha\in \RR$.} concave function. Then \begin{equation*} \EE_{\xi}[\max_{u\in\RR}g(u,\xi)] =\max_{u(\cdot)\in \Gcal(\Xi)}\EE_{\xi}[g(u(\xi),\xi)]. \end{equation*} where $\Gcal(\Xi)=\{u(\cdot):\Xi\to\RR\}$ is the entire space of functions defined on support $\Xi$. \end{lemma} The result implies that one can replace the expected value of point-wise optima by the optimum value over a function space. For the proof of lemma~\ref{lem:switch_correct}, please refer to Appendix~\ref{appendix:dualcontinuity}. More general results of interchange between maximization and integration can be found in \cite[Chapter~14]{RocWet98} and \cite[Chapter~7]{ShaDen14}. \subsection{Saddle Point Reformulation} Let the loss function $\ell_y(\cdot):=\ell(y,\cdot)$ in \eq{eq:target} be a proper, convex and lower semicontinuous for any $y$. We denote $\ell_y^*(\cdot)$ as the convex conjugate; hence $\ell_y(v) = \max_{u}\{uv-\ell^*_y(u)\}$, which is also a proper, convex and lower semicontinuous function. Using the Fenchel duality, we can reformulate problem~\eq{eq:target} as \begin{eqnarray}\label{eq:dual_opt} \min_{f\in\Fcal} \EE_{xy}\bigg[\max_{u\in \RR} \Big[\EE_{z|x}[f(z,x)]\cdot u - \ell_y^*(u) \Big]\bigg], \end{eqnarray} Note that by the concavity and upper-semicontinuity of $-\ell_y^*(\cdot)$, for any given pair $(x,y)$, the corresponding maximizer of the inner function always exists. Based on the interchangeability principle stated in Lemma~\ref{lem:switch_correct}, we can further rewrite (\ref{eq:dual_opt}) as \begin{equation}\label{eq:dual_opt_exchange} \min_{f\in\Fcal} \max_{u(\cdot)\in \Gcal(\Xi)}\Phi(f,u):= \EE_{zx}[f(z,x)\cdot u(x,y)] - \EE_{xy}[\ell_y^*(u(x,y))], \end{equation} where $\Xi=\Xcal\times\Ycal$ and $\Gcal(\Xi)=\{u(\cdot):\Xi\to\RR\}$ is the entire function space on $\Xi$. We emphasize that the $\max$-operator in~\eq{eq:dual_opt} and~\eq{eq:dual_opt_exchange} have different meanings: the one in~\eq{eq:dual_opt} is taking over a single variable, while the other one in~\eq{eq:dual_opt_exchange} is over all possible function $u(\cdot)\in\Gcal(\Xi)$. Now that we have eliminated the nested expectation in the problem of interest, and converted it into a stochastic saddle point problem with an additional dual function space to optimize over. By definition, $\Phi(f,u)$ is always concave in $u$ for any fixed $f$. Since $f(z,x) = \inner{f}{\psi(z,x)}$, $\Phi(f,u)$ is also convex in $f$ for any fixed $u$. Our reformulation (\ref{eq:dual_opt_exchange}) is indeed a convex-concave saddle point problem. \begin{figure*}[!t] \centering \begin{tabular}{ccc} \includegraphics[width=0.315\textwidth, trim=1cm 0.8cm 1.6cm 1.6cm, clip]{figure/illustration_f_reverse_1.pdf}& \includegraphics[width=0.315\textwidth, trim=1cm 0.8cm 1.6cm 1.6cm, clip]{figure/illustration_f_reverse_50.pdf} & \includegraphics[width=0.315\textwidth, trim=1cm 0.8cm 1.6cm 1.6cm, clip]{figure/illustration_f_reverse_150.pdf} \\ (a) $0$-th Iteration &(b) $50$-th Iteration &(c) $150$-th Iteration \\ \includegraphics[width=0.315\textwidth, trim=1cm 0.8cm 1.6cm 1.6cm, clip]{figure/illustration_f_reverse_400.pdf} & \includegraphics[width=0.315\textwidth, trim=1cm 0.8cm 1.6cm 1.6cm, clip]{figure/illustration_f_reverse_1000.pdf} & \includegraphics[width=0.315\textwidth, trim=1cm 0.8cm 1.6cm 1.6cm, clip]{figure/illustration_f_reverse_2000.pdf} \\ (a) $400$-th Iteration &(b) $1000$-th Iteration &(c) $2000$-th Iteration \\ \end{tabular} \caption{Toy example with $f^*$ sampled from a Gaussian processes. The $y$ at position $x$ is obtained by smoothing $f^*$ with a Gaussian distribution condition on location $x$, \ie, $y = \EE_{z|x}\sbr{f^*(z)}$ where $z\sim p(z|x) = \Ncal\rbr{x, 0.3}$. Given samples $\{x, y\}$, the task is to recover $f^*(\cdot)$. The blue dash curve is the ground-truth $f^*(\cdot)$. The cyan curve is the observed noisy $y$. The red curve is the recovered signal $f(\cdot)$ and the green curve denotes the dual function $u(\cdot, y)$ with the observed $y$ plugged for each corresponding position $x$. Indeed, the dual function $u(\cdot, y)$ emphasizes the difference between $y$ and $\EE_{z|x}\sbr{f(z)}$ on every $x$. The interaction between primal $f(\cdot)$ and dual $u(\cdot, y)$ results in the recovery of the denoised signal.} \label{fig:procedure_illustration} \end{figure*} \paragraph{An example.} Let us illustrate this through a concrete example. Let $f^*(\cdot)\in\Fcal$ be the true function, and output $y = \EE_{z|x}\sbr{f^*(z)}$ given $x$. We can recover the true function $f^*(\cdot)$ by solving the optimization problem $$ \min_{f\in \Fcal}\EE_{xy}\sbr{\frac{1}{2}\rbr{y - \EE_{z|x}\sbr{f(z)}}^2}. $$ In this example, $\ell_y(v)=\frac{1}{2}(y-v)^2$ and $\ell_y^*(u)=uy+\frac{1}{2}u^2$. Invoking the saddle point reformulation, this leads to $$ \min_{f\in \Fcal}\max_{ u\in \Gcal(\Xi)}\EE_{xyz}\sbr{\rbr{f(z) - y}u(x,y)} - \frac{1}{2}\EE_{xy}\sbr{u(x,y)^2} $$ where the dual function $u(x,y)$ fits the discrepancy between $y$ and $\EE_{z|x}\sbr{f(z)}$, and thus, promotes the performance of primal function by emphasizing the different positions. See Figure~\ref{fig:procedure_illustration} for the illustration of the interaction between the primal and dual functions. \subsection{Dual Continuation} Although the reformulation in \eq{eq:dual_opt_exchange} gives us more structure of the problem, it is not yet tractable in general. This is because the dual function $u(\cdot)$ can be an arbitrary function which we do not know how to represent. In the following, we will introduce a tractable representation for (\ref{eq:dual_opt_exchange}). First, we will define the function $u^*(\cdot):\Xi=\Xcal\times\Ycal\to\RR$ as the \emph{optimal dual function} if for any pair $(x,y)\in\Xi$, $$u^*(x,y)\in\argmax\nolimits_{u\in\RR} \left\{u\cdot \EE_{z|x}[f(z,x)] - \ell_y^*(u)\right\}.$$ Note the optimal dual function is well-defined since the optimal set is nonempty. Furthermore, $u^*(x,y)$ is related to the conditional distribution via $u^*(x,y)\in \partial \ell_y(\EE_{z|x}[f(z,x)])$. This can be simply derived from convexity of loss function and Fenchel's inequality; see~\cite{HirLem12} for a more formal argument. Depending on the property of the loss function $\ell_y(v)$, we can further derive that (see proofs in Appendix~\ref{appendix:dualcontinuity}): \begin{proposition}\label{prop:dualcontinuity} Suppose both $f(z,x)$ and $p(z|x)$ are continuous in $x$ for any $z$, \begin{enumerate} \item[(1)] (Discrete case) If the loss function $\ell_y(v)$ is continuously differentiable in $v$ for any $y\in\Ycal$, then $u^*(x,y)$ is unique and continuous in $x$ for any $y\in \Ycal$; \item[(2)] (Continuous case) If the loss function $\ell_y(v)$ is continuously differentiable in $(v,y)$, then $u^*(x,y)$ is unique and continuous in $(x,y)$ on $\Xcal\times\Ycal$. \end{enumerate} \end{proposition} This assumption is satisfied widely in real-world applications. For instance, when it comes to the policy evaluation problem in~\ref{eq:RL_obj}, the corresponding optimal dual function is continuous as long as the reward function is continuous, which is true for many reinforcement tasks. The fact that the optimal dual function is a continuous function has interesting consequences. As we mentioned earlier, the space of dual functions can be arbitrary and difficult to represent. Now we can simply restrict the parametrization to the space of continuous functions, which is tractable and still contains the global optimum of the optimization problem in~\eq{eq:dual_opt_exchange}. This also provides us the basis for using an RKHS to approximate these dual functions, and simply optimizing over the RKHS. \subsection{Feature Space Embedding} In the rest of the paper, we assume conditions described in Proposition~\ref{prop:dualcontinuity} always hold. For the sake of simplicity, we focus only on the case when $\Ycal$ is a continuous set. Hence, from Proposition~\ref{prop:dualcontinuity}, the optimal dual function is indeed continuous in $(x,y)\in\Xi=\Xcal\times\Ycal$. As an immediate consequence, we lose nothing by restricting the dual function space $\Gcal(\Xi)$ to be continuous function space on $\Xi$. Recall that with the universal kernel, we can approximate any continuous function with arbitrarily small error. Thus we approximate the dual space $\Gcal(\Xi)$ by the bounded RKHS $\Hcal^{\delta}$ induced by a universal kernel $k((x,y), (x',y')) = \langle \phi(x,y), \phi(x',y')\rangle_{\Hcal}$ where $\phi(\cdot)$ is the implicit feature map. Therefore, $u(x,y) = \inner{u}{\phi(x,y)}_{\Hcal}$. Note that $\Hcal^\delta$ is a subspace of the continuous function space, and hence is a subspace of the dual space $\Gcal(\Xi)$. To distinguish inner product between the primal function space $\Fcal$ and the dual RKHS $\Hcal^\delta$, we denote the inner product in $\Fcal$ as $\langle \cdot, \cdot\rangle_{\Fcal}$. We can rewrite the saddle point problem in~\eq{eq:dual_opt_exchange} as \begin{align}\label{eq:dual_approximate} \min_{f \in\Fcal} \max_{u\in \Hcal^\delta} \Phi(f,u) &= \EE_{xyz}\sbr{\langle f,\psi(z,x)\rangle_{\Fcal}\cdot\langle u, \phi(x,y)\rangle_{\Hcal} \hspace{-1mm}-\hspace{-1mm} \ell_y^*(\langle u, \phi(x,y)\rangle_{\Hcal})} \nonumber\\ &= f^\top \Ccal_{zxy} u - \EE_{xy}[\ell_y^*(\langle u, \phi(x,y)\rangle_{\Hcal})], \end{align} where $f(z, x) = \langle f, \psi(z, x)\rangle_{\Fcal}$ by the definition of $\Fcal$, and $\Ccal_{zxy} = \EE_{zxy}[\psi(z,x)\otimes\phi(x,y)]$ is the joint embedding of $p(z,x, y)$ over $\Fcal\times \Hcal$. The new saddle point approximation (\ref{eq:dual_approximate}) based on dual kernel embedding allows us to efficient represent the dual function and get away from the fundamental difficulty with insufficient sampling from the conditional distribution. There is no need to access either the conditional distribution $p(z|x)$, the conditional expectation $\EE_{z|x}\sbr{\cdot}$, or the conditional embedding operator $\Ucal_{z|x}$ anymore, therefore, reducing both the statistical and computational complexity. Specifically, given a pair of sample $(x,y,z)$, where $(x,y)\sim p(x,y)$ and $z\sim p(z|x)$, we can now easily construct an unbiased stochastic estimate for the gradient, namely, \begin{eqnarray*} \nabla_{{f}}\hat{\Phi}_{x,y,z}({f},u)&=&\psi(z,x) u(x,y),\\ \nabla_{u}\hat{\Phi}_{x,y,z}({f},u)&=&[f(z,x)-\nabla \ell_y^*(u(x,y))] \phi(x,y), \end{eqnarray*} with $\EE\sbr{\nabla\hat{\Phi}_{x,y,z}({f},u)}=\nabla\Phi(f, u)$, respectively. For simplicity of notation, we use $\nabla$ to denote the subgradient as well as the gradient. With the unbiased stochastic gradient, we are now able to solve the approximation problem (\ref{eq:dual_approximate}) by resorting to the powerful mirror descent stochastic approximation framework~\cite{NemJudLanSha09}. \subsection{Sample-Efficient Algorithm} The algorithm is summarized in Algorithm~\ref{alg:stochastic_composite_general}. At each iteration, the algorithm performs a projected gradient step both for the primal variable $f$ and dual variable $u$ based on the unbiased stochastic gradient. The proposed algorithm avoids the need for overwhelmingly large sample sizes from the conditional distributions when estimating the gradient. At each iteration, only one sample from the conditional distribution is required in our algorithm! Throughout our discussion, we make the following standard assumptions: \begin{assumption}\label{asp:bounded} There exists constant scalars $C_\Fcal$, $M_\Fcal$, and $c_\ell$, such that for any ${f}\in{\Fcal},u\in\Hcal^\delta$, \begin{eqnarray*} \EE_{z,x}[\|f(z,x)\|_2^2]\leq M_\Fcal, \quad\EE_{z,x}[\|\psi(z,x)\|_\Fcal^2]\leq C_\Fcal, \quad \EE_{y}[\|\nabla \ell^*_y(u)\|_2^2]\leq c_\ell. \end{eqnarray*} \end{assumption} \begin{assumption}\label{asp:kernel} There exists constant $\kappa>0$ such that $k(w,w')\leq \kappa$ for any $w,w'\in\Xcal$. \end{assumption} Assumption~\ref{asp:bounded} and~\ref{asp:kernel} basically suggest that the variance of our stochastic gradient estimate is always bounded. Note that we do not assume any strongly convexity or concavity of the saddle point problem, or Lipschitz smoothness. Hence, we set the output as the average of intermediate solutions weighted by the learning rates $\{\gamma_i\}$, as often used in the literature, to ensure the convergence of the algorithm. \begin{algorithm}[t!] \caption{\textbf{Embedding-SGD} for Optimization~(\ref{eq:dual_approximate})} \text{\bf Input:} $p(x,y),\, p(z|x),\, \psi(z,x),\,\phi(x,y),\, \{\gamma_i\geq 0\}_{i=1}^t$\\[-4mm] \begin{algorithmic}[1]\label{alg:stochastic_composite_general} \FOR{$i=1,\ldots, t$} \STATE Sample $(x_i,y_i) \sim p(x,y)$ and $z_i \sim p(z|x)$. \STATE ${f}_{i+1} = \Pi_{\Fcal}({f}_i-\gamma_i\psi(z_i,x_i)u_i(x_i,y_i))$. \STATE $u_{i+1} = \Pi_{\Hcal^\delta}(u_i+\gamma_i [f_i(z_i,x_i)-\nabla \ell_{y_i}^*(u_i(x_i,y_i))]\phi(x_i,y_i))$ \ENDFOR\\ \text{\bf Output:} $\bar{f}_t=\frac{\sum_{i=1}^t\gamma_i{f}_i}{\sum_{i=1}^t\gamma_i}$, $\bar u_t=\frac{\sum_{i=1}^t \gamma_iu_i}{\sum_{i=1}^t \gamma_i}$ \end{algorithmic} \end{algorithm} Define the accuracy of any candidate solution $(\bar f,\bar u)$ to the saddle point problem as \begin{equation}\label{eq:sadaccuracy} \epsilon_{\rm gap}(\bar f,\bar u):=\max_{u\in \Hcal^\delta}\Phi(\bar f,u)-\min_{f\in\Fcal}\Phi(f,\bar u). \end{equation} We have the following convergence result, \begin{theorem} \label{thm:main} Under Assumptions~\ref{asp:bounded} and~\ref{asp:kernel}, the solution $(\bar{f}_t,\bar u_t)$ after $t$ steps of the algorithm with step-sizes being $\gamma_t=\frac{\gamma}{\sqrt{t}} (\gamma>0)$ satisfies: \begin{equation}\label{eq:mainbound} \EE[\epsilon_{\rm gap} (\bar{f}_t,\bar u_t)]\leq [(2D_{\Fcal}^2+4\delta)/\gamma+\gamma \Ccal(\delta,\kappa) ]\frac{1}{\sqrt{t}} \end{equation} where $D_{\Fcal}^2=\sup_{{f}\in{\Fcal}}\frac{1}{2}\|{f}_0-{f}\|_2^2$ and $\Ccal(\delta,\kappa)= \kappa(5M_\Fcal+c_\ell)+\frac{1}{8}(\delta+\kappa)^2C_\Fcal$. \end{theorem} The above theorem implies that our algorithm achieves an overall $\Ocal(1/\sqrt{t})$ convergence rate, which is known to be unimprovable already for traditional stochastic optimization with general convex loss function \cite{NemJudLanSha09}. We further observe that \begin{proposition} \label{prop:Lipschitzianity} If $f(z,x)$ is uniformly bounded by $C$ and $\ell_y^*(v)$ is uniformly $K$-Lipschitz continuous in $v$ for any $y$, then $\Phi(f,u)$ is $(C+K)$-Lipschitz continuous on $\Gcal(\Xi)$ with respect to $\|\cdot\|_\infty$, i.e. $$|\Phi(f,u_1)-\Phi(f,u_2)|\leq (C+K)\|u_1-u_2\|_\infty, \forall u_1,u_2\in\Gcal(\Xi).$$ \end{proposition} Let ${f}_*$ be the optimal solution to (\ref{eq:target}). Invoking the Lipschitz continuity of $\Phi$ and using standard arguments of decomposing the objective, we have $$L(\bar {f}_t)-L({f}_*)\leq \epsilon_{\rm gap}(\bar {f}_t,\bar u_t)+2(C+K)\Ecal(\delta).$$ Combining Proposition~\ref{prop:Lipschitzianity} and Theorem~\ref{thm:main}, we finally conclude that under the conditions therein, \begin{equation}\label{eq:combinedbound} \EE[L(\bar{f}_t)-L({f}_*)]\leq \Ocal\left(\frac{\delta^{3/2}}{\sqrt{t}} +\Ecal(\delta)\right). \end{equation} There is clearly a delicate trade-off between the optimization error and approximation error. Using large $\delta$ will increase the optimization error but decrease the approximation error. When $\delta$ is moderately large (which is expected in the situation when the optimal dual function has small magnitude), our dual kernel embedding algorithm can achieve an overall $\Ocal(1/\epsilon^2)$ sample complexity when solving learning problems in the form of (\ref{eq:target}). For the analysis details, please refer to Appendix~\ref{appendix:convergece_rate}. \section{Applications}\label{sec:application} In this section, we discuss in details how the dual kernel embedding can be applied to solve several important learning problems in machine learning, \eg, learning with invariance and reinforcement learning, which are the special cases of the optimization~\eq{eq:target}. {By simple verification, these examples satisify our assumptions for the convergence of algorithm.} We tailor the proposed algorithm for the respective learning scenarios and unify several existing algorithms for each learning problem into our framework. Due to the space limit, we only focus on algorithms with kernel embedding. Extended algorithms with random feature, doubly SGD, neural networks as well as their hybrids can be found in Appendix~\ref{appendix:dual_random_fea},~\ref{appendix:doublySGD} and~\ref{appendix:dual_neural_networks}. \subsection{Learning with Invariant Representations} {\bf Invariance learning.} The goal is to solve the optimization (\ref{eq:invariant}), which learns a function in RKHS $\tilde\Hcal$ with kernel $\tilde k$. Applying the dual kernel embedding, we end up solving the saddle point problem \begin{eqnarray*} \min_{f\in\tilde{\Hcal}}\max_{u\in \Hcal}~\EE_{zxy}\sbr{\langle f, \psi(z)\rangle_{\tilde\Hcal} \cdot u(x,y)}- \EE_{xy}[\ell_y^*(u(x,y))]+ \frac{\nu}{2}\|f\|^2_{\tilde{\Hcal}}, \end{eqnarray*} where $\Hcal$ is the dual RKHS with the universal kernel introduced in our method. \noindent{\bf Remark.} The proposed algorithm bears some similarities to virtual sample techniques~\cite{NiyGirPog98, LooCanBot07} in the sense that they both create examples with prior knowledge to incorporate invariance. In fact, the virtual sample technique can be viewed as optimizing an upper bound of the objective~\eq{eq:invariant} by simply moving the conditional expectation outside, \ie, $\EE_{x, y} [\ell(y, \EE_{z|x}[f(z)] )] \le \EE_{x, y, z}\big[\ell(y,f(z))\big]$, where the inequality comes from convexity of $\ell(y, \cdot)$. \noindent{\bf Remark.} The learning problem~\eq{eq:invariant} can be understood as learning with RKHS $\hat \Hcal$ with Haar-Integral kernel $\hat k$ which is generated by $\tilde k$ as $\hat k(x, x') =\langle \EE_{p(z|x)}[\psi(z)], \EE_{p(z'|x')}[\psi(z')]\rangle_{\tilde \Hcal}$, with implicit feature map $\EE_{p(z|x)}[\psi(z)]$. If $f\in \tilde \Hcal$, then, $ f(x) = \EE_{z|x}[\langle f, \psi(z)\rangle_{\tilde\Hcal}] = \langle f, \EE_{z|x}[\psi(z)]\rangle \in \hat \Hcal$. The Haar-Integral kernel can be viewed as a special case of Hilbertian metric on probability measures on which the output of function should be invariant~\cite{HeiBou05}. Therefore, other kernels defined for distributions, \eg, the probability product kernel~\cite{JebKonHow04}, can be also used in incorporating invariance. \noindent{\bf Remark.} Robust learning with contamined samples can also be viewed as incorporating invariance prior with respect to the perturbation distribution into learning procedure. Therefore, rather than resorting to robust optimization techniques~\cite{BhaPanSmo05,BenGhaNem08}, the proposed algorithm for learning with invariance serves as a viable alternative for robust learning. \subsection{Reinforcement Learning} {\bf Policy evaluation.} The goal is to estimate the value function $V^\pi(\cdot)$ of a given policy $\pi(a|s)$ by minimizing the mean-square Bellman error (MSBE)~\eq{eq:RL_obj}. With $V^\pi\in \tilde\Hcal$ with feature map $\psi(\cdot)$, we apply the dual kernel embedding, which will lead to the saddle point problem \begin{eqnarray}\label{eq:RL_dual} \min_{V^\pi\in\tilde{\Hcal}} \max_{u\in\Hcal} &&\EE_{s',a, s} \sbr{\rbr{R(s,a) - \langle V^\pi, \psi(s) -\gamma \psi(s') \rangle_{\tilde \Hcal}} u(s, a)} -\frac{1}{2}\EE_{s, a}[u^2(s, a)]. \end{eqnarray} In the optimization~\eq{eq:RL_dual}, we simplify the dual $u$ to be function over $\Scal \times \Acal$ due to the fact that $R(s, a)$ is determinastic given $s$ and $a$ in our setting. If $R(s, a)$ is a random variable sampled from some distribution $p(R|s, a)$, then the dual function should be defined over $\Scal \times \Acal \times \RR$. \noindent{\bf Remark.} The algorithm can be extended to off-policy setting. Let $\pi_b$ be the behavior policy and $\rho(a|s) = \frac{\pi(a|s)}{\pi_b(a|s)}$ be the importance weight, then the objctive will be adjusted by $\rho(a|s)$, \ie \begin{eqnarray*} \min_{V^\pi\in \tilde \Hcal}\EE_{s, a}\sbr{\rbr{\EE_{s'|a, s}\sbr{\rho(a|s)\rbr{R(s, a) - \langle V^\pi, \psi(s) -\gamma \psi(s')\rangle_{\tilde \Hcal} }}}^2} \end{eqnarray*} where the successor state $s'\sim P(s'|s, a)$ and actions $a\sim\pi_b(a|s)$ from behavior policy. With the dual kernel embedding, we can derive similar algorithm for off-policy setting, with extra importance weight $\rho(a|s)$ to adjust the sample distribution. \noindent{\bf Remark.} We used different RKHSs for primal and dual functions. If we use the \emph{same finite basis functions} to parametrize both the value function and the dual function, \ie, $V^\pi(s)=\theta^T\psi(s)$ and $u(s)=\eta^T\psi(s)$, where $\psi(s) = [\psi_i(z)]_{i=1}^d\in \RR^{d}$, $\theta, \eta\in \RR^d$, our saddle point problem \eq{eq:RL_dual} reduces to $ \min_{\theta} \big\|\EE_{s, a, s'}[\Delta_\theta(s, a, s')\psi]\big\|^2_{\EE[\psi\psi^\top]^{-1}}, $ where $\Delta_\theta(s, a, s') = R(s, a) + \gamma V^\pi(s') - V^\pi(s)$. This is exactly the same as the objective proposed in~\cite{SutMaePreBhaetal09} of gradient-TD2. Moreover, the update rules in gradient-TD2 can also be derived by conducting the proposed Embedding-SGD with such parametrization. For details of the derivation, please refer to Appendix~\ref{appendix:GTD_special}. From this perspective, gradient-TD2 is simply a special case of the proposed Embedding-SGD applied to policy evaluation with particular parametrization. However, in the view of our framework, there is really no need to, and should not, restrict to the same finite parametric model for the value and dual functions. As further demonstrated in our experiments, with different nonparametric models, the performances can be improved significantly. See details in Section~\ref{subsec:RL_exp}. The residual gradient~(RG)~\cite{Baird95} is trying to apply stochastic gradient descent directly to the MSBE with finite parametric form of value function, \ie, $V^\pi(s)=\theta^T\psi(s)$, resulting the gradient as $ \EE_{s, a, s'}\sbr{\Delta_\theta(s, a, s')\psi(s)} - \gamma\EE_s\sbr{\EE_{s', a|s}\sbr{\Delta_\theta(s, a, s')}\EE_{s'|s}\sbr{\psi(s')}}. $ Due to the inside conditional expectation in gradient expression, to obtain an unbiased estimator of the gradient, it requires two independent samples of $s'$ given $s$, which is not practical. To avoid such ``double sample'' problem, \citet{Baird95} suggests to use gradient as $\EE_{s, a, s'}\sbr{\Delta_\theta(s, a, s')\rbr{\psi(s) - \gamma\psi(s')}}$. In fact, such algorithm is actually optimizing $\EE_{s, a, s'}\sbr{\Delta_\theta(s, a, s')^2}$, which is an upper bound of MSBE~\eq{eq:RL_obj} because of the convexity of square loss. Our algorithm is also fundamentally different from the TD algorithm even in the finite state case. The TD algorithm updates the state-value function directly by an estimate of the temporal difference based on one pair of samples, while our algorithm updates the state-value function based on accumulated estimate of the temporal difference, which intuitively is more robust. \noindent{\bf Optimal control.} The goal is to estiamte the $z(\cdot)$ function by minimizing the error in linear Bellamn equation~\eq{eq:OC_obj}. With $z\in \tilde\Hcal$ with feature map $\psi(\cdot)$, we apply the dual kernel embedding, which will lead to the saddle point problem \begin{eqnarray}\label{eq:OC_dual} \min_{z \in\tilde{\Hcal}} \max_{u\in\Hcal} \Phi(z,u):=\EE_{s', s} \sbr{\langle z, \psi(s) - \exp(-R(s))\psi(s') \rangle_{\tilde \Hcal}\cdot u(s)} -\frac{1}{2}\EE_{s}[u^2(s)]. \end{eqnarray} With the learned $z^*$, we can recover the optimal control via its conditional distribution $p^{\pi_*}(s'|s) = \frac{p(s'|s)z^*(s)}{\EE_{s'|s\sim p(s'|s)}\sbr{z^*(s')}}$. \subsection{Events Prediction in Stochastic Processes} {\bf Expected Hitting Time.} The goal is to estimate the expected hitting time $h(\cdot, \cdot)$ which minimizes the recursive error~\eq{eq:HT_obj}. With $h\in \tilde \Hcal$ with feature map $\psi(\cdot, \cdot)$, we apply the dual kernel embedding similarly to the policy evaluation, which will lead to the saddle point problem \begin{eqnarray}\label{eq:HT_dual} \min_{h \in\tilde{\Hcal}} \max_{u\in\Hcal} \Phi(h,u):=\EE_{s, t}\EE_{s'|s, t} \sbr{ \rbr{1 - \langle h, \psi(s, t) - \psi(s', t) \rangle_{\tilde\Hcal} } u(s, t)}- \frac{1}{2}\EE_{s}\sbr{u(s, t)^2}. \end{eqnarray} \noindent{\bf Remark.} With the proposed algorithm, we can estimate $h(s, t)$, the expected hitting time starting from state $s$ and hitting state $t$, even without observing the hitting events actually happen. We only need to collect the trajectory of the stochastic processes, or just the one-step transition, to feed to the algorithm, therefore, utilize the data more efficietly. \section{Experiments} We test the proposed algorithm for two applications, \ie, learning with invariant representation and policy evaluation. For full details of our experimental setups, please refer to Appendix~\ref{appendix:experiment_setup}. \subsection{Experiments on Invariance Learning} To justify the algorithm for learning with invariance, we test the algorithm on two tasks. We first apply the algorithm to robust learning problem where the inputs are contaminated, and then, we conduct comparison on molecular eneretics prediction problem~\cite{MonHanFazRupetal12}. We compare the proposed algorithm with SGD with virtual samples technique~\cite{NiyGirPog98,LooCanBot07} and SGD with finite sample average for inner expectation~(SGD-SAA). We use Gaussian kernel in all tasks. To demonstrate the sample-efficiency of our algorithm, $10$ virtual samples are generated for each datum in training phase. The algorithms are terminated after going through $10$ rounds of training data. We emphasize that the SGD with virtual samples is optimizing an upper bound of the objective, dan thus, it is predictable that our algorithm can achieve better performance. We plot its result with dot line instead. \paragraph{\bf Noisy measurement.} We generate a synthetic dataset by \begin{eqnarray*} \bar x &\sim&\Ucal([-0.5, 0.5]),\quad x = \bar x + 0.05e,\\ y&=& (\sin(3.53\pi \bar x) + \cos(7.7 \pi \bar x))\exp(-1.6\pi|\bar x|) + 3 \bar x^2+ 0.01e, \end{eqnarray*} where the contamination $e\sim \Ncal(0, 1)$. Only $(x, y)$ are provided to learning methods, while $\bar x$ is unknown. The virtual samples are sampled from $z\sim \Ncal(x, 0.05^2)$ for each observation. The 10 runs average results are illustrated in Figure~\ref{fig:learning_invariant}(a). The proposed algorithm achieves average MSE as low as $0.0029$ after visit $0.1$M data, significantly better than the alternatives. \begin{figure*}[!t] \centering \begin{tabular}{cc} \includegraphics[width=0.35\textwidth]{figure/synthetic_invariant_kernel_best-crop.pdf}&~~~~~~~~~~~~ \includegraphics[width=0.35\textwidth]{figure/quantummachine-crop.pdf}\\ (a) Robust Learning &~~~~~~~~~~~~(b) QuantumMachine\\ \end{tabular} \caption{Learning with invariance.} \label{fig:learning_invariant} \end{figure*} \paragraph{\bf QuantumMachine.} We test the proposed algorithm for learning with invariance task on QuantumMachine 5-fold dataset for atomization energy prediction. We follow~\cite{MonHanFazRupetal12} that the data points are represented by Coulomb matrices, and the virtual samples are generated by random permutation. The average results are shown in Figure~\ref{fig:learning_invariant}(b). The proposed algorithm achieves a significant better solution, while SGD-SAA and SGD with virtual samples stuck in inferior solutions due to the inaccurate inner expectation estimation and optimizing indirect objective, respectively. \subsection{Experiments on Policy Evaluation}\label{subsec:RL_exp} We compare the proposed algorithm to several prevailing algorithms for policy evaluation, including gradient-TD2~(GTD2)~\cite{SutMaePreBhaetal09,LiuLiuGhaMah15}, residual gradient~(RG)~\cite{Baird95} and kernel MDP~\cite{GruLevBalPonetal12} in terms of mean square Bellman error~\cite{DanGerPet14}. It should point out that kernel MDP is not an online algorithm, since it requires to visit the entire dataset when estimating the embedding and inner expectation in each iteration. We conduct experiments for policy evaluation on several benchmark datasets, including navigation, cart-pole swing up and PUMA-560 manipulation. We use Gaussian kernel in the nonparametric algorithms, \ie, kernel MDP and Embedding SGD, while we test random Fourier features~\cite{RahRec08} for the parametric competitors, \ie, GTD2 and RG. In order to demonstrate the sample efficiency of our method, we only use one sample from the conditional distribution in the training phase, therefore, cross-validation based on Bellman error is not appropriate. We perform a parameter sweep to select the hyper-parameters as~\cite{SilLevHeeetal14}. See appendix \ref{appendix:experiment_setup} for detailed settings. Results are averaged over 10 independent trials. \begin{figure*}[!t] \centering \begin{tabular}{ccc} \includegraphics[width=0.312\textwidth]{figure/kernel_synthetic_best_backup-crop.pdf}& \includegraphics[width=0.312\textwidth]{figure/cp_best_backup-crop.pdf}& \includegraphics[width=0.312\textwidth]{figure/puma_best_backup-crop.pdf} \\ (a) Navigation & (b) Cart-Pole &(c) PUMA-560\\ \end{tabular} \caption{Policy evaluation.} \label{fig:policy_evaluation} \end{figure*} \paragraph{\bf Navigation.} The navigation in an unbounded room task extending the discretized MDP in~\cite{GruLevBalPonetal12} to continuous-state continuous-action MDP. Specifically, the reward is $R(s) = \exp(-100\|s\|^2)$ centered in the middle of the room and $s\sim \Ncal(0, 0.2I)$. We evaluate the deterministic policy policy $\pi(s) = -0.2s R(s)$, following the gradient of the reward function. The transition distribution follows Gaussian distribution, $ p(s'|a, s) = \Ncal(s+a, 0.1I)$. Results are reported in Figure~\ref{fig:policy_evaluation}(a). \paragraph{\bf Cart-pole swing up.} The cart-pole system consists of a cart and a pendulum. It is an under-actuated system with only one control act on the cart. The goal is to swing-up the pendulum from the initial position (point down). The reward will be maximum if the pendulum is swing up to $\pi$ angle with zero velocity. We evaluate the linear policy $\pi(s) = As+b$ where $A\in \mathbb{R}^{1\times 4}$ and $b\in\mathbb{R}^{1\times 1}$. Results are reported in Figure~\ref{fig:policy_evaluation}(b). \paragraph{\bf PUMA-560 manipulation.} PUMA-560 is a robotic arm that has 6 degrees of freedom with 6 actuators on each joint. The task is to steer the end-effector to the desired position and orientation with zero velocity. The reward function is maximum if the arm is located to the desired position. We evaluate the linear policy $\pi(s) = {A}s+{b}$ where ${A}\in \mathbb{R}^{6\times 12}$ and ${b}\in\mathbb{R}^{6\times 1}$. Results are reported in Figure~\ref{fig:policy_evaluation}(c). In all experiments, the proposed algorithm performs consistently better than the competitors. The advantages of proposed algorithm mainly come from three aspects: {\bf i)}, it utilizes more flexible dual function space, rather than the constrained space in GTD2; {\bf ii)}, it directly optimizes the MSBE, rather than its surrogate as in GTD2 and RG; {\bf iii)}, it directly targets on value function estimation and forms an one-shot algorithm, rather than a two-stage procedure in kernel MDP including estimating conditional kernel embedding as an intermediate step. \section{Conclusion}\label{sec:conclusion} We propose a novel \emph{sample-efficient} algorithm, {\bf Embedding-SGD}, for addressing learning from conditional distribution problems. Our algorithm benefits from a fresh employ of saddle point and kernel embedding techniques, to mitigate the difficulty with limited samples from conditional distribution as well as the presence of nested expectations. To our best knowledge, among all existing algorithms able to solve such problems, this is \emph{the first} algorithm that allows to take only one sample at a time from the conditional distribution and comes with provable theoretical guarantee. We apply the proposed algorithm to solve two fundamental problems in machine learning, \ie, learning with invariance and policy evaluation in reinforcement learning. The proposed algorithm achieves the state-of-the-art performances on these two tasks comparing to the existing algorithms. As we discussed in Appendix~\ref{appendix:control}, the algorithm is also applicable to control problem in reinforcement learning. In addition to its wide applicability, our algorithm is also very versatile and amenable for all kinds of enhancement. The algorithm can be easily extended with random feature approximation or doubly stochastic gradient trick. Moreover, we can extend the framework by learning complicated nonlinear feature jointly with the function, which results the dual neural network embedding in Appendix~\ref{appendix:dual_neural_networks}. It should be emphasized that since the primal and dual function spaces are designed for different purposes, although we use both RKHS in main text for simplicity, we can also use different function approximators separately for primal and dual functions. \bibliographystyle{plainnat} {
a73ab6b9fd2c2765c7eed77e02add5b7446c498b
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The general idea of controlling transport is associated with the concept of implementing actions to ensure that a system exhibits desired characteristics. For example, Nature is well known in presenting sophisticated mechanisms that regulate phenomena which take place in all scales of time and space \cite{haw2007}. A careful and systematic exploration of these mechanism reveals that sometimes it is more reliable to control the setup by letting it to fluctuate and change its dynamics as little as possible to drive it to the desired state without applying intense external forces. A typical scenario Nature uses is the resource to tailored non-equilibrium forces, as it is the case for molecular and Brownian motors \cite{astumian2002,hanggi2009,spiechowicz2013jstatmech,spiechowicz2014pre, spiechowicz2016jstatmech}. Brownian machinery presents an archetypal scheme in which, even in the absence of an externally applied bias, directed motion emerges from {\it unbiased} environmental non-equilibrium perturbations via the mechanism of breaking the spatial symmetry of the setup \cite{hanggi2009}. The origin of directed transport stems from the fact that the combined action of non-equilibrium perturbations of stochastic or deterministic nature takes the system out of thermal equilibrium and breaks the detailed balance symmetry. This working principle can be seen as a key for understanding various processes occurring in numerous disciplines of science, including e.g. intracellular transport \cite{bressloff2013}. In view of the widespread applications of Brownian motors, directed motion controllability has become a focal point of research in non-equilibrium statistical physics which inspired a plethora of new microscale devices displaying unusual transport features \cite{vlassiouk2007, serreli2007, mahmud2009, costache2010, drexler2013, spiechowicz2014prb, spiechowicz2015njp, spiechowicz2015chaos, grossert2016}. Irregular behaviour generally implies also unpredictability, which, when it comes to applications one typically attempts to avoid or at least to minimize. A particular aspect of unpredictability is the appearance of diffusive spreading constituting yet another salient transport quantifier. Beyond doubt, the problem of diffusion has played a central role in the development of both the foundations of thermodynamics and statistical physics \cite{hanggi2005} and has led to numerous applications within a diversity of scientific disciplines. Inspired by the physics of Brownian motors, we ask to what extent it is possible to influence the diffusive behaviour of the system in presence of thermal noise and deterministic external perturbations and elucidate the mechanisms behind such characteristics. The conventional diffusive spreading characterized by the diffusion coefficient $D$ increases with the temperature of the medium surrounding the system \cite{sutherland1905,einstein1905,smoluchowski1906}. Here, we find that the diffusion coefficient $D$ assumes a non-monotonic behaviour as a function of temperature, possessing a bell-shaped maximum at some intermediate temperature. This form of peculiar diffusive behaviour is markedly distinct from the known case of anomalous diffusion for which the mean square displacement of the particle $\langle \Delta x^2(t) \rangle$ grows asymptotically according to a power law $\langle \Delta x^2(t) \rangle \sim t^\alpha$ with $\alpha \ne 1$ \cite{metzler2014,zaburdaev2015}. The studied non-monotonic diffusion here must also be distinguished from the phenomenon of giant diffusion. The latter has been observed in the diffusive motion of a Brownian particle in a tilted periodic landscape \cite{lindner2001, reimann2001, reimann2002, heinsalu2004, lindner2016}, in disordered potentials \cite{dan2002, garcia2014} or for an adiabatically driven inertial Brownian particle moving in a periodic one-dimensional geometry \cite{marchenko2012}. To investigate our objective of a non-monotonic dependence of diffusion on temperature we consider the non-equilibrium dynamics of a massive Brownian particle moving in a periodic potential under the influence of a time-periodic force. The class of models described by such a dynamics includes periodically driven pendulums \cite{gitterman2010}, super-ionic conductors \cite{fulde1975}, Josephson junctions \cite{kautz1996}, dipoles rotating in external fields \cite{coffey2012}, phase-locked loops \cite{viterbi1966}, dislocations in solid state physics \cite{seeger1980}, solitons described by the Sine-Gordon equation \cite{lamb1980}, the Frenkel-Kontorova lattices \cite{braun1998}, dynamics of adatoms subjected to a time-periodic force \cite{guantes2001}, charge density waves \cite{gruner1981} and cold atoms in optical lattices \cite{denisov2014}, to name but a few. The rich physics contained in this model has become evident over the last decades with numerous studies. In particular, with present periodic driving this class of systems comprises operational regimes that are deterministically chaotic. The sensitive dependence on initial conditions and the abundance of unstable periodic attractors are the most salient characteristics of chaotic behaviour \cite{strogatz}. The combination of these features, possibly assisted in addition with noise driving, make chaotic systems one of the most flexible setups enabling the emergence of the discussed peculiar diffusive behaviour. \section{Model} We formulate the problem in terms of a classical particle of mass $M$ which is (i) subjected to a \emph{spatially periodic} potential $U(x) = U(x + L)$ of period $L$ possessing the \emph{mirror symmetry}, i.e. there exists $x_0$ such that the relation $U(x_0-x) = U(x_0+x)$ holds for all $x$, (ii) additionally being agitated by an external \emph{unbiased} time-periodic deterministic force $A\cos{(\Omega t)}$ with angular frequency $\Omega$ and of amplitude strength $A$, and (iii) coupled to a thermal bath at temperature $\theta$. The overall dynamics of such a Brownian particle is determined by the following Langevin equation \begin{equation} \label{eq:model} M\ddot{x} + \Gamma\dot{x} = -U'(x) + A\cos{(\Omega t)} + \sqrt{2\Gamma k_B \theta}\,\xi(t). \end{equation} Here, the dot and the prime denote differentiation with respect to the time $t$ and the particle coordinate $x$, respectively. Thermal fluctuations due to the coupling of the particle with the thermal bath are modelled by $\delta$-correlated Gaussian white noise $\xi(t)$ of zero mean and unit intensity, i.e., \begin{equation} \langle \xi(t) \rangle = 0, \quad \langle \xi(t)\xi(s) \rangle = \delta(t-s). \label{xi} \end{equation} The parameter $\Gamma$ denotes the friction coefficient and $k_B$ is the Boltzmann constant. The noise intensity factor $2\Gamma k_B \theta$ with temperature $\theta$ of the heat bath follows from the fluctuation-dissipation theorem \cite{HT1982}. The latter ensures the stationarity of the thermal canonical Gibbs state for vanishing driving under the dynamics governed by Eq. (\ref{eq:model}), i.e., when $A = 0$ and in presence of periodic boundary conditions. We focus on the archetypical non-linear situation, namely, a sinusoidal potential with a barrier height $2\Delta U$, \begin{equation} U(x) = \Delta U \sin{\left( \frac{2\pi x}{L} \right)}. \end{equation} Upon introducing the period $L$ and the parameter combination $\tau_0 = \Gamma L^2/\Delta U$ as units for length and time, respectively, Eq. (\ref{eq:model}) is written in a dimensionless form \cite{machura2008}, reading, \begin{equation} \label{eq:dimlessmodel} m\ddot{\hat{x}} + \dot{\hat{x}} = -\hat{U}'(\hat{x}) + a\cos{(\omega \hat{t})} + \sqrt{2Q}\,\hat{\xi}(\hat{t}), \end{equation} where $\hat{x} = x/L$ and $\hat{t} = t/\tau_0$. The dimensionless potential $\hat{U}(\hat{x}) = U(x)/\Delta U = U(L\hat{x})/\Delta U = \hat{U}(\hat{x} + 1)$ assumes the unit period $L = 1$ and the potential amplitude $\Delta \hat{U} = 1$. The remaining re-scaled parameters are the mass $m = (1/\Gamma\tau_0)M$, the amplitude $a = (L/\Delta U)A$ and the angular frequency $\omega = \tau_0\Omega$. The dimensionless thermal noise reads $\hat{\xi}(\hat{t}) = (L/\Delta U)\xi(t) = (L/\Delta U)\xi(\tau_0\hat{t})$ and assumes the same statistical properties, namely it is Gaussian with $\langle \hat{\xi}(\hat{t}) \rangle = 0$ and \mbox{$\langle \hat{\xi}(\hat{t})\hat{\xi}(\hat{s}) \rangle = \delta(\hat{t} - \hat{s})$}. The dimensionless noise intensity $Q = k_B\theta/\Delta U$ is given by the ratio of thermal energy $k_{B}\theta$ and half of the activation energy the particle needs to overcome the original potential barrier $2 \Delta U$. From now on, we will use only the dimensionless variables and therefore shall skip the "hat" for all quantities appearing in Eq. (\ref{eq:dimlessmodel}). The observable of foremost interest in this study is the diffusion coefficient $D$ which characterizes the spread of trajectories and fluctuations around the average position of the particle, namely \cite{spiechowicz2015pre, spiechowicz2016scirep}, \begin{equation} \label{eq:dc} D = \lim_{t \to \infty} \frac{\langle [ x(t) - \langle x (t) \rangle ]^2 \rangle}{2t}, \end{equation} where the averaging $\langle \cdot \rangle$ is over all realizations of thermal fluctuations as well as over initial conditions for the position $x(0)$ and the velocity $\dot{x}(0)$. The latter is necessary because in the deterministic limit of vanishing thermal noise intensity $Q \to 0$ the dynamics may possess several coexisting attractors thus being non-ergodic and implying that the corresponding results may be affected by a specific choice of those selected initial conditions \cite{spiechowicz2016scirep,kostur2008}. In order to explain the discussed peculiar behaviour of the diffusion coefficient we will also investigate velocity of the Brownian particle. Due to the presence of the external time-periodic driving $a\cos{(\omega t)}$ as well as the friction term $\dot{x}$ in Eq. (\ref{eq:dimlessmodel}) the particle velocity $\dot{x}(t)$ approaches a unique asymptotic state in which it is characterized by a temporally periodic probability density. This latter density function has the same period $T$ as the driving \cite{jung1990, jung1993}. In particular, the first statistical moment of the instantaneous particle velocity $\langle \dot{x}(t) \rangle$ assumes for a long time the form \begin{equation} \lim_{t \to \infty} \langle \dot{x}(t) \rangle = \langle \mathbf{v} \rangle + v(t), \end{equation} where $\langle \mathbf{v} \rangle$ is the directed transport velocity. Note that all three terms entering the right hand side of Eq. (\ref{eq:dimlessmodel}) are \emph{unbiased}: the average of the potential force over a spatial period $L$ vanishes as well as that of the time-dependent driving over a temporal period $T$, and also the average of the random force $\xi(t)$ vanishes according to Eq. (\ref{xi}). Moreover, all three elements are symmetric and in consequence there is no directed transport in the long time stationary regime, i.e. $\langle \mathbf{v} \rangle =0$. The deviation $v(t)$ is periodic with period $T$ and has a vanishing time average, i.e. $v(t+T) = v(t)$ and $(1/T) \int_0^T dt\, v(t)= 0$. Therefore it is useful to consider also the {\it period averaged} velocity $\mathbf{v}(t)$ as \begin{equation} \mathbf{v}(t) = \frac{1}{T} \int_{t}^{t + T} ds\, \dot{x}(s), \label{v} \end{equation} where $T = 2\pi/\omega$ denotes a period of the external driving $a\cos{(\omega t)}$. The particle positions $x(n T)$ at integer multiples $n$ of the period $T$ can be exactly expressed by the sequence of period averaged velocities $\mathbf{v}(k T) = (1/T)\int_{kT}^{(k+1)T} ds\,\dot{x}(s)$ defined in Eq. (\ref{v}) as \begin{equation} x(n T) = T \sum_{k=0}^{n-1} \mathbf{v}(k T)\:, \label{xnvk} \end{equation} where we assumed that the particle starts at $x(0)=0$ (for $x(0)\neq 0$ one can re-define $x(nT) \to \delta x(nT) =x(nT)-x(0)$). The period averaged velocities at multiple integers of the period $T$ have a vanishing mean value. Therefore the average position remains zero and the variance of the position is given by its second moment reading \begin{equation} \label{xx} \langle x^2(n T) \rangle = T^2 \sum_{k,l}^{n-1} \langle \mathbf{v}(kT)\mathbf{v}(lT) \rangle. \end{equation} In the asymptotic limit of large times and with the temporally periodic probability density the velocity auto-correlation function depends only on the time difference. In view of Eq. (\ref{xx}) the left hand side approaches $\langle x^2(nT)\rangle = 2nTD$ and hence the diffusion coefficient D can be expressed in terms of the temporal period averaged velocity autocorrelation function, yielding \begin{equation} D = T \sum_{k=0}^\infty \langle \mathbf{v}(kT) \mathbf{v}(0) \rangle. \label{Dvv} \end{equation} An equivalent expression is given in \cite{MKMTHL}. The infinite time limit in Eq. (\ref{eq:dc}) implies that the auto-correlation function has to be determined for the stationary ensemble of the period averaged velocities. Because neither the Langevin equation (\ref{eq:dimlessmodel}) nor the corresponding Fokker-Planck variant can be solved within analytical means we performed comprehensive numerical simulations of the model. We did so by employing a weak version of the stochastic second-order predictor-corrector algorithm with a time step typically set to about $(10^{-3} - 10^{-2}) \times T$ \cite{spiechowicz2015cpc}. Because \mbox{Eq. (\ref{eq:dimlessmodel})} is a second-order differential equation, we need to specify two initial conditions, namely $x(0)$ and $\dot{x}(0)$. We choose $x(0)$ and $\dot{x}(0)$ to be equally distributed over the intervals $[0, 1]$ and $[-2,2]$, respectively. Our quantities of interest were averaged over $10^3 - 10^5$ sample trajectories. All numerical calculations have been performed by use of a CUDA environment as implemented on a modern desktop GPU. This procedure did allow for a speedup of a factor of the order $10^3$ times as compared to a common present-day CPU method \cite{spiechowicz2015cpc}. \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{fig1a} \caption{The dependence of the diffusion coefficient $D$ on the noise intensity $Q \propto \theta$ being proportional to the temperature $\theta$ of the heat-bath. The other parameters are fixed to the following values $m = 0.9, a = 8.7, \omega = 0.275$. The diffusion constant was determined from Eq. (\ref{eq:dc}) on the basis of ensembles of $10^5$ trajectories of Eq. (\ref{eq:dimlessmodel}) reaching up to $t=10^4\,T$ for different values of $Q$.} \label{fig1} \end{figure} \section{Results} The dynamical system described by Eq. (\ref{eq:dimlessmodel}) exhibits an extremely rich behaviour as a function of the four dimensionless parameters $\{m, a, \omega, Q\}$ with dimensionless friction at $\Gamma=1$. A general overview is provided, e.g. with Ref. \cite{kautz1996}. Our objective is not the systematic exploration of the setup at hand but rather the search for the above mentioned {\it non-monotonic behaviour of the diffusion constant} with increasing noise intensity as shown in Fig. \ref{fig1}. The corresponding parameters of the dynamical system (\ref{eq:dimlessmodel}) are chosen as \mbox{$\{m = 0.9, a = 8.7, \omega = 0.275\}$}. At each depicted noise strength $Q \propto \theta$ the diffusion constant $D$ was estimated from $10^5$ trajectories by means of Eq. (\ref{eq:dc}). At low noise intensity $D$ increases with $Q$ until it reaches a local maximum at $Q\approx 2 \cdot 10^{-5}$. From there it decreases to a minimum at $Q\approx 5\cdot 10^{-3}$ turning over to a monotonic function of $Q$, and, finally, at sufficiently large values of $Q$ becoming strictly proportional to $Q$, i.e. to the temperature $\theta$ of the ambient thermal bath. This high temperature behaviour, however, is not depicted in Fig. \ref{fig1}. The decrease of the diffusion constant with increasing temperature $Q \propto \theta$ is counter-intuitive. It stays in clear contrast with the Einstein relation $D \propto \theta$ as well as with other known formulas, e.g. Vogel-Fulcher-like laws \cite{goychuk2014} or Arrhenius-type behaviour for the diffusion of a Brownian particle in periodic potentials \cite{lifsonjackson,festa1978,htb1990}. \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{fig2a} \includegraphics[width=0.49\linewidth]{nfig2b} \caption{The deterministic dynamics in dependence of the driving amplitude $a$ is illustrated as bifurcation diagrams (a) for the stroboscopic position $x(kT)$ projected onto the principal interval $[0,L=1]$ and (b) the period averaged velocity $v(kT)$ defined in Eq. (\ref{v}) both presented after $k = 10^4$ driving cycles. Chosen parameters are the same as in Fig. \ref{fig1} except now the system is deterministic $Q = 0$. The arrows mark specific driving strengths for which the dynamics changes its character (see in text).} \label{fig2} \end{figure} \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{nfig3a} \includegraphics[width=0.49\linewidth]{nfig3b}\\ \includegraphics[width=0.49\linewidth]{nfig3c} \caption{An illustrative trajectory of the period averaged velocity $\mathbf{v}(kT)$, see Eq. (\ref{v}), is presented for the deterministic motion at $Q=0$ in panel (a) and for two different temperatures corresponding to maximal (panel (b), $Q = 2.16 \cdot 10^{-5}$) and minimal (panel (c), $Q = 0.00525$) diffusion coefficient, c.f. Fig. \ref{fig1}. Trajectories are depicted for $k=200$ driving cycles. Each blue dot represents the velocity $\mathbf{v}(kT)$ for the period $k$ at the corresponding abscissa. The remaining parameters are the same as those in Fig. \ref{fig1}. In all three cases the velocity stays predominantly close to $\mathbf{v}(kT) = \pm 0.4$ and $\mathbf{v}(kT) = 0$. In the majority of cases a trajectory does not stay for a longer period of time close to one of the dominant velocities but rather jumps in the vicinity of another one, without following any obvious rule.} \label{fig3} \end{figure} Because relevant structural elements of the stochastic dynamics described by Eq. (\ref{eq:dimlessmodel}) are determined by its deterministic properties as the first step of our analysis we consider the noiseless case. \subsection{Noiseless dynamics: $Q=0$} For the set of parameters presented in Fig. 1, the deterministic system (when $Q=0$) exhibits chaotic behaviour with a dense set of unstable periodic orbits. For $Q>0$, the system 'feels' some unstable orbits which play an important role in controlling diffusion properties. Below we explain this. For this aim, in Fig. \ref{fig2} we present bifurcation diagrams for the positions $x(k T)$ (panel (a)) and the period averaged velocities $\mathbf{v}(kT)$ (panel (b)) as a function of the driving amplitude $a$ of the periodic driving $a\cos{(\omega t)}$ for the deterministic system (\ref{eq:dimlessmodel}) with $Q = 0$. For any given value of the driving amplitude $a$ the red dots in panel (a) represent the stroboscopic positions of the particle $x(kT)\, \mathrm{mod}\, L$, i.e. the projected values onto the spatial principal period $[0,1]$, for $1024$ trajectories with $k=10^4$. The velocities displayed in panel (b) show the according values $\mathbf{v}(kT)$ obtained from Eq. (\ref{v}). The bifurcation diagrams reveal three qualitatively different dynamical regimes. -- For the smallest displayed values of $a<a_0 =8.6867 $ as well as for $a>a_c=8.6893$ the motion is chaotic. The stroboscopic positions projected onto the principal spatial period almost densely cover the available interval. In both regions the transport is diffusive \cite{inoue1982}, i.e. the diffusion constant determined by Eq. (\ref{eq:dc}) has a finite value larger than zero. In the window between $a_0$ and $a_c$ the motion is phase locked with the corresponding winding number $w = \pm 9$: the particle proceeds within one temporal period by 9 spatial periods either in positive or in negative direction depending on the initial condition. This is reflected by the two velocities $\mathbf{v}(kT) \approx \pm 0.4$ visible in panel (b) in the corresponding parameter window. In panel (a) two period doubling cascades start at the parameter value $a_1=8.6878$ and terminate at $a_\infty=8.6886$. In the small parameter interval between $a_\infty$ and $a_c$ the asymptotic motion is phase locked taking place on two coexisting chaotic attractors. The window of phase locked motion ends at $a_c$ with an attractor merging crisis \cite{ott} giving rise to a diffusive attractor covering the full principal period. Also a wide range of velocities suddenly emerges at $a_c$ and continues to be present at larger values of the parameter $a$. However, one still detects a strong concentration of velocities at several values of $\mathbf{v}(kT)$ corresponding to integer winding numbers characterizing unstable periodic orbits. Particularly large is the probability to find temporal period averaged velocities $\mathbf{v}(kT) \approx \pm 0.4$, corresponding to the winding numbers $w = \pm 9$. These two, together with the locked trajectories with $w = 0 $ and a few other phase locked trajectories seem to constitute the backbone of unstable periodic orbits supporting the chaotic motion. As a simplified picture of the chaotic dynamics one may think of a process in which these unstable orbits are visited in a random sequence. \subsection{Influence of thermal noise: $Q>0$} In Fig. \ref{fig3} we present the influence of thermal noise on velocity of the Brownian particle. We depict time series of period averaged velocities $\mathbf{v}(kT)$ resulting from a single trajectory in the large time limit for the same set of parameters specified in Fig. \ref{fig1}. Panel (a) exemplifies the deterministic case, with $Q=0$. Regions close to the dominant winding numbers $w = \pm 9$ and $w = 0$ are significantly more frequently visited than others. Yet, the trajectory almost never dwells near any of these states for a longer period of time but typically leaves the state within the period after it had arrived there. Panels (b) and (c) illustrate the influence of noise on the period averaged velocity time series for the two values $Q =2.16\cdot 10^{-5}$ and $Q=5.25 \cdot 10^{-3}$ which correspond to the positions of the maximum and the minimum of the diffusion constant $D$ displayed in \mbox{Fig. \ref{fig1}}, respectively. At the smaller noise strength displayed in panel (b) still some of the fine details of the deterministic time series are visible while others are washed out. At the larger noise intensity (panel (c)) all fine details have disappeared apart from the fact that positive and negative velocities near $\mathbf{v}(kT) \approx \pm 0.4$ still occur with relatively high probability as well as small velocities $\mathbf{v}(kT) \approx 0$. \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{fig4} \caption{The probabilities $p_R$ and $p_L$ for the particle to be in the running $|\mathbf{v}(kT)| \approx 0.4$ and in the locked $\mathbf{v}(kT) \approx 0$ state, respectively, are plotted against temperature of the system $Q \propto \theta$. The diffusion coefficient $D$ is maximal when the difference between probabilities $p_R - p_L$ has a peak, c.f. Fig. \ref{fig1}. When $p_R$ and $p_L$ intersect the diffusion coefficient $D$ is minimal. Other parameters are the same as in Fig. \ref{fig1}.} \label{fig4} \end{figure} By introducing the thresholds at $\mathbf{v}(kT) = \pm 0.2$ and counting the number of period averaged velocities which are between and outside these values one may estimate the probabilities $p_R = \mathrm{prob}(|\mathbf{v}(kT)| > 0.2)$ and $p_L = \mathrm{prob}(|\mathbf{v}(kT)| \leq 0.2)$ for the occurrence of the running and locked states, respectively. These probabilities are displayed in Fig. \ref{fig4} as a function of the noise intensity $Q$. We note that the running states occur significantly more frequent at weaker noise than at larger noise. Considering as a very rough model an independent sequence $\{v_k\}$ of velocities $\mathbf{v}(kT) \approx \pm 0.4, 0$ occurring with the respective probabilities $p_R, p_L$ one finds a diffusive behaviour for the spreading of the positions $x(n T) = \sum_{k=0}^{n-1} \mathbf{v}(k T)$ with the diffusion constant $D= 0.16 T p_R$. With the observed noise-dependence of the probability $p_R$ one already finds a qualitative agreement with the non-monotonic behaviour of the diffusion constant displayed in Fig. \ref{fig1}. However, in the next Section we will present a slightly more realistic model based on the same three velocity states but with a more adequate dynamics. \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{kqr} \caption{The probabilities $k$ (red) to stay in the locked state and $q$ (blue) to stay in either of the running states, and the transition probability $r$ (green) to switch from a running into the locked state are displayed as functions of the noise strength $Q$. While the latter probability changes only little in the considered region of noise intensities the transition probability $k$ increases approximately in the same range of noise strengths where $q$ decreases.} \label{fig6} \end{figure} \section{Approximate three-state model description of chaos induced non-monotonic diffusion} In order to describe the non-monotonic behaviour of the diffusion constant $D$ in dependence of the noise strength $Q$ we consider a simplified model only, containing transitions between the most relevant unstable orbits. As such we consider the phase locked trajectories with winding numbers $w = \pm 9$ corresponding to the velocities $v_{\pm} \approx \pm 0.4$ and those with winding number $w = 0$ corresponding to $v_0 = 0$. The model is assumed to be symmetric with respect to the two running states $v_\pm$. The sequence of period averaged velocities $\mathbf{v}(nT)$ is then modelled by a Markov process in terms of transitions between these states. Because of the assumed symmetry the matrix $\mathbf{M}$ describing these transitions has the form \begin{equation} \mathbf{M}:= \left (\begin{array}{ccc} q & \frac{1}{2}(1-k)& r\\ 1-q-r &k & 1-q-r\\ r & \frac{1}{2}(1-k) & q \end{array} \right )\:, \label{M} \end{equation} where $q$ denotes the conditional probability to remain staying in the running state $v_+$, $v_+ \rightarrow v_+$, and likewise in $v_-$, $v_- \rightarrow v_-$; further $k$, the conditional probability to remain staying in the resting state $0$, $0 \rightarrow 0$; and $r$, the conditional probability of a transition between opposite running states $v_+ \rightarrow v_-$ and $v_- \rightarrow v_+$. Due to the assumed symmetry of the running states the transitions from the locked into a running state ($0 \rightarrow v_\pm$) is $(1-k)/2$ and from a running into the locked state ($v_\pm \rightarrow 0$) is $1-q-r$. Probabilities $q$ and $k$ may take non-negative values less than or equal to 1, while $r$ is restricted to $0\leq r \leq 1-q$. The transition probabilities can be estimated from simulations of the Langevin equation (\ref{eq:dimlessmodel}) as the relative frequencies with which the period averaged velocity $v = \mathbf{v}(nT)$ visits the three coarse grained regions $V_+=\{v| v\geq 0.2\}$, $V_0=\{v|-0.2 < |v| < 0.2\}$ and $V_-=\{v| v\leq -0.2\}$. The three independent rates $k$, $q$, $r$ are displayed in Fig. \ref{fig6}. Finally, in the present model the states are ``decorated'' with the respective velocities $\{v_\pm,0\}$ such that the particle changes its position within a period by $v T$ with $v\in \{v_\pm,0\}$. \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{pstat1} \caption{The directly estimated probability $p_R/2$ (red) of a particle to be found in the running state as a function of the noise strength $Q$ is compared with the respective stationary probability (\ref{pst}) (blue) following from the three-state model defined by the matrix (\ref{M}) of transition probabilities. The agreement is excellent for a not too large noise intensity. For larger noise intensity the agreement is still qualitatively good.} \label{fig7} \end{figure} Collecting the probabilities of finding the particle at the time $t = nT$ in either of the three states as a vector, $\mathbf{p}(n) = (p_+(n),p_0(n),p_-(n))$ one can write its dynamics as \begin{equation} \mathbf{p}(n+1) = \mathbf{M} p(n) \label{ME} \end{equation} having the formal solution \begin{equation} \mathbf{p}(n) = \mathbf{M}^n \mathbf{p}(0) \label{pn} \end{equation} with the initial probability $\mathbf{p}(0)$. The stationary distribution $\mathbf{p}^{\mathrm{st}}$ is invariant under transitions, and hence is the solution of \begin{equation} \mathbf{p}^{\mathrm{st}} = \mathbf{M} \mathbf{p}^{\mathrm{st}} \label{Mp} \end{equation} given by \begin{equation} \mathbf{p}^{\mathrm{st}} = \left ( \begin{array}{c} \frac{1-k}{2(2-k-q-r)}\\ \frac{1-q-r}{2-k-q-r}\\ \frac{1-k}{2(2-k-q-r)} \end{array} \right )\:. \label{pst} \end{equation} In Fig. \ref{fig7} the stationary probability $p^{\mathrm{st}}_+$ to find the particle in the running state is compared with $p_R/2$ estimated from simulations of the Langevin equation (\ref{eq:dimlessmodel}). The diffusion constant can be determined in the framework of the Markovian three-state model by means of Eq. (\ref{Dvv}) which is based on the stationary auto-correlation function of period averaged velocities. This correlation function follows from the three-state model as \begin{equation} \langle \mathbf{v}(n T)\mathbf{v}(0) \rangle = \sum_{\alpha,\beta} v_\alpha v_\beta (\mathbf{M}^n)_{\alpha,\beta} p^{\mathrm{st}}_\beta\:, \label{vv} \end{equation} where $\alpha$ and $\beta$ label the three velocity states and the matrix element $ (\mathbf{M}^n)_{\alpha,\beta}$ determines the conditional probability to find the state labelled by $\alpha$ after $n$ periods provided the system has started in the state $\beta$. Accordingly, $p^{\mathrm{st}}_\beta$ denotes the probability (\ref{pst}) to find $\beta$ in the stationary state. In order to evaluate the $n^{\mathrm{th}}$ power of the matrix $\mathbf{M}$ it is convenient to use its spectral representation reading \begin{equation} \mathbf{M} = \sum_{\kappa=1}^3 \lambda_\kappa \Pi^{(\kappa)}\:, \label{spM} \end{equation} where $\lambda_\kappa$ are the eigenvalues of $\mathbf{M}$ and $\Pi^{(\kappa)}$ are matrices projecting onto the respective eigenvectors. The eigenvalue $\lambda_1 = 1$ results as a consequence of the conservation of total probability. The corresponding matrix $\Pi^{1}$ projects any vector $\mathbf{q}$ onto the stationary probability $\mathbf{p}^{\mathrm{st}}$, i.e. it acts as $\Pi^{1}\mathbf{q} = \mathbf{p}^{\mathrm{st}} \sum_\alpha q_\alpha$. Explicit expressions of the other eigenvalues and eigenprojectors are given in the Appendix \ref{APP}. Inserting the spectral representation (\ref{spM}) into the expression (\ref{vv}) one finds that the first eigenvalue $\lambda_1$ does not contribute to the auto-correlation function because it yields a factor that is proportional to the equilibrium expectation value of the velocity and hence vanishes. Therefore the period averaged velocity auto-correlation function simplifies to read \begin{eqnarray} \label{vvsp} \langle \mathbf{v}(n T)\mathbf{v}(0) \rangle &= \sum_{\kappa=2}^3 \lambda_\kappa^n \sum_{\alpha,\beta} v_\alpha v_\beta \Pi^{(\kappa)}_{\alpha,\beta} p^{\mathrm{st}}_\beta \nonumber \\ &=2 v^2 \lambda^n_2 p^{\mathrm{st}}_1 \;, \end{eqnarray} where we inserted the velocities $v_1=v$, $v_2=0$ and $v_3=-v$ and made use of the explicit form of the projection matrices $\Pi^{(\kappa)}$ as given in the \ref{APP}. Note that it does not depend on $\lambda^n_3$. This is a consequence of symmetry of the matrix $M$ and the resulting symmetry of the eigen-projector $\Pi^{(3)}$, see equation (A6). These symmetries reflect the symmetry of the model with respect to two states with non-zero velocities $v_+$ and $v_-$. \begin{figure}[t] \centering \includegraphics[width=0.49\linewidth]{DQ} \caption{The diffusion coefficient $D$ (red) obtained from simulation of the Langevin equation (\ref{eq:dimlessmodel}) is displayed as a function of the noise strength $Q$ and compared with the respective diffusion constant (blue) given by Eq. (\ref{D3S}) resulting from the three state model. The two curves qualitatively agree with each other in displaying the same bell-shape with matching extrema, however a quantitative discrepancy by a factor of roughly 1.5 is apparent. Possible reasons for this discrepancy are explained in the Discussion.} \label{fig8} \end{figure} The sum on $n$ is readily performed to yield for the diffusion matrix \begin{eqnarray} D&= 2 v^2 T\frac{p^{\mathrm{st}}_1}{1-\lambda_2} \nonumber \\ &= v^2 T \frac{1-k}{(2-k-q-r)(1-q+r)} \;, \label{D3S} \end{eqnarray} where we used the Eqs. (\ref{pst}) and (\ref{lambda}). In Fig. \ref{fig8} the resulting expression is compared with the diffusion constant obtained from simulations of the Langevin equation (\ref{eq:dimlessmodel}). It turns out that (\ref{D3S}) reproduces the non-monotonic behaviour of the diffusion constant with its maximum and minimum in good agreement with the simulation result. However, the absolute magnitude of $D$ resulting from (\ref{D3S}) is too large by approximately a factor of 1.5. \section{Discussion} In this work we studied the spreading of a cloud of inertial, massive Brownian particles independently moving in a periodic potential. As Brownian particles they are subject to a friction as well as to a fluctuating force, both being imposed by the interaction with a thermal heat bath held at a temperature $\theta$. Moreover, they are driven by a force periodically varying in time. In the present investigation we considered a parameter regime in which the system behaves chaotic in the absence of thermal noise, i.e., if formally the temperature is set to zero. If the temperature is gradually increased the diffusion constant characterizing the spreading rate of the particle cloud first increases until it reaches a local maximum and then decreases again until a minimum is reached. A further increase of the temperature causes a continuous growth of the diffusion constant which eventually becomes proportional to temperature. As the reason for the counter-intuitive non-monotonic behaviour of the diffusion coefficient versus temperature we identified the dynamics of the populations of certain regions in phase space. The latter contain certain deterministically unstable periodic orbits. Unstable periodic orbits are known to constitute the backbone of deterministic chaos \cite{ACEGP}. They allow one to reconstruct the chaotic dynamics of a system in a hierarchical way. Using a simplified model we considered three unstable periodic orbits. Two of them move with opposite temporal period averaged velocities and a third one that is locked. The frequencies of transitions change with temperature, thereby leading to a non-trivial dependence of the diffusion coefficient on temperature. Considering Markovian transitions between these states with transition rates estimated from numerical simulations of the Langevin equation (\ref{eq:dimlessmodel}) then leads to a qualitatively agreeing description, however, with a too large diffusion coefficient in the considered range of temperatures. This discrepancy may have different reasons. First of all, non-Markovian memory effects might lead to a smaller diffusion constant. Additionally, a more systematic partitioning of the phase space \cite{badii,BFBS} and the inclusion of further unstable periodic orbits likely will improve the agreement. Here, however, we have limited ourselves to the generic simplest stochastic model with three states in discrete time. The presented analysis has been restricted to one set of the system parameters. It is not the only one and exceptional region where a non-monotonic dependence of the diffusion coefficient $D$ on temperature $\theta$ is detected. We have found other regimes but are not presented here because the temperature dependence of $D$ and its mechanisms are similar. Non-monotonic diffusion was also recently found in a rocked ratchet \cite{spiechowicz2015pre}; i.e. a driven Brownian particle moving in a periodic potential lacking any mirror symmetry. Due to the absence of the latter property a non-trivial directed transport effect $\langle \mathbf{v} \rangle \neq 0$ may emerge in this system. Contrary to the case considered in this paper, the deterministic limit of the mentioned regime is non-chaotic and possesses three stable coexisting states. The normal diffusion of the particle is caused by thermal noise produced by a heat bath in the small-to-moderate temperature limit. In the limit of weak noise a three-state modeling along the reasoning put forward here might as well provide satisfactory insight into the diffusion behavior there. In contrast, the present setup exhibits chaotic motion in the deterministic limit leading already to a finite diffusion constant $D \neq 0$ at $\theta=0$. At low temperatures the combination of chaotic motion with very weak noise leads to an increase of the diffusion. With increasing noise the conditional probability to stay in the locked state eventually increases whereas that of the running state decreases going hand in hand with a decrease of the diffusion constant. For this particular mechanism to work all terms of the Langevin equation (\ref{eq:dimlessmodel}) are relevant. Without the inertial term, i.e. in the overdamped limit, the dynamics take place in a two-dimensional state space and therefore cannot display chaos. Likewise, the absence of the periodic driving also would restrict the system to a two dimensional phase space without the option of exhibiting chaos. Moreover, the periodic forcing drives the system out of thermal equilibrium in which the diffusion constant is a strictly monotonic function of temperature. For the case with vanishing friction the model becomes simple; then a dose of finite noise corresponds to infinite temperature rendering any bounded potential ineffective. In this sense with our setup in Eq. (\ref{eq:dimlessmodel}) we deal with a minimal model for the peculiar diffusive behaviour. Since the latter is readily implemented in many diverse physical situations \cite{gitterman2010,fulde1975,kautz1996,coffey2012,viterbi1966,seeger1980,lamb1980,braun1998,guantes2001,gruner1981,denisov2014} we expect that a similar non-monotonic behaviour of the diffusion may be observed in various contexts recurring in the most diverse areas of science such as physics, chemistry, biology, engineering, computer science or even sociology. We expect that our findings can be experimentally corroborated with any of the physical systems mentioned in the introductory part of the article. As a possibility we propose a promising setup for this purpose, namely the dynamics of cold atoms dwelling optical lattices \cite{lutz2013}. These systems are known for their high tunability, providing a precise control of the amplitude and period of a defect-free symmetric, spatially periodic optical potential. We consider the simplest and the most common model of a dissipative optical lattice consisting of atoms of mass $m$ with a two level structure illuminated by the optical molasses, i.e. by counter-propagating light fields with the same frequency tuned slightly below the electronic transition of the atoms. This trap generates the potential force $-U'(x)$ with precisely adjusted period and amplitude. The red-detuning of the laser fields leads to the Doppler cooling mechanism \cite{metcalf} which can be described as a classical damping term $\propto -\dot{x}$. The random photon absorption and re-emission events can be modelled by the Gaussian white noise $\xi(t)$ with noise intensity $Q$. Thus, the light beams play a role of the bath at an effective temperature to which atoms are coupled to \cite{kindermann2016}. Temperature of the atoms is mainly controlled by the molasses frequency or by heating of the optical lattice due to the noisy electronic phase shifting of the beam, see Ref. \cite{kindermann2016} and Eq. (6) therein. Finally, in order to generate the external driving a time dependent phase modulation $\varphi(t) = a\cos{(\omega t)}$ can be applied to one of the lattice building beams. In the laboratory reference frame such a laser configuration generates a moving optical potential. In the co-moving frame the potential is static and the atoms experience an inertial force which is proportional to the time dependent phase modulation $\varphi(t) = a\cos{(\omega t)}$ \cite{renzoni2003, wickenbrock2012}. In this way, the predicted non-monotonic behaviour of the diffusion constant as a function of temperature should manifest itself in an optical lattice experiment. Last but not least, we stress that the detected peculiar diffusive behaviour manifesting in non-monotonic temperature dependence of the diffusion coefficient can be observed in \emph{deep} optical lattices. This non-intuitive diffusion behaviour is therefore quite distinct from the well explored regime with shallow potentials where typically anomalous diffusion processes occur \cite{lutz2003, lutz2004}. \section*{Acknowledgements} This work was supported in part by the MNiSW program via the Diamond Grant (J.S.) and the NCN grant 2015/19/B/ST2/02856 (J. {\L}). P.T. thanks the Foundation for Polish Science (FNP) for granting him an Alexander von Humboldt Honorary Research Fellowship.
71c21d304dc6074d16886fdee8ad0932e97a6723
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:intro} Galaxy clusters form from large primordial density fluctuations that have collapsed and virialised by the present epoch, with more massive clusters forming from larger and rarer fluctuations. This makes them especially sensitive to fundamental cosmological parameters, such as the matter density, the amplitude of the matter power spectrum and the equation of state of dark energy \citep[see][]{Voit2005,AllenEvrardMantz2011,KravtsovBorgani2012,Weinberg2013}. The observable properties of a galaxy cluster result from a non-trivial interplay between gravitational collapse and astrophysical processes. The diverse range of formation histories of the cluster population leads to scatter in the observable-mass scaling relations and, as surveys select clusters based on an observable, this can lead to a biased sample of clusters, resulting in systematics when using them as a cosmological probe \citep[e.g.][]{Mantz2010}. Many previous studies have shown that the relationship between a cluster observable, such as its temperature or X-ray luminosity, and a quantity of interest for cosmology, e.g. its mass, has a smaller scatter for more massive, dynamically relaxed objects \citep{EkeNavarroFrenk1998,Kay2004,Crain2007,NagaiVikhlininKravtsov2007a,Planelles2013}. Therefore, the fundamental requirement when probing cosmological parameters with galaxy clusters is a sample of relaxed, massive clusters with well calibrated mass-observable scaling relations. However, galaxy clusters are rare objects, becoming increasingly rare with increasing mass, and to observe a sample large enough to be representative of the underlying population requires a survey with significant size and depth. Currently ongoing and impending observational campaigns, such as the Dark Energy Survey \citep{DEScol2005}, \textit{eRosita} \citep{Merloni2012}, \textit{Euclid} \citep{Laureijs2011}, \textsc{SPT-3G} \citep{Benson2014} and Advanced ACTpol \citep{Henderson2016}, will be the first to have sufficient volume to yield significant samples of massive clusters. Due to their rarity, the majority of these massive clusters will be at high redshift and it is therefore critical to understand how the cluster observables and their associated scatter evolve. Additionally, the most massive clusters will be the brightest and easiest to detect objects at high redshift, making it vital to understand the selection function of the chosen cluster observable and whether the most massive clusters are representative of the underlying cluster population. Theoretical modelling of the formation of clusters and their observable properties is required to understand these issues and to further clusters as probes of cosmology. Due to the range of scales involved in cluster formation, the need to incorporate astrophysical processes and to self-consistently predict observable properties, cosmological hydrodynamical simulations are the only viable option. Recent progress in the modelling of large-scale structure formation has been driven mainly by the inclusion of supermassive black holes and their associated Active Galactic Nucleus (AGN) feedback, which has been shown to be critical for reproducing many cluster properties \citep{Bhattacharya2008,PuchweinSijackiSpringel2008,McCarthy2010,Fabjan2010}. A number of independent simulations are now able to produce realistic clusters that simultaneously reproduce many cluster properties in good agreement with the observations \citep{LeBrun2014,Pike2014,Planelles2014}. Results from the recent BAryons and HAloes of MAssive Systems (BAHAMAS) simulations \citep{McCarthy2016} have shown that by calibrating the subgrid model for feedback to match a small number of key observables, in this case the global galaxy stellar mass function and the gas fraction of clusters, simulations of large-scale structure are now able to reproduce many observed scaling relations and their associated scatter over two decades in halo mass. However, full gas physics simulations of large-scale structure formation, with sufficient resolution, are still computationally expensive. This has limited previous studies to either small samples with $<50$ objects or to volumes of $596\,\mathrm{Mpc}$, all of which are too small to contain the representative sample of massive clusters that is required for cosmological studies above $z=0$. This paper introduces the Virgo consortium's MACSIS project, a sample of $390$ massive clusters selected from a large volume dark matter simulation and resimulated with full gas physics to enable self-consistent observable predictions. The simulations extend the BAHAMAS simulations to the most massive clusters expected to form in a $\Lambda\rm{CDM}$ cosmology. In this paper we study the cluster scaling relations and their evolution. We combine the MACSIS and BAHAMAS simulations to produce a sample that spans the complete mass range and that can be studied to high redshift, using the progenitors of the MACSIS sample. We also select the hottest clusters from the combined sample and a relaxed subset of them to examine the impact of such selections on the scaling relations and their evolution. We then study the gas profiles to further understand the differences between the samples. This paper is organised as follows. In Section \ref{sec:MACsamp} we introduce the MACSIS sample and discuss the parent dark matter simulation from which the sample was selected, the selection criteria used, the model used to resimulate the haloes, how we produced the observable quantities and the three samples we use in this work. In Section \ref{sec:screlations} we investigate how the scaling relations evolve and how this evolution changes when a hot cluster sample or relaxed, hot cluster sample is selected. We then study the hot gas profiles to understand the differences in the evolution of the relations for the different samples in Section \ref{sec:gasprofs}. Finally, in Section \ref{sec:sad} we discuss our results and summarise our main findings. \section{Parent simulation and sample selection} \label{sec:MACsamp} In this section we describe the parent simulation, the selection of the MACSIS sample, the baryonic physics used in the resimulation of the sample and the calculation of the observable properties of the resimulated clusters. Additionally, we describe how MACSIS and BAHAMAS clusters were selected to produce the combined sample and the cuts made to yield a hot sample and its relaxed subset. \subsection{The parent simulation} To obtain a population of massive clusters we require a simulation with a very large volume $(> 1\,\rm{Gpc}^{3})$. With current computational resources it is unfeasible to simulate such a volume with hydrodynamics and the required gas physics, such as radiative cooling, star formation and feedback, at a resolution high enough to accurately capture the cluster properties. An alternative option is to apply the zoomed simulation technique to a representative sample of objects from a larger volume. Therefore, we select a sample of massive haloes from a dark matter only simulation that has sufficient volume to yield a population of massive clusters and the resolution to ensure they are well characterized. We label this simulation the `parent' simulation. The parent simulation is a periodic cube with a side length of $3.2\,\rm{Gpc}$. Its cosmological parameters are taken from the Planck 2013 results combined with baryonic acoustic oscillations, WMAP polarization and high multipole moments experiments \citep{Planck2014I} and are $\Omega_{\rm{b}}=0.04825$, $\Omega_{\rm{m}}=0.307$, $\Omega_{\Lambda}=0.693$, $h\equiv H_0/(100\,\rm{km}\,\rm{s}^{-1}\,\rm{Mpc}^{-1})=0.6777$, $\sigma_{8}=0.8288$, $n_{\rm{s}}=0.9611$ and $Y=0.248$. We note that there are minor differences between these values and the Planck-only cosmology used for the BAHAMAS simulations, but this has negligible impact on the results presented here. The simulation contained $N=2520^3$ dark matter particles that were arranged in an initial glass-like configuration and then displaced according to second-order Lagrangian perturbation theory $(\mathrm{2LPT})$ using the \textsc{ic\_2lpt\_gen} code \citep{Jenkins2010} and the public Gaussian white noise field \textit{Panphasia} \citep{Jenkins2013,JenkinsBooth2013}. \footnote{The phase descriptor for this volume is [Panph1, L14, (2152, 5744, 757), S3, CH1814785143, EAGLE\_L3200\_VOL1].} The particle mass of this simulation is $m_{\rm{DM}}=5.43\times10^{10}\,\mathrm{M}_{\rm{\odot}}/h$ and the comoving gravitational softening length was set to $40\,\mathrm{kpc}/h$. The simulation was evolved from redshift $z=127$ using a version of the Lagrangian TreePM-SPH code \textsc{p-gadget3} \citep[last described in][]{Springel2005}. Haloes were identified at $z=0$ using a \textit{Friends-of-Friends} (FoF) algorithm with a standard linking length of $b=0.2$ in units of the mean interparticle separation \citep{Davis1985}. \begin{figure} \includegraphics[width=\columnwidth]{Parent_Mass_Function.png} \caption{Comparison of the \textit{Friends-of-Friends} mass function of the parent simulation against those from \citet{Jenkins2001}, \citet{Angulo2012}, \citet{Watson2013} and \citet{Heitmann2015} (top) with the residual differences (bottom). We find good agreement with \citet{Heitmann2015}, but for values of $\ln(\sigma^{-1})>0.4$ we find a growing discrepancy between the parent simulation and the other simulations. This is likely due to our use of $2^{\mathrm{nd}}$ order Lagrangian perturbation theory when generating the initial conditions of the parent simulation and cosmic variance for the rarest haloes.} \label{fig:massfunc} \end{figure} We plot the FoF mass function of the parent simulation at $z=0$ in Fig. \ref{fig:massfunc}. We compare it to the published relations of \citet{Jenkins2001}, \citet{Angulo2012}, \citet{Watson2013} and \citet{Heitmann2015}. We plot the scaled differential mass function \begin{equation} f(\sigma) = \frac{M}{\bar{\rho}}\frac{dn}{d\ln\sigma^{-1}}(M,z)\:, \end{equation} where $M$ is halo mass, $\bar{\rho}$ is the mean density of the Universe at $z=0$, $n$ is the number of haloes per unit volume, and $\sigma^{2}$ is the variance of the linear density field when smoothed with a top-hat filter. We plot the mass function as a function of the variable $\ln(\sigma^{-1})$ as it is insensitive to cosmology \citep{Jenkins2001}. For $\ln(\sigma^{-1}) < 0.3$ we find that all of the mass functions show reasonable agreement with differences of $\sim5-10\%$ between them, with the small differences likely due to the mass function not being exactly universal \citep{Tinker2008,Courtin2011}. However, for larger values the mass functions begin to diverge, as the parent simulation has an excess of massive clusters compared to the other simulations. This is likely due to two effects. First, the MACSIS simulation is the only one to use $\mathrm{2LPT}$ when generating the initial conditions. It has been shown that not using $\mathrm{2LPT}$ results in a significant underestimation of the abundance of the rarest objects \citep{Crocce2006,Reed2013}. The second effect is simply statistics: even in a very large volume there are still low numbers of the rarest and most massive clusters, where there is likely to be significant variance between the simulation volumes. \subsection{The MACSIS sample} To select the MACSIS sample, all haloes with $M_{\rm{FoF}}>10^{15}\,\mathrm{M}_{\odot}$ were grouped in logarithmically spaced bins , with $\Delta\log_{10}\,M_{\rm{FoF}}=0.2$. If a bin contained less than one hundred haloes then all of the objects in that bin were selected. For bins with more than one hundred objects the bin was then further subdivided into bins of 0.02 dex and ten objects from each sub-bin were then selected at random. The subdividing of the bins ensured that our random selection was not biased to low masses by the steep slope of the mass function. This selection procedure results in a sample of $390$ haloes that is mass limited above $10^{15.6}\,\rm{M}_{\odot}$ and randomly sampled below this limit. Table \ref{tab:smpcomp} shows the fraction of haloes selected from the parent simulation in each mass bin. We have compared the properties of the selected haloes with those of the underlying population and found the MACSIS sample to be representative. Additionally, in Appendix \ref{app:seleff} we demonstrate that selecting by a halo's FoF mass does not bias our results when binning clusters by their $M_{500}$. \begin{table} \caption{Table showing the fraction of haloes from the parent simulation that are part of the MACSIS sample for the selection mass bins. The sample is complete above $M_{\rm{FoF}} > 10^{15.6}\,\mathrm{M}_{\odot}$. The parent simulation contains $9754$ haloes with $M_{\rm{FoF}} > 10^{15.0}\,\mathrm{M}_{\odot}$ at $z=0$.} \centering \begin{tabular}{l r r c} \hline Mass Bin & \multicolumn{1}{c}{Sample} & \multicolumn{1}{c}{Total} & Fraction \\ & \multicolumn{1}{c}{Size} & \multicolumn{1}{c}{Haloes} & Selected \\ \hline $15.0 \leq \log_{10}(M_{\rm{FoF}}) < 15.2$ & $100~~$ & $7084~$ & $0.01$ \\ $15.2 \leq \log_{10}(M_{\rm{FoF}}) < 15.4$ & $100~~$ & $2095~$ & $0.05$ \\ $15.4 \leq \log_{10}(M_{\rm{FoF}}) < 15.6$ & $100~~$ & $485~$ & $0.21$ \\ $15.6 \leq \log_{10}(M_{\rm{FoF}}) < 15.8$ & $83~~$ & $83~$ & $1.00$ \\ $15.8 \leq \log_{10}(M_{\rm{FoF}})$ & $7~~$ & $7~$ & $1.00$ \\ \hline \end{tabular} \label{tab:smpcomp} \end{table} \begin{figure*} \includegraphics[width=\textwidth]{BigBoxZoom.png} \caption{Slice of depth $40\,\rm{Mpc}$ through the parent simulation showing the projected dark matter density at $z=0$. The left inset shows a $50\,\rm{Mpc}$ cube centred on the most massive halo. The right inset shows the stellar particles of the same halo in yellow, re-simulated using the BAHAMAS model and resolution, with X-ray emission from the hot gas overlaid in purple.} \label{fig:zoomed} \end{figure*} Due to current computational constraints the BAHAMAS simulations are limited to periodic cubes with a side length of $596\,\mathrm{Mpc}$. There are very few clusters with a mass greater than $10^{15}\,\mathrm{M}_{\odot}$ in a volume of this size, and those that are present may be affected by the loss of power from large-scale modes that are absent due to their wavelengths being greater than the box size. The zoom simulations of the MACSIS project provide an extension to the BAHAMAS periodic simulations. They provide the most massive clusters and allow the mass-observable scaling relations to be studied across the complete cluster mass range. We use the zoomed simulation technique \citep{KatzWhite1993,Tormen1997} to re-simulate the chosen sample at increased resolution. We perform both DM only and full gas physics re-simulations. The Lagrangian region for every cluster was selected so that its volume was devoid of lower resolution particles beyond a cluster centric radius of $5r_{200}$.\footnote{We define $r_{200}$ as the radius at which the enclosed average density is two hundred times the critical density of the Universe.} The resolution of the Lagrangian region was increased such that the particles in the DM only simulations had a mass of $m_{\rm{DM}}=5.2\times10^9\,\mathrm{M}_{\odot}/h$ and in the hydrodynamic re-simulations the dark matter particles had a mass of $m_{\rm{DM}}=4.4\times10^9\,\mathrm{M}_{\odot}/h$ and the gas particles had an initial mass of $m_{\rm{gas}}=8.0\times10^8\,\mathrm{M}_{\odot}/h$. In all simulations the Plummer equivalent gravitational softening length for the high-resolution particles was fixed to $4\mathrm{kpc}/h$ in comoving units for $z>3$ and in physical coordinates thereafter. The smoothed particle hydrodynamics interpolation used $48$ neighbours and the minimum smoothing length was set to one tenth of the gravitational softening. A schematic view of the zoom approach is shown in Fig. \ref{fig:zoomed}. The resolution and softening of the zoom re-simulations were deliberately chosen to match the values of the periodic box simulations of the BAHAMAS project \citep{McCarthy2016}, which is a calibrated version of the OWLS code \citep{Schaye2010}, which was also used for cosmo-OWLS \citep{LeBrun2014}. The subgrid models for feedback from star formation and AGN used in the BAHAMAS simulations was calibrated to obtain a good fit to the observed galaxy stellar mass function and the amplitude of the gas fraction-total mass relation, respectively, at $z=0$. Without any further tuning, the simulations then produce a population of groups and clusters that shows excellent agreement with the observations for a range of galaxy-halo, hot gas-halo and galaxy-hot gas relations. \subsection{Baryonic physics} The BAHAMAS simulations were run with a version of \textsc{p-gadget3} that has been heavily modified to include new subgrid physics as part of the OWLS project \citep{Schaye2010}. We now briefly describe the subgrid physics, but refer the reader to \citet{Schaye2010}, \citet{LeBrun2014} and \citet{McCarthy2016} for greater detail, including the impact of varying the free parameters in the model and the calibration strategy. Radiative cooling is calculated on an element-by-element basis following \citet{WiersmaSchayeSmith2009}, interpolating the rates as a function of density, temperature and redshift from pre-computed tables generated with \textsc{cloudy} \citep{Ferland1998}. It accounts for heating and cooling due to the primary cosmic microwave background and a \citet{HaardtMadau2001} ultra-violet/X-ray background. The background due to reionization is assumed to switch on at $z=9$. Star formation is modelled stochastically in a way that by construction reproduces the observations, as discussed in \citet{SchayeDallaVecchia2008}. Lacking the resolution and physics to correctly model the cold interstellar medium, gas particles with a density $n_{\rm{H}} > 0.1\,\rm{cm}^{-3}$ follow an imposed equation of state with $P\propto\rho^{4/3}$. These gas particles then form stars at a pressure-dependent rate that reproduces the observed Kennicutt-Schmidt law \citep{Schmidt1959,Kennicutt1998}. Stellar evolution and the resulting chemical enrichment are implemented using the model of \citet{Wiersma2009}, where $11$ chemical elements (H, He, C, N, O, Ne, Mg, Si, S, Ca and Fe) are followed. The mass loss rates are calculated assuming Type Ia and Type II supernovae and winds from massive and asymptotic giant branch stars. Stellar feedback is implemented via the kinetic wind model of \citet{DallaVecchiaSchaye2008}. The BAHAMAS simulations used the calibrated mass-loading factor of $\eta_{\rm{w}}=2$ and wind velocity $v_{\rm{w}}=300\,\rm{km/s}$. This corresponds to $20$ percent of available energy from Type II supernovae, assuming a \citet{Chabrier2003} IMF, and yields an excellent fit to the observed galaxy mass function. The seeding, growth and feedback from supermassive black holes (BH) is implemented using the prescription of \citet{BoothSchaye2009}, a modified version of the method developed by \citet{SpringelDiMatteoHernquist2005}. A FoF algorithm is run on-the-fly and BH seed particles, with $m_{\mathrm{BH}}=10^{-3}m_{\mathrm{gas}}$, are placed in haloes that contain at least $100$ DM particles, which corresponds to a halo mass of $\sim5\times10^{11}\,\rm{M}_{\odot}$. BHs grow via Eddington-limited accretion of gas at the Bondi-Hoyle-Littleton rate, with a boost factor that is a power-law of the local density for gas above the star formation density threshold. They also grow by direct mergers with other BHs. A fraction, $\epsilon$, of the rest mass energy of the accreted gas is then used to heat $n_{\rm{heat}}$ neighbour particles by increasing their temperature by $\Delta T_{\rm{heat}}$. Changes to these parameters have a significant impact on the hot gas properties of clusters. The calibrated values of these parameters in the BAHAMAS simulations are $n_{\rm{heat}}=20$ and $\Delta T_{\rm{heat}}=10^{7.8}\,\rm{K}$. The feedback efficiency $\epsilon=\epsilon_{\rm{r}}\epsilon_{\rm{f}}$, where $\epsilon_{\rm{r}}=0.1$ is the radiative efficiency and $\epsilon_{\rm{f}}=0.15$ is the fraction of $\epsilon_{\rm{r}}$ that couples to the surrounding gas. The choice of the efficiency, assuming it is non-zero, is generally of little consequence as the feedback establishes a self-regulating scenario, but determines the black hole masses \citep{BoothSchaye2009}. \subsection{Calculating observable properties} Previous studies have shown that there can be significant biases in the observable properties of clusters due to issues such as multi-temperature structures and gas inhomogeneities \citep[e.g.][]{NagaiVikhlininKravtsov2007b,Khedekar2013}. Therefore, when investigating cluster properties it is critical that, as far as possible, we make a like-with-like comparison with the observations. Following \citet{LeBrun2014}, we do this by producing synthetic observational data for each cluster and analysing it in a manner similar to what is done for real data. Using the particle's temperature, density and metallicity, where the metallicity is smoothed over a particle's neighbours, we first compute a rest-frame X-ray spectrum in the $0.05-100.0\,\rm{keV}$ band for all gas particles, using the Astrophysical Plasma Emission Code \citep[\textsc{apec};][]{Smith2001} via the \textsc{pyatomdb} module with atomic data from \textsc{atomdb} v3.0.2 \citep[last described in][]{Foster2012}. A particle's spectrum is a sum of the individual spectra for each chemical element tracked by the simulations, scaled by the particle's elemental abundance. We ignore particles with a temperature lower than $10^5\,\rm{K}$ as they make a negligible contribution to the total X-ray emission. We then estimate the density, temperature and metallicity of the hot gas in $25$ logarithmically spaced radial bins by fitting a single temperature \textsc{apec} model, with a fixed metallicity, to the summed spectra of all particles that fall within that radial bin. We then scale the spectra by the relative abundance of the heavy elements as the fiducial spectra assume solar abundance \citep{AndersGrevesse1989}. The spectra have an energy resolution of $150\,\rm{eV}$ in the range $0.05-10.0\,\rm{keV}$ and are logarithmically spaced between $10.0-100.0\,\rm{keV}$. To get a closer match to the observations, we multiply the spectra by the effective area of \textit{Chandra}. To derive temperature and density profiles of a cluster, we fit the spectrum in the range $0.5-10.0\,\rm{keV}$ for each radial bin with a single temperature model using a least-squares approach. The temperature and density profiles derived from the X-ray spectra are then used to perform a hydrostatic mass analysis of the cluster. The profiles are fit with the density and temperature models proposed by \citet{Vikhlinin2006} to produce a hydrostatic mass profile. We then derive various mass and radius estimates, such as $M_{500}$ and $r_{500}$, from the hydrostatic mass profiles. With these estimates we calculate quantities, such as $M_{\rm{gas}}$ or $Y_{\rm{SZ}}$, by summing the properties of the particles that fall within the set. Core-excised quantities are calculated in the radial range $0.15-1.0$ of the aperture. Luminosities are calculated by integrating the spectra of all particles within the aperture in the requisite energy band, for example, bolometric luminosities are calculated in the range $0.05-100.0\,\rm{keV}$. Averaged X-ray temperatures are calculated by fitting a single temperature model to the sum of the spectra of all particles within the aperture. We repeat this analysis for all clusters in the combined sample at all redshifts of interest. All quantities derived in this manner are labelled with the sub-script `spec'. \subsection{Cluster sample selection} \begin{figure*} \includegraphics[width=\textwidth]{Observations.png} \caption{Gas mass-total mass relation (top left), core-excised bolometric X-ray luminosity-core-excised X-ray temperature relation (top right) and the integrated Sunyaev-Zel'dovich signal-total mass relation at $z=0$ (bottom left) and $z=1$ (bottom right) for the combined sample. The median relation of the BAHAMAS sample is given by the red line, with the red hatch region enclosing $68\%$ of the population, and the median MACSIS result is shown by the blue line, with the blue hatched region enclosing $68\%$ of the sample. The median MACSIS line becomes dashed when there are less than $10$ clusters in a bin. The black triangles, crosses, squares, right-facing triangles, circles, left-facing triangles, hexagons and pluses are observational data from \citet{Vikhlinin2006}, \citet{Maughan2008}, \citet{Pratt2009}, \citet{Lin2012}, \citet{Maughan2012} and the second \textit{Planck} SZ catalogue \citep{PlanckSZ2015} respectively.} \label{fig:observations} \end{figure*} We select clusters from MACSIS and BAHAMAS to form a `combined' sample with which we can investigate the cluster scaling relations. We perform our analysis at $z=0.0,0.25,0.5,1.0$ and $1.5$. We create this sample at each redshift by selecting all clusters with a mass $M_{500,\rm{spec}}\geq10^{14}\,\rm{M}_{\odot}$. Additionally, we introduce a mass cut at every redshift below which we remove any MACSIS clusters. For example, at $z=0$ ($z=1$) this cut is made at $M_{500,\mathrm{spec}}=10^{14.78}\,\rm{M}_{\odot}$ ($M_{500,\mathrm{spec}}=10^{14.3}\,\rm{M}_{\odot}$). This removes a tail of clusters with low $M_{500,\mathrm{spec}}$, but have high $M_{\mathrm{FoF}}/M_{500,\mathrm{spec}}$ ratios (see Appendix \ref{app:seleff}). For the luminosity-temperature relation, we use the temperature-mass relation of the combined sample to convert the mass cut into a temperature cut. At $z=0$ this results in a sample of $1294$ clusters, containing $1098$ clusters from BAHAMAS and $196$ MACSIS clusters, and at $z=1$ a sample of $225$ clusters, $99$ from BAHAMAS and $126$ from MACSIS. The MACSIS clusters enable the investigation of the behaviour of the most massive clusters at low redshift. These clusters are commonly selected in cosmological analyses because their deep potentials are expected to reduce the impact of non-gravitational processes and as the brightest clusters they require shorter exposures. We select a hot, and therefore massive, cluster sample by selecting all clusters in the combined sample with a core-excised X-ray temperature greater than $5\,\rm{keV}$. At $z=0$ ($z=0.5$) this yields a sample of $244$ ($186$) clusters, with $190$ ($173$) coming from the MACSIS sample. Finally, we examine the impact of selecting a relaxed subset of the hot cluster sample. Theoretically, there are many ways to define a relaxed halo \citep[see][]{Neto2007,Duffy2008,Klypin2011,DuttonMaccio2014,Klypin2016}. For this study we use the following criteria \begin{equation} X_{\rm{off}} < 0.07\,;~f_{\rm{sub}} < 0.1~\rm{and}~\lambda < 0.07\,, \nonumber \end{equation} where $X_{\rm{off}}$ is distance between the cluster's minimum gravitational potential and centre of mass, divided by its virial radius; $f_{\rm{sub}}$ is the mass fraction within the virial radius that is bound to substructures; and $\lambda$ is the spin parameter for all particles inside $r_{200}$. These criteria are not designed to select a small subset that comprises the most relaxed objects, but to simply remove those clusters that are significantly disturbed. This results in a subsample at $z=0$ ($z=0.5$) that contains $213$ ($117$) clusters, with $177$ ($111$) coming from the MACSIS sample. \section{The scaling relations of massive clusters} \label{sec:screlations} In this section we present our main results, measuring the scaling relations of our cluster samples across a range of redshifts. \subsection{Comparison to observational data} Fig. \ref{fig:observations} shows the gas mass, $M_{\rm{gas},500,\rm{spec}}$, the integrated Sunyaev-Zel'dovich (SZ) signal, $Y_{\rm{SZ}}$, measured in a $5r_{500,spec}$ aperture as a function of estimated total mass, $M_{500,\rm{spec}}$, (at $z=0$ and $z=1$) and the core-excised bolometric X-ray luminosity, $L^{\rm{X,ce}}_{500,\rm{spec}}$, as a function of core-excised X-ray temperature, $T^{\rm{X,ce}}_{500,\rm{spec}}$, for the combined sample. We compare the sample to the relevant observational data. At all redshifts the MACSIS sample provides a consistent extension to the BAHAMAS clusters with similar scatter. At low redshift, \citet{McCarthy2016} have shown that the BAHAMAS sample shows good agreement with the observed median relations and shows similar intrinsic scatter. The MACSIS sample continues this agreement to observed high-mass clusters, though there are significantly fewer clusters to compare against. In detail, it appears that the $M_{500,\mathrm{gas,spec}}$-$M_{500,\mathrm{spec}}$ and $L^{\mathrm{X,ce}}_{500,\mathrm{spec}}$-$T^{\mathrm{X,ce}}_{500,\mathrm{spec}}$ relations are slightly steeper than observed. However, we would exercise caution as we have not applied the same selection criteria as was used for the observational X-ray analyses. At high redshift observational data becomes sparse and currently only SZ surveys have detected a reasonable number of clusters. At $z=1$ these clusters are all significantly more massive than any cluster in the BAHAMAS volume. However, the progenitors of the very massive MACSIS clusters provide a sample that can be compared with these observations. We find that the median relation shows good agreement with the observations and the intrinsic scatter of the clusters about the median relation is consistent with the scatter in the observations. Overall, we find that all quantities computed in a like-with-like manner show good agreement with the observations. \renewcommand\arraystretch{1.5} \newcolumntype{C}{>{\centering\arraybackslash}X} \begin{table*} \caption{The normalization and slope of the best-fit relations presented in this work and the scatter about them for the three samples at $z=0$. All quantities presented in this table are `$\rm{spec}$' values calculated via the synthetic X-ray analysis within an aperture of $r_{500,\rm{spec}}$. The scatter $\langle\sigma_{\log_{10}Y}\rangle$ is averaged over all masses.} \centering \begin{tabularx}{\textwidth}{l r C C r C C r C C} \hline Scaling relation & \multicolumn{3}{c}{Combined sample} & \multicolumn{3}{c}{Hot Clusters} & \multicolumn{3}{c}{Relaxed, Hot Clusters} \\ & \multicolumn{1}{c}{$A$} & \multicolumn{1}{c}{$\alpha$} & \multicolumn{1}{c}{$\langle\sigma_{\log_{10}Y}\rangle$} & \multicolumn{1}{c}{$A$} & \multicolumn{1}{c}{$\alpha$} & \multicolumn{1}{c}{$\langle\sigma_{\log_{10}Y}\rangle$} & \multicolumn{1}{c}{$A$} & \multicolumn{1}{c}{$\alpha$} & \multicolumn{1}{c}{$\langle\sigma_{\log_{10}Y}\rangle$} \\ \hline $L^{\rm{X,ce}}_{500}-M_{500}$ & $44.50^{+0.01}_{-0.01}$ & $1.88^{+0.03}_{-0.05}$ & $0.15^{+0.01}_{-0.02}$ & $44.71^{+0.02}_{-0.02}$ & $1.36^{+0.08}_{-0.07}$ & $0.12^{+0.01}_{-0.02}$ & $44.69^{+0.03}_{-0.03}$ & $1.43^{+0.13}_{-0.09}$ & $0.11^{+0.01}_{-0.01}$ \\ $k_{\rm{B}}T^{\rm{X,ce}}_{500}-M_{500}$ & $0.68^{+0.01}_{-0.01}$ & $0.58^{+0.01}_{-0.01}$ & $0.048^{+0.003}_{-0.003}$ & $0.71^{+0.01}_{-0.01}$ & $0.51^{+0.04}_{-0.04}$ & $0.05^{+0.01}_{-0.01}$ & $0.70^{+0.01}_{-0.01}$ & $0.55^{+0.06}_{-0.03}$ & $0.04^{+0.01}_{-0.01}$ \\ $M_{\rm{gas},500}-M_{500}$ & $13.67^{+0.01}_{-0.01}$ & $1.25^{+0.01}_{-0.03}$ & $0.07^{+0.01}_{-0.01}$ & $13.77^{+0.01}_{-0.01}$ & $1.02^{+0.03}_{-0.03}$ & $0.06^{+0.01}_{-0.01}$ & $13.75^{+0.01}_{-0.01}$ & $1.05^{+0.04}_{-0.04}$ & $0.05^{+0.01}_{-0.01}$ \\ $Y_{\rm{X},500}-M_{500}$ & $14.33^{+0.01}_{-0.01}$ & $1.84^{+0.02}_{-0.05}$ & $0.12^{+0.01}_{-0.01}$ & $14.47^{+0.02}_{-0.02}$ & $1.51^{+0.07}_{-0.08}$ & $0.11^{+0.01}_{-0.01}$ & $14.45^{+0.02}_{-0.02}$ & $1.59^{+0.12}_{-0.06}$ & $0.08^{+0.01}_{-0.01}$ \\ $Y_{\rm{SZ},500}-M_{500}$ & $-4.51^{+0.01}_{-0.01}$ & $1.88^{+0.02}_{-0.03}$ & $0.10^{+0.01}_{-0.01}$ & $-4.39^{+0.02}_{-0.02}$ & $1.60^{+0.07}_{-0.05}$ & $0.10^{+0.01}_{-0.02}$ & $-4.42^{+0.02}_{-0.02}$ & $1.69^{+0.07}_{-0.07}$ & $0.09^{+0.01}_{-0.01}$ \\ $L^{\rm{X,ce}}_{500}-T^{\rm{X,ce}}_{500}$ & $44.80^{+0.02}_{-0.01}$ & $3.01^{+0.04}_{-0.04}$ & $0.14^{+0.01}_{-0.01}$ & $44.93^{+0.01}_{-0.01}$ & $2.41^{+0.12}_{-0.12}$ & $0.11^{+0.01}_{-0.01}$ & $44.89^{+0.02}_{-0.02}$ & $2.53^{+0.12}_{-0.13}$ & $0.10^{+0.01}_{-0.01}$ \\ \hline \end{tabularx} \label{tab:z0bestfit} \end{table*} \renewcommand\arraystretch{1.0} \subsection{Modelling cluster scaling relations} As a baseline for understanding how the scaling relations evolve as a function of mass and redshift, we adopt the following self-similar scalings \begin{equation} M_{\rm{gas},\Delta}\propto M_{\Delta}\,, \end{equation} \begin{equation} \label{eq:T-M} T_{\Delta}\propto M^{2/3}_{\Delta}E^{2/3}(z)\,, \end{equation} \begin{equation} Y_{\rm{X},\Delta}\propto M_{\Delta}^{5/3}E^{2/3}(z)\,, \end{equation} \begin{equation} Y_{\rm{SZ},\Delta}\propto M_{\Delta}^{5/3}E^{2/3}(z)\,, \end{equation} \begin{equation} \label{eq:L-M} L_{\Delta}^{\rm{X,bol}}\propto M^{4/3}_{\Delta}E^{7/3}(z)\,, \end{equation} \begin{equation} L_{\Delta}^{\rm{X,bol}}\propto T^{2}E(z)\,, \end{equation} where $E(z)\equiv H(z)/H_0=\sqrt{\Omega_{m}(1+z)^3+\Omega_{\Lambda}}$, $\Delta$ is the chosen overdensity relative to the critical density and $Y_{\rm{X}}$ is the X-ray analogue of the integrated SZ effect. These are derived in Appendix \ref{app:ssr}. Although shown to be too simplistic by the first X-ray studies of clusters \citep{Mushotzky1984,EdgeStewart1991,David1993}, the self-similar relations allow us to investigate if astrophysical processes are less significant in more massive clusters or at higher redshift. To enable a comparison with the self-similar predictions, and previous work, we fit the scaling relations of our samples at each redshift. We derive a median relation by first binning the clusters into bins of log mass (width $0.1$ dex) or log temperature (width $0.07$ dex) and then computing the median in each bin with more than ten clusters. We also remove the evolution in normalization predicted by self-similar relations. The medians of the bins are then fit with a power-law of the form \begin{equation} E^{\beta}(z)Y=10^A\left(\frac{X}{X_0}\right)^{\alpha}\,, \label{eq:plfit} \end{equation} where $A$ and $\alpha$ describe the normalization and slope of the best fit respectively, $\beta$ removes the expected self-similar evolution with redshift, $X$ is either the total mass or temperature and $Y$ is the observable quantity ($M_{\rm{gas}}$, $L^{\rm{X,bol}}$, etc.). $X_0$ is the pivot point, which we set to $4\times10^{14}\,\rm{M}_{\odot}$ for observable-mass relations and to $6\,\rm{keV}$ for observable-temperature relations. We note that we fix the pivot for all samples and all redshifts. Fitting to the medians of bins, rather than individual clusters, prevents the fit from being dominated by low-mass objects, which are significantly more abundant due to the shape of the mass function. For the hot sample and its relaxed subset there are too few bins with ten or more clusters to reliably derive a best-fit relation at $z\geq1$. By limiting our sample to systems with $M_{500}\geq10^{14}\,\rm{M}_{\odot}$ we avoid any breaks in the powerlaw relations that have been seen both observationally and in previous simulation work \citep{LeBrun2016}. We compute the scatter about the best-fit relation at each redshift by calculating the root mean squared (\textit{rms}) dispersion in each bin according to \begin{equation} \sigma_{\log_{10}Y}=\sqrt{\frac{1}{N}\sum_{i=1}^N\left[\log_{10}(Y_i)-\log_{10}(Y_{\rm{BF}})\right]^2}\,, \end{equation} where $i$ runs over all clusters in the bin, $Y_{\rm{BF}}$ is the best fit relation for a cluster with a value $X_i$ and we note that $\sigma_{\ln Y}=\ln(10)\sigma_{\log_{10}Y}$. We obtain the uncertainties for our fit parameters by bootstrap re-sampling the clusters $10,000$ times. The best-fit values of all the scaling relations considered for the three samples (combined, hot and relaxed) at $z=0$ are summarized in Table \ref{tab:z0bestfit} and other redshifts are listed in Appendix \ref{app:fitpar}. We now discuss each relation in turn. \subsection{Gas Mass-Total Mass} \label{ssec:Mg-M} \begin{figure*} \includegraphics[width=\textwidth]{Mg-M_ScalRel_spec.png} \caption{Evolution of the gas mass-total mass scaling relation for the three samples as a function of redshift. The top left panel shows the median gas mass in bins of total mass at $z=0$ (blue) and $z=1$ (red) for the combined sample, with error bars showing the $16^{\rm{th}}$ and $84^{\rm{th}}$ percentiles of the distribution in each bin. The solid (dashed) line shows the best-fit relation at $z=0$ ($z=1$). Note that only two redshifts are shown for clarity. The top right panel shows the \textit{rms} scatter in each mass bin at each redshift for the combined sample. The bottom panels show the best-fit normalization, $A$, (left) and slope, $\alpha$, (right) of the scaling relation as a function of $\log_{10}(1+z)$ for the three different samples: combined (blue squares), hot clusters (red triangles) and relaxed hot sample (black diamonds). We have offset the points for clarity. The dot-dashed magenta line shows the value of the predicted self-similar slope.} \label{fig:MgMsr} \end{figure*} We plot the hot gas mass-total mass scaling relation for the three samples in Fig. \ref{fig:MgMsr}. The best-fit normalization for the combined sample shows significant evolution with redshift, with clusters of a fixed mass containing $25\%$ more hot gas at $z=1$ than at $z=0$. With the inclusion of star formation, radiative cooling and feedback from supernovae and AGN, the departure from self-similarity is not unexpected. The increasing normalization with redshift is due to either the impact of AGN feedback or the conversion of gas to stars. As the normalization of the baryonic mass exhibits a similar trend, this evolution is being driven by AGN feedback. A plausible explanation is as follows. The mean density of the Universe increases with redshift and cluster potentials at a fixed mass get deeper with increasing redshift. This reduces the efficiency with which AGN expel gas from the cluster with increasing redshift, leading to a higher gas mass at higher redshift for clusters at a fixed mass. In addition, AGN have less time to act on and expel gas from clusters that form at higher redshifts. The AGN breaks the self-similar assumption of a constant gas fraction, resulting in the normalization of the gas mass-total mass relation increasing with increasing redshift. However, we note that this behaviour appears to be dependent on the implementation of the subgrid physics. \citet{LeBrun2016} use the same subgrid implementation, but with different parameters, and obtain similar behaviour. However, \citet{Planelles2013} see a constant baryon fraction with redshift suggesting that feedback is not expelling gas beyond $r_{500}$. The bottom left panel of Fig. \ref{fig:MgMsr} shows that the normalizations of the best-fit relations for the hot sample of clusters and for the relaxed subset of hot clusters are higher at $z=0$ than the normalization of the combined sample and evolve less with redshift. This is because hotter clusters are generally more massive and have deeper potential wells, reducing the amount of gas the AGN can permanently expel from the cluster during its formation. This flattens the slope of the relation leading to a higher normalization at the pivot. The bottom right panel of Fig. \ref{fig:MgMsr} shows that the slope of the best-fit relation of the combined sample is significantly steeper than the self-similar prediction of unity. At a given redshift AGN feedback has expelled more gas from lower mass clusters, due to their shallower potentials, leading to a tilt in the relation. We find a slope of $\alpha=1.25^{+0.01}_{-0.03}$. Our slope is mildly shallower than found in previous a simulation work, where \citet{LeBrun2016} find a slope of $1.32$ for their AGN8.0 simulation, but consistent with observations, where \citet{Arnaud2007} found a slope of $1.25\pm0.06$ for a sample of clusters observed with \textit{XMM}. We find negligible evolution in the slope of the relation for the combined sample. The hot cluster sample and the relaxed subset have best-fit slopes that are consistent with the self-similar prediction. The increased depth of the potential well in massive clusters means that their gas mass is approximately a constant fraction of their total mass. Specifically, we find that most massive clusters have a median gas fraction $f_{\mathrm{gas}}=0.89\pm0.09$ of the universal baryon fraction at $z=0$. This results in slopes of $\alpha=1.02\pm0.03$ and $1.05\pm0.04$ for the hot cluster sample and the relaxed subset respectively. We find good agreement with the slope of $1.05\pm0.05$ found by \citet{Mantz2016} and the self-similar slope found by \citet{Vikhlinin2009} for relaxed cluster samples. The slope of the best-fit relation for both samples shows no significant evolution with redshift. The top right panel of Fig. \ref{fig:MgMsr} shows that the scatter about the best-fit relation is independent of both mass and redshift. Averaged over all mass bins it has a value of $\sigma_{\log_{10}Y}=0.07$ at $z=0$. The scatter reduces slightly for the hot cluster sample, with a value of $0.06$, and further still for the relaxed subset, with a value of $0.05$. The scatter is in reasonable agreement with the scatter of $0.04$ found by \citet{Arnaud2007} for a sample of clusters observed with \textit{XMM}. \begin{figure*} \includegraphics[width=\textwidth]{Tx-M_ScalRel_spec.png} \caption{Evolution of the core-excised X-ray temperature-total mass scaling relation for the three samples as a function of redshift. The panels are arranged as described in Fig. \ref{fig:MgMsr}.} \label{fig:TxMsr} \end{figure*} \subsection{X-ray Temperature-Mass} \label{ssec:Tx-M} The evolution of the core-excised spectroscopic temperature-total mass scaling relations, and their scatter, for the three samples are shown in Fig. \ref{fig:TxMsr}. The normalization of the best-fit relation of the combined sample shows a minor evolution with redshift, being $15\%$ lower at $z=1$ compared to $z=0$ (bottom left panel). In the self-similar model the temperature of the ICM is related to the depth of the gravitational potential of the cluster, under the assumption of hydrostatic equilibrium. Previous simulation work has shown that the non-thermal pressure in mass-limited samples grows with redshift due to the increasing importance of mergers and resulting incomplete thermalisation \citep{Stanek2010,LeBrun2016}. Therefore, clusters increasingly violate the assumption of hydrostatic equilibrium with redshift and require a lower temperature at a fixed mass to balance gravitational collapse, which leads to a normlization that decreases with redshift compared to self-similar. The effective temperature of the non-thermal pressure can be estimated via \begin{equation} T_{\rm{kin}}=\left(\frac{\mu m_{\rm{p}}}{k_{\rm{B}}}\right)\sigma_{\rm{gas}}^2\, \end{equation} where $\sigma_{\rm{gas}}$ is the 1D velocity dispersion of the gas particles, $\mu=0.59$ is the mean molecular weight, $m_{\rm{p}}$ is the mass of the proton and $k_{\rm{B}}$ is the Boltzmann constant. Fig. \ref{fig:TxTkinMsr} shows the evolution of the temperature-mass normalization once this effective kinetic temperature has been added to the spectral temperature. For all three samples the addition of the kinetic temperature results in a normalization that shows significantly reduced evolution with respect to self-similar. \begin{figure} \includegraphics[width=\columnwidth]{Tx+Tkin-M_Norm_spec.png} \caption{Evolution of the normalization of the spectroscopic temperature-total mass relation when the effective non-thermal support temperature is included. All three samples show negligible evolution with redshift relative to self-similar once non-thermal pressure support is included.} \label{fig:TxTkinMsr} \end{figure} The normalizations of the best-fit relations for the hot cluster and the relaxed hot samples are slightly higher than for the combined sample, but they show a similar trend with redshift that is removed when the kinetic temperature is included. The higher normalization occurs because, again, the hot sample has a flatter slope with mass. This flatter slope is driven by two processes. First, non-thermal pressure support becomes more important in higher mass clusters at a fixed redshift, as they have had less time to thermalise, and this lowers their temperatures. Second, we find that the bias between the spectroscopic and mass-weighted temperatures increases mildly with mass. This does not appear to be caused by cold clumps due to the SPH method, but is due to the presence of cooler gas in the outskirts of massive clusters that is hotter than the $0.5\,\rm{keV}$ lower limit, contributing to the X-ray spectrum, and biasing the measured temperature low for the most massive clusters. Fig. \ref{fig:Tx_bias} shows the fractional difference between the spectroscopic and mass-weighted core-excised temperatures as a function of mass. Similar to \citet{Biffi2014}, we find that for low-mass clusters the spectroscopic temperature estimate agrees well with the mass-weighted estimate at $z=0$. However, as cluster mass increases we find that the spectroscopic estimate is increasingly biased low compared to the mass-weighted estimate. This will also impact the hydrostatic mass estimate of the cluster and we refer the reader to \citet{Henson2016} for a more in-depth study. Both of these effects lead to a flattening of the slope with mass and a higher normalization for the hot samples. We note that removing the most disturbed clusters produces a marginal decrease in the normalization of the relation, which is due to the steeper slope yielding a lower normalization at the pivot point. \begin{figure} \includegraphics[width=\columnwidth]{Tx_bias.png} \caption{Plot of fractional difference between the spectroscopic and mass-weighted temperature estimates as a function of $M_{500}$ for the combined sample at $z=0$ (blue squares) and $z=1$ (red triangles). Error bars show $68\%$ of the population.} \label{fig:Tx_bias} \end{figure} We find the slope of the best-fit relation for the combined sample to be $\alpha=0.58\pm0.01$ at $z=0$. This is in good agreement with the slope found by previous simulation work, where values of $0.55\pm0.01$ \citep{Short2010}, $0.576\pm0.002$ \citep{Stanek2010}, $0.54\pm0.01$ \citep{Planelles2014}, $0.56\pm0.03$ \citep{Biffi2014}, $0.60\pm0.01$ \citep{Pike2014} and $0.58$ \citep{LeBrun2016} were found. All of these are in agreement with the observed temperature-total mass relation found for volume-limited samples, with values of $0.58\pm0.03$ for a sample of clusters observed with \textit{XMM} \citep{Arnaud2007} and $0.56\pm0.07$ for a sample of low-redshift clusters \citep{Giles2015}. We note that a caveat to these comparisons is the differing mass ranges will alter the slope as the relation is not a perfect power law. All of these relations are slightly flatter than the predicted self-similar slope of $2/3$ due to non-thermal pressure support and temperature bias. \begin{figure*} \includegraphics[width=\textwidth]{Yx-M_ScalRel_spec.png} \caption{Evolution of the X-ray analogue $Y_{\rm{X}}$ signal-total mass scaling relation for the three samples as a function of redshift. The panels are arranged as described in Fig. \ref{fig:MgMsr}.} \label{fig:YxMsr} \end{figure*} Selecting only hot clusters produces a best-fit relation with a slope of $0.51\pm0.04$, flatter than the combined relation. The best-fit slope of $0.55^{+0.06}_{-0.03}$ for the relaxed subset, is compatible with the combined sample. The slope of the relaxed subset is compatible with the slope found by \citet{Mantz2016} of $0.66\pm0.05$ and the slope of $0.65\pm0.04$ found by \citet{Vikhlinin2009} for relaxed clusters. However, we note that our relaxation criteria only remove the most disturbed objects, as opposed to the criteria of \citet{Mantz2015} which select the most relaxed objects. Therefore, we would likely recover a steeper slope with stricter relaxation criteria. Both samples are equally affected by the spectroscopic temperature being biased low. The slopes of the hot sample and the relaxed subset show no clear trend with redshift. The temperature-mass scaling relations shows very low scatter, which is independent of both mass and redshift. The average scatter across all mass bins is $\sigma_{\log_{10}Y}=0.046$, $0.045$ and $0.039$ for the combined sample, hot sample and relaxed subset, respectively, at $z=0$. These values are consistent with the values found by both observations and previous simulations \citep{Arnaud2007,Giles2015,Stanek2010,Short2010}. \subsection{$Y_{\rm{X}}$- Mass} The power-law fits to the X-ray analogue of the integrated SZ effect-total mass relations for the three samples, and their scatter, are shown in Fig. \ref{fig:YxMsr}. The X-ray analogue signal, $Y_{\rm{X}}$, is the product of the core excised spectral temperature and the gas mass, and the relation should reflect the combination of the two previously presented relations. We indeed find this to be the case. For the combined sample, the decreasing temperature-total mass normalization with increasing redshift offsets the increasing gas mass-total mass normalization, producing almost no evolution of the normalization for the $Y_{\rm{X}}$-total mass relation. The same trend was found by \citet{LeBrun2016}. Therefore, the normalization evolves in a close to self-similar manner. Selecting a sample of hot clusters or a relaxed subset of them leads to higher overall normalization of the best-fit relation. This is mainly due to the reduced impact of AGN feedback on the gas mass-total mass relation, which flattens the relation and leads to a higher normalization at the pivot. Both samples agree very well with the predicted self-similar evolution of the normalization of the relation, with the normalization of the relaxed subset changing by less than one percent between $z=0$ and $z=0.5$. The slope of the $Y_{\rm{X}}$-total mass relation is simply the sum of the slopes of the temperature-mass and gas mass-total mass relations and for the combined sample the slope is significantly steeper than the $5/3$ value predicted by self-similar theory. We find a value of $\alpha=1.84^{+0.02}_{-0.05}$ at $z=0$. The slope of our best-fit relation is consistent with those of previous simulations, who found values of $1.78\pm0.01$ \citep{Short2010}, $1.73\pm0.01$ \citep{Planelles2014} and $1.89$ \citep{LeBrun2016}. Our result is also in agreement with the observational value found by \citet{Arnaud2007} of $1.82\pm0.1$ using the REXCESS cluster sample. The physical reason for the steeper slope is that gas is preferentially removed from lower mass clusters by feedback. In response to gas expulsion the remaining gas increases in temperature, offsetting some of the losses, but the loss of gas dominates and steepens the relation. The value of the slope for the best-fit relation is approximately constant with redshift, within the uncertainty of the fits. \begin{figure*} \includegraphics[width=\textwidth]{Ysz-M_ScalRel_spec.png} \caption{Evolution of the integrated Sunyaev-Zel'dovich signal-total mass scaling relation for the three samples as a function of redshift. The panels are arranged as described in Fig. \ref{fig:MgMsr}.} \label{fig:YszMsr} \end{figure*} Selecting a sample of hot clusters leads to a significant flattening of the slope of the relation, slightly flatter than the self-similar prediction of $5/3$. With the gas mass-total mass relations of the hot sample and relaxed subset being very close to self-similar, the shallower than self-similar slope is due to the temperature-mass relation. The best-fit slope of both samples shows no significant trend with redshift. The scatter about the best-fit relation is independent of both mass and redshift for all three samples, but it is noisy. We find an average value of $0.12$ at $z=0$ for the scatter for the combined sample, $0.11$ for the hot cluster sample and $0.08$ for the relaxed subset. These values are larger than those found previously for both simulations, where values of $0.04$ \citep{Short2010}, $0.08$ \citep{Planelles2014} and $0.04$ \citep{LeBrun2016} were found, and observations, where a value of $0.04$ was found for a sample of clusters observed with \textit{XMM} \citep{Arnaud2007}. \subsection{$Y_{\rm{SZ}}$-Total Mass} The integrated SZ effect-total mass relations for the three samples are shown in Fig. \ref{fig:YszMsr}. Both the integrated SZ signal and its X-ray analogue measure the total energy of the hot gas in the ICM, however the SZ signal depends on the mass-weighted temperature rather than the X-ray spectral temperature. Our best-fit relation for the combined sample shows a mild evolution with redshift, with clusters at $z=1$ yielding an integrated signal that is $27\%$ higher than clusters at $z=0$ for a fixed mass. The evolution reflects the evolution in the gas mass-total mass relation. The increased evolution of its normalization compared to its X-ray analogue suggests that the normalization of the mass-weighted temperature evolves more self-similarly then the spectroscopic X-ray temperature and is indeed confirmed by the study of the mass-weighted temperature-total mass relation. Selecting a sample of hot clusters or a relaxed subset of them significantly reduces the evolution in the normalization. The normalization of both samples, within the uncertainty of the fits, evolves in agreement with the self-similar prediction. Selecting a hot sample leads to a $25\%$ higher normalization than the combined sample at $z=0$, due to the flatter slope of the gas mass-total mass relation yielding a flatter $Y_{\rm{SZ}}$ slope and a higher normalization at the pivot point. The best-fit relation for the combined sample produces a slope of $\alpha=1.88^{+0.02}_{-0.04}$ at $z=0$, which is significantly steeper than the $5/3$ value predicted by the self-similar model. The value for the slope of the relation is consistent with previous values from both simulations, where values of $1.825\pm0.003$ \citep{Stanek2010}, $1.71\pm0.03$ \citep{Battaglia2012}, $1.74\pm0.01$ \citep{Planelles2014}, $1.70\pm0.02$ \citep{Pike2014}, $1.68\pm=0.05$ \citep{Yu2015} and $1.94$ \citep{LeBrun2016} have been found, and observations, where $1.79\pm0.08$ was found for the Planck clusters \citep{Planck2014XX} and $\alpha=1.77\pm0.35$ was found for the clusters in the $2500\,\rm{deg}^2$ SPT survey. The steeper than self-similar slope is the result of the gas mass-total mass relation having a steeper slope. We find that the slope of the relation is independent of redshift. \begin{figure*} \includegraphics[width=\textwidth]{Lx-M_ScalRel_spec.png} \caption{Evolution of the core-excised bolometric X-ray luminosity-total mass scaling relation as a function of redshift for the three samples of clusters. The panels are arranged as described in Fig. \ref{fig:MgMsr}.} \label{fig:LxMsr} \end{figure*} The best-fit slopes of the hot cluster sample and the relaxed subset are consistent with the slope predicted by self-similar theory. The slopes of both samples are consistent with no evolution. The scatter of the clusters about the best-fit relation shows no trend with either mass or redshift for all three samples. We find an average scatter of $\sigma_{\log_{10}Y}=0.10$, $0.10$ and $0.09$ for the combined, hot and relaxed samples, respectively, at $z=0$. This is larger than the scatter reported by previous simulations, where \citet{Battaglia2012}, \citet{Pike2014}, \citet{Planelles2014} and \citet{LeBrun2016} found values of $0.06$, $0.03$, $0.07$ and $0.04$ respectively, but in reasonable agreement with the values of $0.12\pm0.03$ and $0.08$ observed by \citet{Yu2015} and \citet{Planck2014XX} respectively. \subsection{Bolometric X-ray Luminosity-Total Mass} Fig. \ref{fig:LxMsr} shows the core-excised bolometric X-ray luminosity-total mass scaling relations for the three samples and their evolution with redshift. The normalization of the best-fit relation for the combined sample shows significant evolution with redshift, being $80\%$ higher at $z=1$ compared to $z=0$. The same physics driving the gas mass-total mass relation, increased binding energy, is driving the departure from self-similar. The X-ray emission of a cluster is particularly sensitive to the thermal structure of the ICM, which depends on processes such as radiative cooling and feedback. Therefore, it is not surprising that the luminosity-mass relation shows significantly more evolution than other observable-mass relations. Selecting a sample of hot clusters significantly reduces the evolution in the normalization. Both the hot sample and the relaxed subset have a normalization that is $\approx60\%$ larger at $z=0$ compared to the combined sample. The deeper potentials of more massive clusters reduces the impact of the AGN feedback and flattens the relation. This flattening leads to a higher luminosity at the pivot point. The normalizations of the best-fit relations for both the hot sample and its relaxed subset show very minor evolution, which is consistent with the self-similar prediction. \begin{figure*} \includegraphics[width=\textwidth]{Lx-Tx_ScalRel_spec.png} \caption{Evolution of the bolometric X-ray luminosity-X-ray temperature scaling relation for the three samples as a function of redshift. The panels are arranged as described in Fig. \ref{fig:MgMsr}.} \label{fig:LxTxsr} \end{figure*} The slope of the best-fit relation for the combined sample is significantly steeper than the $4/3$ slope predicted by self-similar theory. At $z=0$ we find a slope of $\alpha=1.88^{+0.03}_{-0.05}$ for the combined sample. This steepening is driven by AGN feedback being more effective in lower mass clusters. The slope at $z=0$ is in reasonable agreement with the slopes found in volume-limited observational samples, such as \citet{Pratt2009} who found a slope of $1.80\pm0.05$ for the REXCESS sample and \citet{Giles2015} who found a slope of $2.14\pm0.21$ for a sample of $34$ low-redshift clusters. Previous simulation work by \citet{Short2010}, using the semi-analytic feedback model of the Millennium Gas project, found a bolometric luminosity-total mass slope of $1.77\pm0.03$ and \citet{Stanek2010}, using the preheating model of the Millennium Gas project, found a slope of $1.87\pm0.01$. \citet{Biffi2014} found a slope of $1.45\pm0.05$ for the MUSIC simulations. The slope of the best-fit relation for the combined sample is approximately independent of redshift, with a very mild steepening of the slope with redshift occurring due to the reduction in fitting range with increasing redshift. The slopes of the best-fit relation follow the same trend as the gas mass-total mass relation, with the hot sample and its relaxed subset producing shallower slopes that are in much better agreement with self-similar theory. Our best-fit slope is consistent with the observational result of \citet{Mantz2016}, who found a self-similar slope for the core-excised luminosity-total mass relation for a sample of $40$ relaxed clusters with $k_{\rm{B}}T\geq\,5\,\rm{keV}$. The scatter about the best fit relation is approximately independent of both mass and redshift for all three samples, although it is relatively noisy. Averaging the scatter for the combined sample across all mass bins produces a value of $\sigma_{\log_{10}Y}=0.15$. This is in reasonable agreement with the scatter found in low-redshift observational samples, where \citet{Pratt2009} find a value of $0.17\pm0.03$ and \citet{Giles2015} find a value of $0.22\pm0.03$. Selecting hot clusters and a relaxed subset produces a small reduction in scatter about the best-fit relation with values of $0.12$ and $0.11$ respectively. \subsection{X-ray Luminosity-Temperature} Finally, we study the redshift evolution of the X-ray luminosity-spectroscopic temperature relation. Both quantities of the luminosity-temperature scaling relation are observable, with the temperature tracing the depth of the potential of the cluster. This makes it a useful relation to study the impact of non-gravitational physics. In Fig. \ref{fig:LxTxsr} we plot the bolometric X-ray luminosity-spectroscopic temperature scaling relation for the three samples of clusters. The normalization of the best-fit relation for the combined sample shows significant evolution with redshift relative to self-similar. Clusters with a temperature of $6\,\rm{keV}$ at $z=1$ have a luminosity $94\%$ greater than clusters with the same temperature at $z=0$. This evolution can be thought of as being due to a combination of the evolution of the temperature-mass and luminosity-mass relations. The decreasing temperature-mass normalization and increasing luminosity-mass normalization with redshift combine to yield a significant evolution of the luminosity-temperature normalization relative to self-similar. Selecting a sample of hot clusters, or a relaxed subset of them, reduces the evolution, but there is still a mild evolution in the normalization. Hot clusters at a fixed temperature at $z=0.5$ are $\approx15\%$ more luminous than those at $z=0$. Combining equations (\ref{eq:T-M}) and (\ref{eq:L-M}), but allowing the slope of the relations to vary from their self-similar values yields \begin{equation} \label{eq:Lx-Txevo} L_{\rm{X},\Delta}^{\rm{bol}}\propto T^{\alpha_{\rm{LM}}/\alpha_{\rm{TM}}}E^{7/3-2\alpha_{\rm{LM}}/3\alpha_{\rm{TM}}}(z)\,, \end{equation} where $\alpha_{\rm{LM}}$ and $\alpha_{\rm{TM}}$ are the slopes of the luminosity-mass and temperature-mass relations respectively. Hence, deviations of their slopes from self-similar leads to evolution of the normalization of the luminosity-temperature relation that is not self-similar. With the luminosity-mass relation being self-similar for the hot cluster sample and its relaxed subset, the evolution of the normalization is being driven by the flatter than self-similar slope of the temperature-mass relation, which is due to the increased importance of non-thermal pressure support and the increasingly biased spectroscopic temperatures of more massive clusters. We find a slope of $\alpha=3.01\pm0.04$ for the best-fit relation at $z=0$. This is significantly steeper than the slope of $2$ predicted by self-similar theory. However, this value is reasonable agreement with previous simulation work, $3.30\pm0.07$ \citep{Short2010}, and those found by observations, $2.95\pm0.15$ for the REXCESS sample \citep{Pratt2009} and $\alpha=3.63\pm0.27$ for a sample of $114$ clusters observed with \textit{Chandra} \citep{Maughan2012}. It is clear from equation (\ref{eq:Lx-Txevo}) that the slope of the relation depends on the slopes of the luminosity-mass and temperature-mass relations. The steeper than expected slope for the combined sample is due to the combined effects of AGN feedback on the luminosity slope and non-thermal pressure support and temperature bias on the temperature slope, both of which lead to a steepening of the relation compared to the self-similar prediction. We find that the best-fit relation steepens slightly with redshift, increasing to $3.35\pm0.07$ at $z=1$. This evolution is due to the removal of high-mass objects with redshift. The best-fit slope of the hot cluster sample and the relaxed subset are flatter than the combined relation with slopes of $2.41\pm0.12$ and $2.53\pm0.13$. This is still significantly steeper than the slope predicted by self-similar theory, but in good agreement with the slope of $2.44\pm0.43$ observed by \citet{Maughan2012} for their relaxed cool core cluster sample. With both samples exhibiting self-similar slopes for the luminosity-mass relations, the deviation from self-similarity is being driven by their temperature-mass relations. The scatter about the best-fit relation demonstrates a trend with both temperature and redshift. Although somewhat noisy, the scatter appears to increase with decreasing temperature. The average scatter at $z=0$ for the combined sample is $\sigma_{\log_{10}Y}=0.14$ . This scatter is consistent with the simulations of \citet{Short2010}, who found a scatter $0.10$, and the intrinsic observational scatter of $0.12$ found by \citet{Pratt2009}. However, it is significantly lower than the scatter of $0.29$ found by \citet{Maughan2012}. The scatter reduces for the hot cluster sample and the relaxed subset to $0.11$ and $0.10$ respectively. \subsection{Summary} Overall, the scaling relations of the combined sample show good agreement with previous work, both simulations and observations. Departures from self-similarity are driven by the increased efficiency of gas expulsion by AGN feedback in clusters with shallower potentials, due to being less massive or forming at a lower redshift; the increased contribution of non-thermal pressure that supports the ICM against gravity in more massive clusters or those at higher redshifts; and the increase in the spectroscopic temperature bias for the most massive clusters. The MACSIS sample enabled the scaling relations to be studied to higher redshifts, as their progenitors are still clusters at high redshift, and the examination of the impact of selecting a sample of hot clusters on the evolution of the scaling relations. This demonstrated that massive clusters are more self-similar and evolve more self-similarly with redshift compared to the overall cluster population, as the efficiency of gas expulsion by AGN feedback is reduced due to their deeper potentials. However, it also highlighted that non-thermal pressure support becomes more important in these clusters and that their spectroscopic temperatures are biased low. \section{Evolution of gas profiles} \label{sec:gasprofs} Most of the scaling relations of hot, and therefore massive, clusters evolve in a way that is consistent with the predictions of the self-similar model. However, the combined sample showed significant deviations from the self-similar model due to the impact of non-gravitational processes. To further understand the differences between the samples in the evolution of their scaling relations, we now examine the gas profiles of the different cluster samples. To enable a quantitative comparison with the observational data requires us to compare like-with-like. Therefore, we restrict the mass range of the combined sample to $2.0\times10^{14}\,\mathrm{M}_{\odot}\leq M_{500,\mathrm{spec}}\leq1.0\times10^{15}\,\mathrm{M}_{\odot}$, yielding a sample with a median mass of $2.44\times10^{14}\,\mathrm{M}_{\odot}$. We compare this to the REXCESS cluster sample which has a median mass of $2.68\times10^{14}\,\mathrm{M}_{\odot}$ and a sample of clusters from \citet{Giles2015} with a median mass of $5.43\times10^{14}\,\mathrm{M}_{\odot}$. Although this mass matching does not account for selection effects, it should allow for a quantitative comparison. We do not alter the hot sample or the relaxed subset. We factor out the expected self-similar evolution in the profiles by dividing by the appropriate quantity, e.g. $\rho_{\rm{crit}}$, $k_{\rm{B}}T_{500}$, $P_{500}$ or $K_{500}$. We define these quantities as \begin{equation} \rho_{\rm{crit}}(z)\equiv E^2(z)\frac{3H^2_0}{8\pi G}\,, \end{equation} \begin{equation} k_{\rm{B}}T_{500}=\frac{GM_{500}\mu m_{\rm{p}}}{2r_{500}}\,, \end{equation} \begin{equation} P_{500}=500f_{\rm{b}}k_{\rm{B}}T_{500}\frac{\rho_{\rm{crit}}}{\mu m_{\rm{p}}}\,, \end{equation} \begin{equation} K_{500}=\frac{k_{\rm{B}}T_{500}}{\left(500f_{\rm{b}}(\rho_{\rm{crit}}/\mu_{\rm{e}}m_{\rm{p}})\right)^{2/3}}\,, \end{equation} where $H_0$ is the Hubble constant, $G$ is the gravitational constant, $\mu_{\rm{e}}$ is the mean atomic weight per free electron and $f_{\rm{b}}=\Omega_{\rm{b}}/\Omega_{\rm{m}}$ is the universal baryon fraction. Therefore, any changes in the profiles are due to non-gravitational physics, such as AGN feedback or non-thermal pressure support. \subsection{Density profiles} \begin{figure} \includegraphics[width=\columnwidth]{Density_spec.png} \caption{Median gas density profiles for the combined (grey dash-dot), hot (dark red dashed) and relaxed hot (red solid) samples at $z=0$ (top panel) and $z=1$ (middle panel), scaled by $(r/r_{500,\rm{spec}})^2$ to reduce dynamic range. The grey hatched region shows the $16^{\rm{th}}$ to $84^{\rm{th}}$ percentiles of the combined sample profile. Overlaid as black squares, triangles and circles are the median observed profiles from the REXCESS sample \citep{Croston2008}, a sample of low redshift clustes observed with \textit{Chandra} \citep{Giles2015} and a high-redshift, SPT-selected sample \citep{McDonald2013,McDonald2014} respectively, with the error bars showing the $16^{\rm{th}}$ and $84^{\rm{th}}$ percentiles. The bottom panel shows the $\log_{10}$ of the ratio of the profiles at $z=0$ and $z=1$ for each sample.} \label{fig:gas_prof} \end{figure} The three-dimensional dimensionless density profiles for the three cluster samples at $z=0$ and $z=1$ are shown in Fig. \ref{fig:gas_prof}. We have scaled the profiles by $r^2$ to reduce the dynamic range. At $z=0$, we compare the median profile of the combined sample with the observed median profiles from \citet{Croston2008} for the REXCESS sample and \citet{Giles2015} for a sample of low-redshift clusters observed with \textit{Chandra}. The combined sample shows good agreement with the observed profiles and has similar intrinsic scatter. Beyond a radius of $0.15r_{500,\rm{spec}}$ the median profiles of the hot sample and its relaxed subset have a similar shape as the combined sample, but the densities are higher as they are on average more massive clusters. Inside this radius the profiles of both samples have a shallower gradient compared to the combined sample. This is caused by the accretion of low-entropy, high-density gas that sinks to the centre of the cluster potential, becoming increasingly important below $z=1$ \citep{Power2014}. This effect is not offset in massive clusters by the AGN feedback and so their density profiles have a shallower gradient in the core. We note that this effect can potentially impact the relations we presented in section \ref{sec:screlations}. However, we presented core-excised temperatures and luminosities, which should minimise any bias introduced by the accretion of poorly mixed gas. At $z=1$, we compare the median density profiles of the three samples to the observed profile from \citet{McDonald2013}, which has been derived from a sample of $40$ clusters with a mean redshift of $z=0.82$. These clusters were selected from the SPT $2500\,\rm{deg}^2$ survey catalogue and observed with \textit{Chandra}. There is a reasonable agreement between the combined sample's median profile and the observations, but the observations are higher between $0.2-1.0r_{500}$. The observed profile is in better agreement with the median profiles of the hot sample and its relaxed subset. This suggests that the observed clusters are more representative of more massive objects at $z=1$. There is better agreement between the density profiles of the three samples at $z=1$ because the mass cut of $M=10^{14}\,\rm{M}_{\odot}$ causes the samples to converge with increasing redshift. Selecting relaxed hot clusters leads to a median profile that is slightly more centrally concentrated than for all hot clusters. In the bottom panel of the plot we show the $\log_{10}$ of the ratio of the median density profile at $z=0$ and the median profile at $z=1$ for each sample. For the hot cluster sample and the relaxed subset the profiles have evolved in a self-similar way beyond $0.2\,r_{500}$, showing very little change. Inside of this radius the impact of accreting low-entropy, high-density gas that sinks to the centre of the cluster is apparent as an increase in the density profiles from $z=1$ to $z=0$. For the combined sample the difference between the two profiles shows the increase of the depth of the potential with redshift. This leads to higher densities at $z=1$ and a negative change density profile at all radii with decreasing redshift. \subsection{Temperature profiles} \begin{figure} \includegraphics[width=\columnwidth]{Temperature_spec.png} \caption{Median temperature profiles for the three samples. The details are the same as for Fig. \ref{fig:gas_prof}, except that the REXCESS data was taken from \citet{Arnaud2010}.} \label{fig:temp_prof} \end{figure} Fig. \ref{fig:temp_prof} shows the three-dimensional temperature profiles divided by the predicted self-similar temperature. At $z=0$ the profiles all have a similar shape, but the normalization of the combined sample is somewhat higher than those of the hot sample and its relaxed subset. This is due to the lower gas density of the combined sample, which requires a higher temperature to balance gravitational collapse. Also, there is likely to be a small effect due to the mass dependence of non-thermal pressure support, with more massive clusters having more non-thermal support and lower temperatures. The accretion of low-entropy, cold gas that sinks to the cluster core produces a steeper temperature gradient in the central profiles of the hot sample and its relaxed subset. Overlaid are the observed median temperature profiles from two cluster samples, the REXCESS sample \citep{Arnaud2010} and a sample of clusters observed with \textit{Chandra} \citep{Giles2015}. The median profile of the combined sample and its intrinsic scatter show good agreement with the observed temperature profiles and their scatter. At $z=1$ all samples have a similar profile shape, but the hot sample has a lower normalization compared to the combined and relaxed hot sample. This is because non-thermal pressure support becomes increasingly important in clusters of a fixed mass with redshift, leading to a lower temperature in hot clusters. The relaxed sample removes the most disturbed objects with greatest level of non-thermal support, producing a higher median temperature profile. We compare to the observed median profile of \citet{McDonald2014}. The median profiles of the combined sample and the relaxed hot sample slightly under predict the observations at $0.3\,\rm{r}_{500}$ and over predict the observations at large radii, but the observed profile is within the scatter of the combined sample. Within $r_{500,\rm{spec}}$ the median temperature profiles show significantly less evolution between the two redshifts than the density profiles. The combined and hot samples deviate from self-similarity and show an increase in temperature from $z=1$ to $z=0$ at all radii, consistent with the decreasing temperature-mass normalization with increasing redshift found in Section \ref{ssec:Tx-M}. This is because non-thermal pressure support decreases with increasing redshift. Therefore, as clusters thermalise their temperatures must increase to balance gravitational collapse, resulting in a hotter temperature profile at $z=0$ compared to $z=1$. Selecting a relaxed subset reduces the non-thermal pressure support and the median profile changes significantly less from $z=1$ to $z=0$ inside $r_{500}$. \subsection{Pressure profiles} \begin{figure} \includegraphics[width=\columnwidth]{Pressure_spec.png} \caption{Median pressure profiles for the three samples. The details are the same as for Fig. \ref{fig:gas_prof}, except that the REXCESS data was taken from \citet{Arnaud2010}. The green curve shows the best-fit pressure profile from \citet{Planck2013}.} \label{fig:pres_prof} \end{figure} The dimensionless pressure profiles, scaled by $r^{3}$, of the three cluster samples are shown in Fig. \ref{fig:pres_prof}. The increased mass of the hot sample and its relaxed subset lead to median pressure profiles that are higher in the centre at $z=0$ due to their higher densities. We compare the median profiles to the observed median pressure profiles from \citet{Arnaud2010} and \citet{Giles2015} and the best-fit profile from \citet{Planck2013}. We note that the \textit{Planck} result is based on the stacked profile of nearby systems. For \citet{Giles2015} we have combined their published density and temperature profiles to produce a pressure profile for each cluster. There is good agreement between the combined sample and the observed profiles, with a slight over prediction at large radii. For comparison to the Planck best-fit parameters we fit the mean profiles of our clusters at both redshifts with a generalised Navarro-Frenk-White pressure profile \citep{NFW1997,NagaiVikhlininKravtsov2007a} of the form \begin{equation} P(x)=\frac{P_{0}}{(c_{500}x)^{\gamma}\left[1+(c_{500}x)^{\alpha}\right]^{(\beta-\gamma)/\alpha}}\:. \label{gNFWeq} \end{equation} We fit a four parameter model with $\gamma=0.31$ fixed. The results are shown in table \ref{tab:Ppro_fit}. \begin{table} \caption{Table showing the best-fit generalised Navarro-Frenk-White pressure profile parameters (see eq. \ref{gNFWeq}) for the combined, hot and relaxed hot samples of clusters present in this work. We fix $\gamma=0.31$.} \centering \begin{tabular}{c l r c c c} \hline $z$ & Sample & \multicolumn{1}{c}{$P_{0}$} & $c_{500}$ & $\alpha$ & $\beta$ \\ \hline & \textit{Planck} & $6.41$ & $1.81$ & $1.33$ & $4.13$ \\ \hline $0$ & Combined & $8.80$ & $1.56$ & $1.09$ & $4.01$ \\ & Hot & $20.66$ & $0.52$ & $0.70$ & $6.69$ \\ & Relaxed Hot & $24.01$ & $0.54$ & $0.69$ & $6.79$ \\ \hline $1$ & Combined & $6.96$ & $0.99$ & $1.26$ & $5.84$ \\ & Hot & $6.44$ & $0.51$ & $1.14$ & $9.44$ \\ & Relaxed Hot & $9.28$ & $1.97$ & $1.61$ & $4.11$ \\ \hline \end{tabular} \label{tab:Ppro_fit} \end{table} At $z=1$ the median profiles of the three samples are in closer agreement with each other, because the minimum mass limit of $M=10^{14}\,\rm{M}_{\odot}$ causes the samples to converge at high redshift. We compare our median pressure profiles with the observed profile of \citet{McDonald2014}. They find a median pressure profile that is in good agreement with the median profiles, but it is most consistent with the relaxed hot sample of massive clusters. The pressure profile of the relaxed subset shows very little evolution between $z=1$ and $z=0$, except for the core where the increasing density leads to an increased pressure with decreasing redshift. The hot sample shows an increased pressure in the core with decreasing redshift, due to the increased density, but a negative change in pressure from $z=1$ to $z=0$ at larger radii. The combined sample shows a negative pressure change between $z=1$ and $z=0$ at all radii. The decreased pressure with decreasing redshift is caused by the decrease in density from $z=1$ to $z=0$. \subsection{Entropy Profiles} \begin{figure} \includegraphics[width=\columnwidth]{Entropy_spec.png} \caption{Median entropy profiles for the three samples. The details are the same as Fig. \ref{fig:gas_prof}, except that the REXCESS data was taken from \citet{Pratt2010}. We also show the prediction from non-radiative simulations for $z=0$ \citep{VoitKayBryan2005}.} \label{fig:enty_prof} \end{figure} The median entropy profiles are shown in the bottom right panel of Fig. \ref{fig:enty_prof} and they have been normalized by the predicted self-similar entropy. We note that we define entropy as \begin{equation} K_{\Delta}\equiv\frac{k_{\rm{B}}T_{\Delta}}{n^{2/3}_{e,\Delta}}\, \end{equation} where $n_e$ is the electron number density and $\Delta$ is the chosen overdensity relative to the critical density of the Universe. At $z=0$ the the combined sample shows a higher normalization compared to the hot sample and its relaxed subset. This is due to it lower density profile and higher temperature profile. The gradients of the hot sample and the relaxed subset profiles steepen in the centre due to the accretion of low entropy gas. We compare with the observed median profiles of \citet{Pratt2010} and \citet{Giles2015}, and the baseline profile of \citet{VoitKayBryan2005} derived from non-radiative SPH simulations. The combined sample is in good agreement with the observations and tends to the non-radiative predictions at large radii. At $z=1$ the three samples are in reasonable agreement with each other, all having a similar shape with the hot sample showing a marginally lower normalization. This change from $z=0$ is in agreement with the evolution in their density and temperature profiles. We compare the profiles to the observations of \citet{McDonald2014}. The combined and relaxed hot sample show good agreement with the observed profile for $r<0.5r_{500,\rm{spec}}$, but over predict the entropy at larger radii. In contrast the median profile of the hot sample is consistent with the observations at large radii, but under predicts the entropy in the centre of the cluster. The departure from self-similarity for the three samples is due to a combination of the evolution in their temperature and density profiles. The relaxed hot sample shows a mild increase in entropy from $z=1$ to $z=0$ at large radii, due to change in its temperature profile, and a decrease in entropy in the core due to the increase in density at $z=0$. The increased normalization of the hot sample's temperature profile at $z=0$ compared to $z=1$ leads to an increased entropy profile with decreasing redshift, except in the core. The combined sample shows an increase in entropy at all radii at $z=0$ compared to $z=1$ and is produced by the decreased density and increased temperature with decreasing redshift. \section{Summary \& Discussion} \label{sec:sad} In this work we have presented the MACSIS clusters, a sample of 390 zoomed simulations of the most massive and rarest clusters run with the state-of-the-art, calibrated baryonic physics model from the BAHAMAS project \citep{McCarthy2016} that yields realistic clusters. Such massive clusters are absent from the BAHAMAS simulation volumes of $596\,\mathrm{Mpc}$ as the simulated volume is too small. After introducing the selection of the sample from the parent $3.2\,\mathrm{Gpc}$ volume simulated with the \textit{Planck} 2013 cosmology, and demonstrating the agreement of the properties of our massive cluster sample with the properties of observed massive clusters, we examined the evolution of the cluster scaling relations and the evolution of the cluster gas profiles. By combining the MACSIS sample with the clusters in the BAHAMAS volume, we were able to examine the cluster scaling relations over the full observed mass range for the first time. Additionally, the MACSIS clusters enabled the study of the evolution of the cluster scaling relations to unprecedentedly high redshifts. Finally, the MACSIS sample enabled clusters to be selected in ways which mimic a cosmological study, such as selecting the hottest clusters, to examine if the scaling relations of such objects evolve differently from the underlying cluster population. Our main results are: \begin{itemize} \item As shown in Fig. \ref{fig:observations}, the MACSIS simulations yield realistic massive clusters at low redshift and their progenitors are in good agreement with the limited observational data that is available at high redshift (i.e. $z=1$). \item Scaling relations for the combined sample that spans the full observed cluster mass range show significant deviations from the simple self-similar theory (see Figs. \ref{fig:MgMsr}-\ref{fig:LxTxsr}). Both the slope of the relations and the redshift evolution of the normalization are significantly affected by non-gravitational physics. The low redshift relations are in good agreement with observations and with most previous simulation work. \item The main drivers of non-self-similar evolution are AGN feedback, non-thermal pressure support and a mild mass dependence of the spectroscopic temperature bias. Shallower potentials of clusters that are less massive or form at lower redshifts allows feedback from AGN to eject more gas. Non-thermal pressure lowers a cluster's temperature for a given potential and is more important in more massive clusters that have had less time to thermalise. We found that the spectroscopic temperature bias increases for the most massive clusters. \item With the exception of the luminosity-temperature relation, we found the scatter about the best-fit scaling relations is insensitive to mass and redshift for all of the cluster samples. \item Selecting a hot cluster sample, i.e. core-excised spectroscopic temperatures $k_{\rm{B}}T^{\mathrm{X,ce}}_{500\mathrm{spec}}\geq5\,\rm{keV}$, significantly alters the scaling relations and their evolution. Excluding the spectroscopic temperature-total mass relation, we find that the scaling relations of the hot cluster sample evolve in a much more self-similar manner. After accounting for the expected self-similar evolution with redshift, we find that the normalizations are consistent with no evolution. The slopes of the best-fit relations at each redshift are also broadly consistent with the slopes predicted by self-similar theory. However, the spectroscopic temperature-total mass relation of the hot sample deviates further from self-similarity than the combined sample. Selecting hot clusters removes the less massive clusters from the sample, so the hot sample is dynamically younger than the combined sample as more massive clusters form later in the hierarchical merger scenario. This increases the average level of non-thermal support in the hot sample, leading to a flatter spectroscopic temperature-total mass relation. Additionally, the spectroscopic temperature bias flattens the relation for the most massive clusters and this has a larger impact in a sample of only hot clusters. \item Selecting a relaxed subset of hot clusters, where the most dynamically disturbed objects are removed, leads to a small reduction in the scatter for most scaling relations. Removing the most disturbed objects also leads to a reduction in the level of non-thermal support in the sample compared to the complete hot sample. This leads to steeper slope of the spectroscopic temperature-total mass relation compared to the hot sample and a value that is closer to the self-similar prediction. \item The median hot gas profiles of the combined sample in general shows good agreement with observed radial profiles. The low redshift data is in very good agreement, while the data at $z=1$ shows reasonable agreement with the relaxed hot sample. \item Comparison of the hot gas profiles at $z=0$ and $z=1$ show evolution different from self-similar prediction (see Figs. \ref{fig:gas_prof}-\ref{fig:enty_prof}). The combined sample shows a decreasing density profile with decreasing redshift, suggesting the impact of AGN feedback. Selecting a sample of hot clusters produces a median density profile that evolves in much more self-similar manner. The combined and hot samples have a median temperature profile that increases with decreasing redshift. This is likely driven by decreasing importance of non-thermal pressure support with decreasing redshift. Selecting relaxed hot cluster sample produces a median profile that evolves in better agreement with the self-similar prediction. \end{itemize} MACSIS enables the study of the observable properties of the most massive and rarest galaxy clusters. We have demonstrated that their progenitors provide a good match to the currently limited observational data at high redshift and that their observable properties evolve in a significantly more self-similar manner than for lower-mass and less-relaxed clusters. We have shown how the selection function can impact the derived scaling relations and radial profiles. The size of the parent simulation enables the creation of synthetic lightcones with an area comparable to currently ongoing surveys. This will allow the impact of selection biases to be fully examined and the covariance of observable properties to be studied. Another route for future work is to improve our understanding of structure in the ICM, as the limited resolution and traditional SPH scheme used in this work limits our ability resolve structures and understand their impact on observable properties. \section*{Acknowledgements} This work used the DiRAC Data Centric system at Durham University, operated by the Institute for Computational Cosmology on behalf of the STFC DiRAC HPC Facility (www.dirac.ac.uk). This equipment was funded by BIS National E-infrastructure capital grant ST/K00042X/1, STFC capital grants ST/H008519/1 and ST/K00087X/1, STFC DiRAC Operations grant ST/K003267/1 and Durham University. DiRAC is part of the National E-Infrastructure. DJB and STK acknowledge support from STFC through grant ST/L000768/1. MAH is supported by an STFC quota studentship. IGM is supported by a STFC Advanced Fellowship. The research was supported in part by the European Research Council under the European Union's Seventh Framework Programme (FP7/2007-2013) / ERC Grant agreement 278594-GasAroundGalaxies. ARJ acknowledges support from STFC through grant ST/L00075X/1. \bibliographystyle{mnras}
78e828700db98d40569e3b8ade9cfcb556414852
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{intro} Converse and saturation results are an important part in regularization theory for solving ill-posed problems. Related results for Tikhonov regularization were developed many years ago and are well known, see Groetsch~\cite[Chapter 3]{Groetsch[84]} and the references therein, or Neubauer~\cite{Neubauer[97]}. In the present paper we show that similar results can be obtained for \lavmet{}\xspace when accretive\xspace linear bounded, and possibly non-selfadjoint, operators on Hilbert spaces are involved. Our work is inspired by the two papers \cite{Neubauer[97],Raus[84]}. At several steps, however, the technique used in the present paper differs substantially from the one used in the two papers \cite{Neubauer[97],Raus[84]} since no spectral decomposition is available in our setting, in general. As a byproduct we obtain a new result on the optimality of a posteriori parameter choices for \lavmet{}\xspace. We start more generally with the consideration of equations on Banach spaces, i.e., \begin{align} A u = f, \label{eq:maineq} \end{align} where $ A: \mathcal{X} \to \mathcal{X} $ is a bounded linear operator on a real or complex Banach space $ \mathcal{X} $ with norm $ \norm{\cdot} $, and $ f \in \mathcal{R}(A) $. Our focus is on operators having a non-closed range $ \mathcal{R}(A) $ which in fact implies that the considered equation \refeq{maineq} is ill-posed. Note, however, that this range condition will be explicitly stated in this paper whenever needed. In the sequel we restrict the considerations to the following class of operators: \begin{definition} \label{th:postype-def} A bounded linear operator $ A: \mathcal{X} \to \mathcal{X} $ on a Banach space $ \mathcal{X} $ is called \emph{\postype{}\xspace}, if for any parameter $ \gamma > 0 $ the operator $ A + \gamma I: \mathcal{X} \to \mathcal{X} $ has a bounded inverse on $ \mathcal{X} $, and \begin{align} \norm{(A + \gamma I)^{-1} } \le \tfrac{ M }{\gamma} \quad \text{for } \gamma > 0, \label{eq:postype} \end{align} holds, with some constant $ M \ge 1 $ that is independent of $ \gamma $. \end{definition} The notation ``nonnegative\xspace'' is introduced by Komatsu~\cite{Komatsu[69.1]}; see also Martinez/Sanz~\cite{Martinez_Sanz[00]}. In many papers, no special notation is used for property \refeq{postype}. \begin{example} \label{th:integration_abel_example} Prominent examples of nonnegative\xspace operators are given by the classical integration operator $ (Vu)(x) =\int_{0}^{x} u(y) dy $ for $ 0 \le x \le 1 $, and the Abel integral operators $ (V^\alpha u)(x) = \frac{1}{\Gamma(\alpha)} \int_{0}^{x}{(x-y)^{-(1-\alpha)} u(y) }{dy} $ for $ 0 \le x \le 1 \ (0 < \alpha < 1) $. Both operators $ V $ and $ V^\alpha $ are considered either on the space of functions $ \mathcal{X} = L^p(0,1) $ with $ 1 \le p \le \infty $, or the space of continuous functions $ \mathcal{X} = C[0,1] $. See, e.g. \cite[Section 1.3]{Plato[95]} for details. \quad $ \vartriangle $ \end{example} For the regularization of the considered equation $ Au = f $ with a \postype bounded linear\xspace operator $ A $, we consider Lavrentiev's method \begin{align} (A + \gamma I) \uparadel = f^\delta, \label{eq:lavmet} \end{align} where $ \gamma > 0 $ is a regularization parameter. In addition we have \begin{align} f^\delta \in \mathcal{X}, \quad \norm{ f - f^\delta } \le \delta, \label{eq:noisy_data} \end{align} where $ \delta > 0 $ is a given noise level. We next consider fractional powers of the operator $ A $ that may serve as a tool to describe smoothness of solutions for equation \refeq{maineq}. \begin{definition} \label{th:frac-power} For $ 0 < p < 1 $, the \emph{fractional power} $ A^p: \mathcal{X} \to \mathcal{X} $ of a nonnegative\xspace bounded linear operator $ A: \mathcal{X} \to \mathcal{X} $ on a Banach space $ \mathcal{X} $ is given by (see, e.g., Kato~\cite[formula (12)]{Kato[62]}) the improper operator-valued integral \begin{align} A^p := \frac{\sin \pi p }{\pi} \int_0^\infty s^{p-1} (A + sI)^{-1} A \ ds, \quad 0 < p < 1 . \label{eq:frac-power} \end{align} For arbitrary values $ p > 0 $, the fractional power $ A^p $ of the operator $ A $ is defined by $ A^p := A^{p - \lfloor p \rfloor} A^{\lfloor p \rfloor}, $ where $ \lfloor p \rfloor $ denotes the largest integer which does not exceed $p$. \quad $ \vartriangle $ \end{definition} For each $ 0 < p < 1 $, the identity \refeq{frac-power} defines a bounded linear operator $ A^p: \mathcal{X} \to \mathcal{X} $. In inverse problems, smoothness of a solution $ u $ of equation \refeq{maineq} is often described in the form $ u \in \mathcal{R}(A^p)$ for some $ p > 0 $. This allows to deduce convergence rates for Lavrentiev regularization\xspace in the case of noise-free data (with respect to $ \gamma $) as well as in the case of noisy data (in terms of $ \delta $ then). The subject of this paper is to present converse and saturation results for those convergence rates. In other terms, the impact of the convergence rates on the smoothness of the solution is considered, and the maximal possible rates are identified. The outline of the paper is as follows. Section \ref{conv_saturation_exdat} deals with converse and saturation results for Lavrentiev regularization\xspace in case of exact data\xspace. In Section \ref{optimality}, a theorem is presented which provides the basis for the converse and saturation results in case of noisy data\xspace. This theorem has also impact on the optimality of parameter choices for Lavrentiev regularization\xspace, and related results are stated in Section \ref{optimality} as a byproduct. In Sections \ref{conv_noisdat} and \ref{saturation_noisdat}, converse and saturation results in case of noisy data\xspace are presented, and Section \ref{auxiliary} serves as an appendix which provides some auxiliary results. The main results of this paper are formulated in Theorems \ref{th:pseuopt=orderopt}, \ref{th:converse_noisy_data_1} and \ref{th:saturation_noisy_data}. \section{Converse and saturation results in case of exact data\xspace} \label{conv_saturation_exdat} \subsection{Introductory remarks} Throughout this section let $ A: \mathcal{X} \to \mathcal{X} $ be a nonnegative\xspace bounded linear operator on a Banach space $ \mathcal{X} $. Our main interest are operators with a non-closed range $ \mathcal{R}(A) $, but nowhere in this section this is explicitly required. Comments on the closed range case can be found at the end of this section, cf.~Remark \ref{th:closed_range_exact_data}. In a first step, and also as preparation for converse and saturation results related with noisy data presented in the following sections, we consider \lavmet{}\xspace in case of exact data, and we also introduce the corresponding approximation error: for any $ u \in \mathcal{X} $ let $ \upara \in \mathcal{X} $ and $ \fpara \in \mathcal{X} $ be given by \begin{align} (A+\gamma I)\upara = f, \qquad \fparau{u} := \fpara := \upara-u, \quad \gamma > 0. \label{eq:upara-epara-def} \end{align} We note that the approximation error $ \fpara $, sometimes also called bias, can be represented as follows: \begin{align} \fpara = -\gamma (A + \gamma I)^{-1}u \quad \textup{ for } \gamma > 0. \label{eq:epara-rep} \end{align} The smoothness of a solution $ u $, given in the form $ u \in \mathcal{R}(A^p)$ for some $ p > 0 $, has impact on the speed of convergence $ \upara \to u $ as $ \gamma \to 0 $. We cite the following well-known result; for a proof, see, e.g., \cite[Example 4.1]{Plato[96]}. \begin{proposition} \label{th:direct_exact_data} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a Banach space $ \ix $. If $ u \in \overline{\mathcal{R}(A)} $ then $ \upara \to u $ as $ \gamma \to 0 $. If moreover $ u \in \mathcal{R}(A^p) $ for some $ 0 < p \le 1 $ then $ \norm{\upara-u} = \mathcal{O}(\gamma^p) $ as $ \gamma \to 0 $. \end{proposition} For recent results on the convergence of Lavrentiev regularization\xspace with adjoint source conditions $ u \in \mathcal{R}((A^*)^p)$ in Hilbert spaces, see, e.g., Hofmann/Kaltenbacher/Resmerita~\cite{Hofmann_Kaltenbacher_Resmerita[16]} and Plato/Hofmann/Math\'{e}~\cite{Plato_Hofmann_Mathe[16]}. Two natural questions arise in the context of Proposition \ref{th:direct_exact_data}: \begin{mylist_indent} \item Are the given convergence results in that proposition optimal, or, in other terms, are the conditions $ u \in \overline{\mathcal{R}(A)} $ and $ u \in \mathcal{R}(A^p) $ considered there also necessary, respectively? \item Is the range of values for $ p $, considered in Proposition \ref{th:direct_exact_data}, maximal? \end{mylist_indent} We show in this section that the answers to those questions are basically affirmative. The related results are called converse and saturation results for Lavrentiev regularization\xspace in case of exact data, respectively. \subsection{Converse results in case of exact data\xspace} We start with three converse results related with exact data. \begin{theorem} \label{th:converse_exact_data_1} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $, and let $ u \in \ix $. If $ \upara \to u $ as $ \gamma \to 0 $, then necessarily $ u \in \overline{\mathcal{R}(A)} $ holds. \end{theorem} \begin{proof} We consider the decomposition $ u = u_{R} + u_{N} $ with $ u_{R} \in \overline{\mathcal{R}(A)}, \, u_{N} \in \mathcal{N}(A) $, see Lemma \ref{th:r-n-decomp} in the appendix. This decomposition yields \begin{align*} \gamma (A + \gamma I)^{-1}u = \gamma (A + \gamma I)^{-1}u_{R} + u_{N} \to u_{N} \quad \textup{as } \gamma \to 0 \end{align*} according to Proposition \ref{th:direct_exact_data}. The assumption of the theorem now implies $ u_{N} = 0 $, and from this the statement of the theorem follows. \end{proof} \begin{theorem} \label{th:converse_exact_data_2} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $, and let $ u \in \ix $. If $ \norm{\upara-u} = \mathcal{O}(\gamma^p) $ as $ \gamma \to 0 $ holds for some $ 0 < p < 1 $, then we have $ u \in \mathcal{R}(A^q) $ for each $ 0 < q < p $. \end{theorem} \begin{proof} We shall make use of some of the results in Komatsu~\cite[Sections 2 and 3]{Komatsu[66]}. The negative fractional power $ A^{-q} : \mathcal{X} \supset \mathcal{D} \to \mathcal{X} $ is defined in a direct way there, with a domain of definition $ \mathcal{D} $ that, under the assumptions on the asymptotical behaviour of $ \norm{\upara-u} $ made in our theorem, contains $ u $, cf.~\cite[estimate (3.7)]{Komatsu[66]}. We have $ A^{-q}u = \frac{\sin \pi q }{\pi} \int_0^\infty s^{-q} (A + sI)^{-1} u \ ds $ in fact (see \cite[equation (4.10)]{Komatsu[66]} for the details). From \cite[Proposition 4.13]{Komatsu[66]} the identity $ A^{q}(A^{-q} u) = A^{q-q} u = u $ then follows which means $ u \in \mathcal{R}(A^q) $, and this completes the proof of the theorem. \end{proof} \bigskip \noindent We note that the statement of Theorem \ref{th:converse_exact_data_2} cannot be extended to the case $ q = p $, in general. For a counterexample related with Tikhonov regularization, see Neubauer~\cite[p.~521]{Neubauer[97]}. In the case $ p = 1 $, the situation is different: \begin{theorem} \label{th:converse_exact_data_3} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $, and let $ u \in \ix $. In the case $ \norm{\upara-u} = \mathcal{O}(\gamma) $ as $ \gamma \to 0 $ we necessarily have $ u \in \mathcal{R}(A) $. \end{theorem} \proof From Theorem \ref{th:converse_exact_data_1} we obtain $ u \in \overline{\mathcal{R}(A)} $, and we next show that $ u \in \mathcal{R}(A) $ holds. By assumption we have $ \norm{(A + \gamma I)^{-1} u} = \mathcal{O}(1) $ as $ \gamma \to 0 $, and thus there exists an element $ v \in \mathcal{X} $ and a sequence $ (\gamma_n) $ of positive real numbers with $ \gamma_n \to 0 $ as $ n \to \infty $ such that we have weak convergence $ (A + \gamma_n I)^{-1} u \rightharpoonup v $ as $ n \to \infty $. From this, weak convergence $ A (A + \gamma_n I)^{-1} u \rightharpoonup Av $ as $ n \to \infty $ follows. On the other hand, due to $ u \in \overline{\mathcal{R}(A)} $ we have strong convergence $ A (A + \gamma_n I)^{-1} u \to u $ as $ n \to \infty $, and this shows $ A v = u $. \endproof \subsection{Saturation in case of exact data\xspace} We next present a saturation result for Lavrentiev regularization\xspace with exact data. \begin{theorem} \label{th:saturation_exact_data} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $, and let $ u \in \ix $. If $ \norm{\upara-u} = \mbox{\scriptsize $ \mathcal{O} $}(\gamma) $ as $ \gamma \to 0 $, then necessarily $ u = 0 $ holds. \end{theorem} \proof By assumption we have $ (A + \gamma I)^{-1}u \to 0 $ and thus $ (A + \gamma I)^{-1} Au \to 0 $ as $ \gamma \to 0 $. For the same term there also holds $ (A + \gamma I)^{-1} A u = u - \gamma (A + \gamma I)^{-1} u \to u $ as $ \gamma \to 0 $ according to Proposition \ref{th:direct_exact_data}, and this implies $ u = 0 $. Note that it follows from Theorem \ref{th:converse_exact_data_1} that $ u \in \overline{\mathcal{R}(A)} $ holds, thus Proposition \ref{th:direct_exact_data} indeed may be applied here. This completes the proof. \endproof \subsection{Some additional observations} Some conclusions of this section remain true under weaker hypotheses. Details are given in the following corollary. As a preparation we introduce the notation $ \mathbb{R}_+ = \{ \gamma \in \mathbb{R} \mid \gamma > 0 \} $. \begin{corollary} \label{th:liminf-weakening-exact-data} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $, and let $ u \in \ix $. \begin{myenumerate_indent} \item \label{liminf-weakening-exact-data_1} If there exists some sequence $ (\gamma_n) \subset \mathbb{R}_+ $ with $ \lim_{n \to \infty} \gamma_n = 0 $ such that $ \norm{\uparan-u} = \mathcal{O}(\gamma_n) $ as $ n \to \infty $ holds, then we necessarily have $ u \in \mathcal{R}(A) $. \item \label{liminf-weakening-exact-data_2} Suppose that for some sequence $ (\gamma_n) \subset \mathbb{R}_+ $ with $ \lim_{n \to \infty} \gamma_n =0 $ there holds \linebreak $ \norm{\uparan-u} = \mbox{\scriptsize $ \mathcal{O} $}(\gamma_n) $ as $ n \to \infty $. Then we necessarily have $ u = 0 $. \item \label{liminf-weakening-exact-data_3} If $ u \not \in \mathcal{R}(A) $ holds, then we have $ \norm{(A + \gamma I)^{-1}u } \to \infty $ as $ \gamma \to 0 $. \item \label{liminf-weakening-exact-data_4} If $ u \neq 0 $, then there exists a constant $ c > 0 $ such that $ \norm{\fpara} \ge c \gamma $ for $ \gamma > 0 $ small. \end{myenumerate_indent} \end{corollary} \begin{proof} Parts \ref{liminf-weakening-exact-data_1} and \ref{liminf-weakening-exact-data_2} follow similarly to the proofs of Theorems \ref{th:converse_exact_data_3} and \ref{th:saturation_exact_data}, respectively; one has to consider subsequences in those proofs then. Parts \ref{liminf-weakening-exact-data_3} and \ref{liminf-weakening-exact-data_4} are the logical negation of parts \ref{liminf-weakening-exact-data_1} and \ref{liminf-weakening-exact-data_2}, respectively. \end{proof} \bigskip \noindent We note that in the case ``$ \mathcal{X} $ Hilbert space, $ M = 1 $'' (the operator $ A $ is accretive\xspace then, cf.~the following section), the modified hypotheses in parts \ref{liminf-weakening-exact-data_1} and \ref{liminf-weakening-exact-data_2} of the preceding corollary coincide with the original hypotheses considered in Theorems~\ref{th:converse_exact_data_3} and \ref{th:saturation_exact_data}, respectively. This is an immediate result of the monotonicity of the functional $ \gamma \to \norm{\fpara}/\gamma $, cf.~Lemma~\ref{th:resolvent-monotone} in the appendix. We conclude this section with a remark on the closed range case. \begin{remark} \label{th:closed_range_exact_data} Note that throughout this section we do not require that the range $ \mathcal{R}(A) $ is non-closed. In case of a closed range, i.e., $ \mathcal{R}(A) = \overline{\mathcal{R}(A)} $, the results of Proposition \ref{th:direct_exact_data} and Theorems~\ref{th:converse_exact_data_1} and \ref{th:converse_exact_data_3} can be summarized as follows: \begin{align*} \lim_{\gamma \to 0} \upara = u \ \Longleftrightarrow \ u \in \mathcal{R}(A) \ \Longleftrightarrow \ \norm{\upara-u} = \mathcal{O}(\gamma) \textup{ as } \gamma \to 0. \end{align*} The case $ 0 < p < 1 $ considered in Theorem \ref{th:converse_exact_data_2} is not relevant in the closed range case, while the saturation case considered in Theorem \ref{th:saturation_exact_data} still is. \quad $ \vartriangle $ \end{remark} \section{Optimality concepts} \label{optimality} \subsection{Preliminaries} This section serves on the one hand as a preparation for the noisy data related converse and saturation results presented in the subsequent sections. The results of the present section, however, may be of independent interest: the impact on the optimality of parameter choices for Lavrentiev regularization\xspace is also established. Our main results in this section are obtained for operators on Hilbert spaces, but the preliminaries presented in this first subsection are considered for Banach spaces. So, throughout the present subsection we assume that $ A: \mathcal{X} \to \mathcal{X} $ is a nonnegative\xspace bounded linear operator on a Banach space $ \mathcal{X} $. Our main focus is on operators $ A $ with a non-closed range $ \mathcal{R}(A) $ or a nontrivial nullspace $ \mathcal{N}(A) $. The \textit{maximal best possible error} of \lavmet{}\xspace with respect to a given $ u \in \mathcal{X} $ and a noise level $ \delta > 0 $ is given by \begin{align} \Pdeltau & := \sup_{f^\delta: \, \norm{Au-f^\delta} \le \delta } \ \inf_{\gamma > 0 } \norm{\uparadel - u } \nonumber \\ & = \sup_{\Delta \in \mathcal{X}: \, \norm{\Delta} \le \delta } \ \inf_{\gamma > 0 } \norm{\fpara + (A+\gamma I)^{-1} \Delta }. \label{eq:pdelta} \end{align} The quantity $ \Pdeltau $ may serve as a tool for considering \quasioptimal{}ity\xspace of special parameter choices for Lavrentiev regularization\xspace, cf.~Definition \ref{th:optimality-def} below. First, however, we introduce other quantities that are often used in this direction: for $ u \in \mathcal{X}, \ \delta > 0 $ and $ 1 \le p \le \infty $ we define \begin{align} \Rdeltau{p} := \left\{ \begin{array}{lr} \displaystyle \inf_{\gamma > 0 } \big\{ \norm{\upara-u }^p + \big(M\tfrac{\delta}{\gamma}\big)^p \}^{1/p}, & \textup{if } p < \infty, \\ \displaystyle \inf_{\gamma > 0 } \max\big\{ \norm{\upara-u }, M \tfrac{\delta}{\gamma} \}, & \textup{if } p = \infty, \end{array} \right. \label{eq:rdelta} \end{align} where $ M $ is the constant from \refeq{postype}. Similar to some relations between $ p $-norms on $ \mathbb{R}^2 $, we obviously have $ \Rdeltau{\infty} \le \Rdeltau{p} \le \Rdeltau{1} \le 2\Rdeltau{\infty} $ for each $ u \in \mathcal{X} $ and $ \delta > 0 $, with $ 1 < p < \infty $. The most important quantity from this set of numbers is $ \Rdeltau{1} $. This is due to the fact that $ \norm{\uparadel - u } \le \norm{\upara-u } + M\frac{\delta}{\gamma} $ holds for each $ f^\delta \in \mathcal{X} $ with $ \norm{Au-f^\delta} \le \delta $. We next introduce two notations related with the optimality of parameter choices; see Raus/H\"amarik~\cite{Raus_Haemarik[07]} for similar notations. Other optimality concepts can be found in Vainikko~\cite{Vainikko[87.1]}. \begin{definition} \label{th:optimality-def} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a Banach space $ \ix $. We call a parameter choice $ 0 < \gamma = \gamma(\delta,f^\delta) \le \infty $ (with the notation $ \uparadel[\infty] := 0 $) for Lavrentiev regularization\xspace \begin{mylist_indent} \item \emph{orderoptimal\xspace}, if there exists a constant $ c > 0 $ such that for each $ u \in \mathcal{X}, \, \delta > 0 $ and $ f^\delta \in \mathcal{X} $ with $ \norm{Au-f^\delta} \le \delta $, we have $ \norm{\uparadel[\parac] - u } \le c \Pdeltau $, \item \emph{quasioptimal\xspace}, if there exists a constant $ c > 0 $ such that for each $ u \in \mathcal{X}, \, \delta > 0 $ and $ f^\delta \in \mathcal{X} $ with $ \norm{Au-f^\delta} \le \delta $, we have $ \norm{\uparadel[\parac] - u } \le c \Rdeltau{1} $. \end{mylist_indent} \end{definition} We obviously have $ \Pdelta(u) \le \Rdeltau{1} $ for each $ u \in \mathcal{X} $ and $ \delta > 0 $, so each orderoptimal\xspace parameter choice is quasioptimal\xspace. We now consider a modified discrepancy principle (sometimes called MD rule) which turns out to be a quasioptimal\xspace parameter choice strategy for Lavrentiev regularization\xspace in Banach spaces. \begin{example} \label{th:discrepancy_principle_mod} Fix real numbers $ b_1 \ge b_0 > M $, and let $ \Deltapara = A\uparadel - f^\delta $ for $ \gamma > 0 $. Consider the following parameter choice strategy: \begin{mylist} \item If $ \norm{f^\delta } \le b_1 \delta, $ then take $ \gamma = \infty $. \item Otherwise choose $ 0 < \gamma = \sigma < \infty $ such that $ b_0 \delta \le \norm{ \gamma (A+\gamma I)^{-1} \Deltapara } \le b_1 \delta $ holds. \end{mylist} It is shown in Plato/H\"amarik~\cite[Parameter Choice 4.1 and Theorem 4.4]{Plato_Haemarik[96]} that this parameter choice strategy is quasioptimal\xspace, if $ A : \mathcal{X} \to \mathcal{X} $ is a \postype bounded linear\xspace operator on a Banach space $ \mathcal{X} $. It is an open problem, in case of ill-posed problems, if the modified discrepancy principle is orderoptimal\xspace in such a general setting. For accretive\xspace operators on Hilbert spaces, however, strong \quasioptimal{}ity\xspace{}\xspace can be verified. Details are given in the following subsection. \quad $ \vartriangle $ \end{example} \subsection{\Quasioptimal{}ity\xspace in Hilbert spaces} For the following investigations we need to restrict the considered class of operators. \begin{definition} \label{th:accretive Let $ \mathcal{H} $ be a real or complex Hilbert space, with inner product $ \skp{\cdot}{\cdot} $. A bounded linear operator $ A: \mathcal{H} \to \mathcal{H} $ is called \emph{accretive\xspace}, if \begin{align} \textup{Re} \, \skp{Au}{u} \ge 0 \ \text{for } u \in \mathcal{H}. \label{eq:accretive} \end{align} \end{definition} We note that for real Hilbert spaces, condition \refeq{accretive} means $ \skp{Au}{u} \ge 0 $ for each $ u \in \mathcal{H} $, and the operator $ A $ is called \emph{monotone} then. \begin{example} The classical integration operator and the Abel integral operator (see Example~\ref{th:integration_abel_example}), considered on the space $ L^2(0,1) $ are accretive\xspace. For the integration operator this follows, e.g., from Halmos~\cite[Solution 150]{Halmos[78]}, and for the Abel integral operator see, e.g., \cite[Theorem 1.3.3]{Plato[95]}. \end{example} A bounded linear operator $ A: \mathcal{H} \to \mathcal{H} $ is accretive\xspace if and only it satisfies \refeq{postype} with $ M = 1 $. This follows, e.g., from Pazy~\cite[Theorem 1.4.2]{Pazy[83]}, in combination with Lemma~\ref{th:r-n-decomp} in the appendix, applied to the operator $ A + \gamma I $. Throughout this subsection we consider an accretive\xspace bounded linear operator $ A: \mathcal{H} \to \mathcal{H} $ on a Hilbert space $ \mathcal{H} $. Our main result of this subsection is Theorem \ref{th:pseuopt=orderopt} below, but first we introduce another quantity which is related to the concept of weak \quasioptimality{}\xspace (and variants of it sometimes are used as definition in fact, see, e.g., Hohage/Weidling~\cite{Hohage_Weidling[16]}): \begin{align} \Qdeltau & := \inf_{\gamma > 0 } \ \sup_{f^\delta: \, \norm{Au-f^\delta} \le \delta } \norm{\uparadel - u } \nonumber \\ & = \inf_{\gamma > 0 } \sup_{\Delta \in \mathcal{H}: \, \norm{\Delta} \le \delta } \norm{\upara-u + (A+\gamma I)^{-1} \Delta }, \quad u \in \mathcal{H}. \label{eq:qdelta} \end{align} The quantities $\Qdeltau $ and $\Pdeltau $ in \refeq{pdelta} differ in such a way that $ \inf $ and $ \sup $ are interchanged. The following proposition relates $\Qdeltau $ with $ \Rdeltau{p} $ from \refeq{rdelta}, i.e., weak \quasioptimality{}\xspace of a parameter choice for Lavrentiev regularization\xspace can by characterized by $\Qdeltau $. Note that in the current situation (accretive operators) we may consider those numbers $ \Rdeltau{p} $ with $ M = 1 $. \begin{proposition} Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $, with $ \mathcal{R}(A) \not = \ixh $. Then the following holds: \begin{myenumerate_indent} \item We have $ \Rdelta{2} \le \Qdelta \le \Rdelta{1} $ on $ \mathcal{H} $. \item A parameter choice strategy for Lavrentiev regularization\xspace is quasioptimal\xspace if and only if there exists a constant $ c > 0 $ such that for any $ u \in \mathcal{H}, \, \delta > 0 $ and $ f^\delta \in \mathcal{H} $ with $ \norm{Au-f^\delta} \le \delta $, we have $ \norm{\uparafdelb - u } \le c \Qdeltau $. \end{myenumerate_indent} \end{proposition} \begin{proof} The proof of part (a) is elementary, and details are left to the reader. We only note that under the given conditions on the operator $ A $ we have $ \norm{ (A+\gamma I)^{-1}} = \tfrac{1}{\gamma} $ for each $ \gamma > 0 $. The statement in part (b) is an immediate consequence of part (a). \end{proof} \bigskip \noindent We note that the assumption $ \mathcal{R}(A) \not = \mathcal{H} $ in the preceding proposition is essential. \subsection{Strong versus weak \quasioptimality{}\xspace} Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $, with $ \mathcal{R}(A) \not = \ixh $. For the proof of our main theorem of this section, we need to consider perturbations $ f^\delta = f + \delta \Delta_{\varepsilon} $ with $ \Delta_{\varepsilon} \in \mathcal{H}, \norm{\Delta_{\varepsilon}} = 1 $, such that the data error $ \delta \Delta_{\varepsilon} $ is nearly amplified by a factor $ 1/\gamma $ when the operator $ (A + \gamma I)^{-1} $ with $ \gamma > 0 $ is applied to it. The following lemma provides the basic ingredient. \begin{lemma} \label{th:resolvent_maximize Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $, with $ \mathcal{R}(A) \not = \ixh $. For parameters $ 0 < \varepsilon \le \gamma $ and $ v_{\varepsilon} \in \mathcal{H} $ with $ \norm{\Delta_{\varepsilon}} = 1 $ and $ \norm{A \Delta_{\varepsilon}} = \varepsilon $ we have \begin{align} \norm{(A + \gamma I)^{-1} \Delta_{\varepsilon}} \ge \big(1 - \frac{\varepsilon}{\gamma}\big) \frac{1}{\gamma} . \label{eq:resolvent_maximize} \end{align} \end{lemma} \proof We have \begin{align*} \norm{ \gamma (A + \gamma I)^{-1} \Delta_{\varepsilon} } = \norm{ \Delta_{\varepsilon} - A (A + \gamma I)^{-1} \Delta_{\varepsilon} } \ge 1 - \norm{ (A + \gamma I)^{-1} A \Delta_{\varepsilon} } \ge 1 - \frac{\varepsilon}{\gamma}, \end{align*} and this already completes the proof. \endproof \bigskip \noindent We note that the assumption $ \mathcal{R}(A) \neq \mathcal{H} $ in the preceding lemma is essential. This property is equivalent with $ \mathcal{R}(A) \not = \overline{\mathcal{R}(A)} $ or $ \mathcal{N}(A) \not = \{0\} $, cf.~Lemma~\ref{th:r-n-decomp} in the appendix. It is also equivalent with $ 0 \in \sigma(A) $, the spectrum of $ A $. This assumption guarantees, for arbitrarily small $ \varepsilon > 0 $, the existence of elements $ \Delta_{\varepsilon} \in \mathcal{H} $ with the properties stated in Lemma \ref{th:resolvent_maximize}. \bigskip \noindent In the proof of Theorem \ref{th:pseuopt=orderopt} considered below, we apply Lemma \ref{th:resolvent_maximize} with some specific $ \Delta_{\varepsilon} \in \mathcal{H} $ that in fact is obtained by an accretive\xspace transformation of the element $ u $. This guarantees that an inner product that occurs in the mentioned proof takes nonnegative values only. The following lemma provides the basic ingredient for the construction of those special elements $ v_{\varepsilon} $. \begin{lemma} \label{eq:phibeta-def} \ahassump. Let $ u \in \mathcal{H}, \, u \not \in \mathcal{R}(A) \cup \mathcal{N}(A) $, and let \begin{align} \varphi_{\beta} := \frac{(A + \beta I)^{-1} u}{\norm{(A + \beta I)^{-1} u}}, \quad \beta > 0. \end{align} For each real number $ 0 < \varepsilon < \frac{\norm{Au}}{\norm{u}} $ there exists a parameter $ \beta = \beta(\varepsilon) $ with $ \norm{A\varphi_{\betapara}} = \varepsilon $. \label{th:beta_alpha_choice} \end{lemma} \proof The function $ \beta \mapsto \norm{A\varphi_{\beta}} $ obviously is continuous on $ \mathbb{R}_+ $, and the lemma then follows from the asymptotic behaviours \begin{align} & \lim_{\beta \to 0} \norm{A\varphi_{\beta}} = 0, \qquad \lim_{\beta \to \infty} \norm{A\varphi_{\beta}} = \frac{\norm{Au}}{\norm{u}}. \label{eq:abetaalpha_behaviour} \end{align} In the sequel, the two statements in \refeq{abetaalpha_behaviour} will be verified. We consider first the case $ \beta \to 0 $. Obviously $ A (A + \beta I)^{-1} u = u - \beta (A + \beta I)^{-1} u $ is uniformly bounded with respect to $ \beta > 0 $, and in addition we have $ \norm{(A + \beta I)^{-1} u } \to \infty $ as $ \beta \to 0 $, cf. part \ref{liminf-weakening-exact-data_3} of Corollary \ref{th:liminf-weakening-exact-data}. This already completes the proof of the first statement in \refeq{abetaalpha_behaviour}. We next consider the case $ \beta \to \infty $. From a simple expansion and Lemma \ref{th:bias-monotone} in the appendix we obtain \begin{align*} \norm{A\varphi_{\beta}} = \frac{\norm{\beta (A + \beta I)^{-1} Au}}{\norm{\beta(A + \beta I)^{-1} u}} \to \frac{\norm{Au}}{\norm{u}} \quad \textup{as } \beta \to \infty, \end{align*} which is the second statement in \refeq{abetaalpha_behaviour}. This completes the proof of the lemma. \endproof \bigskip \noindent We next show that for accretive\xspace ill-posed operators on Hilbert spaces, the notions ``orderoptimal\xspace'' and ``quasioptimal\xspace'' are equivalent. This theorem provides the main result of this section. \begin{theorem} \label{th:pseuopt=orderopt} Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $, with $ \mathcal{R}(A) \not = \ixh $. Then we have $ \Rdeltau{2} \le \Pdeltau $ for each $ u \in \mathcal{H} $ and $ \delta > 0 $. \end{theorem} \begin{proof} 1)~In a first part we show that $ \Rdeltau{2} \le \Pdeltau $ holds for each $ u \not \in \mathcal{R}(A) $. \begin{myenumerate} \item We consider the trivial case $ u \in \mathcal{N}(A) $ first. Then we have $ \fpara = -u $ for each $ \gamma > 0 $, and in \refeq{pdelta} we may consider $ \Delta = - \delta \frac{ u }{\norm{u}} $ then. From this, $ \Pdeltau \ge \norm{u} = \Rdeltau{2} $ easily follows. \item Let us now consider the case $ u \not \in \mathcal{R}(A) \cup \mathcal{N}(A) $. We first show that \begin{align} \Pdeltau^2 \ge \ \inf_{\gamma > 0 } \Big\{ \normqua{\fpara} + \delta^2 \normqua{(A + \gamma I)^{-1} \Delta_{\varepsilon}} \Big\} \label{eq:pseuopt=orderopt-a} \end{align} holds, where $ \Delta_{\varepsilon} = -\varphi_{\betapara} $ for $ 0 < \varepsilon < \frac{\norm{Au}}{\norm{u}} $, and $ \varphi_{\betapara} $ is chosen as in Lemma \ref{th:beta_alpha_choice}. In fact, we have \begin{align*} \Delta_{\varepsilon} & = -c_\varepsilon (A+\beta(\varepsilon) I)^{-1}u, \quad \textup{with } c_\varepsilon = \frac{1}{\norm{(A+\beta(\varepsilon) I)^{-1}u}}, \end{align*} and then we obviously have \begin{align} \Pdeltau \ge \ \inf_{\gamma > 0 } \norm{\fpara + \delta (A+\gamma I)^{-1} \Delta_{\varepsilon} }. \label{eq:optimal-lemma-a} \end{align} We now expand, for $ \gamma > 0 $ fixed, the term on the right-hand side of \refeq{optimal-lemma-a}: \begin{align*} & \normqua{\fpara + \delta (A + \gamma I)^{-1} \Delta_{\varepsilon}} \\ & \qquad = \normqua{\fpara} + 2 \delta \textup{Re} \, \skp{€\fpara}{(A + \gamma I)^{-1} \Delta_{\varepsilon}} + \delta^2 \normqua{(A + \gamma I)^{-1} \Delta_{\varepsilon}}. \end{align*} For the inner product we have, by definition, \begin{align*} \skp{€\fpara}{(A + \gamma I)^{-1} \Delta_{\varepsilon}} = c_\varepsilon \gamma \skp{€(A + \gamma I)^{-1} u} {(A+\beta(\varepsilon) I)^{-1} (A + \gamma I)^{-1} u}, \end{align*} which has a nonnegative real part since the operator $ (A+\beta(\varepsilon) I)^{-1} $ is accretive\xspace. This implies \refeq{pseuopt=orderopt-a}. \item We next show that the inequality \refeq{pseuopt=orderopt-a} remains valid if the infimum on the right-hand side is considered for $ \gamma $ away from zero. For this purpose we choose some $ \gamma_0 $ with \begin{align*} 0 < \gamma_0 < \min \Big\{ \frac{\delta}{2\Pdeltau}, 2\frac{\norm{Au}}{\norm{u}} \Big\} \end{align*} and show in the sequel that \begin{align} \Pdeltau^2 \ge \ \inf_{\gamma \ge \gamma_0 } \Big\{ \normqua{\fpara} + \delta^2 \normqua{(A + \gamma I)^{-1} \Delta_{\varepsilon}} \Big\} \quad \textup{ for } \ 0 < \varepsilon \le \frac{\gamma_0}{2}, \label{eq:pseuopt=orderopt-b} \end{align} holds. In fact, we have \begin{align*} \norm{(A + \gamma I)^{-1} \Delta_{\varepsilon}} & \ge \norm{(A + \gamma_0 I)^{-1} \Delta_{\varepsilon}} \ge \big(1 - \frac{\varepsilon}{\gamma_0}\big) \frac{1}{\gamma_0} \ge \frac{1}{2\gamma_0} > \frac{\Pdeltau}{\delta} \quad \textup{ for } \ 0 < \gamma \le \gamma_0, \end{align*} by monotonicity of the norm of the resolvent operator, see Lemma \ref{th:resolvent-monotone} in the appendix, and Lemma \ref{th:resolvent_maximize} has also been applied. From this, \refeq{pseuopt=orderopt-b} follows easily. \item We proceed now with an estimation of the right-hand side of \refeq{pseuopt=orderopt-b}: For $ \gamma \ge \gamma_0 $ and $ \varepsilon \le \tfrac{\gamma_0}{2} $ we have, by Lemma \ref{th:resolvent_maximize}, \begin{align*} \norm{(A + \gamma I)^{-1} \Delta_{\varepsilon}} \ge \big(1 - \frac{\varepsilon}{\gamma}\big) \frac{1}{\gamma} \ge \big(1 - \frac{\varepsilon}{\gamma_0}\big) \frac{1}{\gamma}, \end{align*} and from \refeq{pseuopt=orderopt-b} we then obtain \begin{align*} \Pdeltau^2 & \ge \inf_{\gamma \ge \gamma_0 } \Big\{ \normqua{\fpara} + \big(1 - \frac{\varepsilon}{\gamma_0}\big)^2 \big(\frac{\delta}{\gamma}\big)^2 \Big\} \ge \big(1 - \frac{\varepsilon}{\gamma_0}\big)^2 \inf_{\gamma \ge \gamma_0 } \Big\{ \normqua{\fpara} + \big(\frac{\delta}{\gamma}\big)^2 \Big\} \\ & \ge \big(1 - \frac{\varepsilon}{\gamma_0}\big)^2 \inf_{\gamma > 0 } \Big\{ \normqua{\fpara} + \big(\frac{\delta}{\gamma}\big)^2 \Big\} = \big(1 - \frac{\varepsilon}{\gamma_0}\big)^2 \Rdeltau{2}^2 \quad \textup{ for } 0 < \varepsilon \le \frac{\gamma_0}{2}. \end{align*} Letting $ \varepsilon \to 0 $ now gives $ \Rdeltau{2} \le \Pdeltau $, and this completes the first part of the proof. \end{myenumerate} 2) In the second part of the proof, we show that the inequality $ \Rdelta{2} \le \Pdelta $ holds not only on $ \mathcal{H}\backslash\mathcal{R}(A) $ but all over the Hilbert space $ \mathcal{H} $. \begin{myenumerate} \item As a preparation we observe that, for $ \delta > 0 $ fixed, the functionals $ \Pdeltau $ and $ \Rdeltau{2} $ both are continuous with respect to $ u $. In fact, we have $ \norm{\fparau{u} - \fparau{\widetilde{u}}} \le \norm{u - \widetilde{u}} $ for each $ u, \widetilde{u} \in \mathcal{H} $, and from this the two inequalities \begin{align*} \vert \Pdeltau -\Pdeltautil \vert & \le \norm{u - \widetilde{u}}, \\ \vert \Rdeltau{2}^2 -\Rdeltautil{2}^2 \vert & \le 2 \max\{\norm{u},\norm{\widetilde{u}}\} \norm{u - \widetilde{u}}, \quad u, \, \widetilde{u} \in \mathcal{H}, \end{align*} are easily obtained. \item We are now in a position to verify that $ \Rdelta{2} \le \Pdelta $ holds over $ \mathcal{H} $. In fact, we already know that this estimate holds on $ \mathcal{H}\backslash\mathcal{R}(A) $ (see the first part of this proof), and in addition the functionals $ \Pdelta $ and $ \Rdelta{2} $ are continuous on $ \mathcal{H} $ for $ \delta > 0 $ fixed, see (a) of the second part of this proof. The assertion now follows from the fact that each nontrivial\xspace linear subspace of a normed space has an empty interior so that any $ u \in \mathcal{R}(A) $ is the limit of a sequence of elements not belonging to $ \mathcal{R}(A) $. \end{myenumerate} This completes the proof of the theorem. \end{proof} \bigskip \noindent For symmetric, positive semidefinite operators, a result similar to that of Theorem \ref{th:pseuopt=orderopt} can be found in Raus~\cite{Raus[84]}. As an immediate consequence of Theorem \ref{th:pseuopt=orderopt} we obtain the following result. \begin{corollary} \label{th:pseuopt=orderopt-cor}% Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $, with $ \mathcal{R}(A) \not = \ixh $. Then any parameter choice strategy for \lavmet{}\xspace is quasioptimal\xspace if and only if it is orderoptimal\xspace. \end{corollary} \begin{example} Under the conditions of Corollary \ref{th:pseuopt=orderopt-cor}, the parameter choice strategy considered in Example \ref{th:discrepancy_principle_mod} is quasioptimal\xspace (cf.~again Example \ref{th:discrepancy_principle_mod}) and therefore also orderoptimal\xspace. For symmetric, positive semidefinite operators, this is already observed in Raus~\cite{Raus[84]}. \quad $ \vartriangle $ \end{example} \section{Converse results in case of noisy data\xspace} \label{conv_noisdat} \subsection{Introductory remarks} The degree of smoothness of a solution $ u $, described here by the property $ u \in \mathcal{R}(A^p)$ for some $ p > 0 $, has impact on the decay rate of the best possible maximal error $ \Pdeltau $ as $ \delta \to 0 $. We cite the following well-known result; for a proof, see, e.g., \cite[Example 4.1]{Plato[96]}. As a preparation we note that our main interest are operators having a non-closed range $ \mathcal{R}(A) $, but this is nowhere explicitly required in this section. Further notes on the closed range case are given at the end of this section, cf.~Remark \ref{th:closed_range_noisy_data}. \begin{proposition} \label{th:direct_noisy_data} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a Banach space $ \ix $. \begin{myenumerate_indent} \item If $ u \in \overline{\mathcal{R}(A)} $ then $ \Pdeltau \to 0 $ as $ \delta \to 0 $. \item Let $ 0 < p \le 1 $. If $ u \in \mathcal{R}(A^p) $ then $ \Pdeltau = \mathcal{O}(\delta^{p/(p+1)}) $ as $ \delta \to 0 $. \end{myenumerate_indent} \end{proposition} \begin{proof} The modified discrepancy principle, cf.~Example \ref{th:discrepancy_principle_mod}, satisfies, see~\cite[Theorems 2.5 and 4.4]{Plato_Haemarik[96]}, $ \uparadel \to u $ as $ \delta \to 0 $ in the case $ u \in \overline{\mathcal{R}(A)} $. In addition, for $ 0 < p \le 1 $ we have $ \norm{\uparadel-u} = \mathcal{O}(\delta^{p/(p+1)}) $ as $ \delta \to 0 $ for each $ u \in \mathcal{R}(A^p) $. The statement of the proposition now easily follows. \end{proof} \bigskip \noindent We note that standard a priori parameter choices may be used as well in this proof. We may address the same topics as for exact data: \begin{mylist_indent} \item Are the given convergence results in Proposition \ref{th:direct_noisy_data} optimal, or, in other terms, are the conditions $ u \in \overline{\mathcal{R}(A)} $ and $ u \in \mathcal{R}(A^p) $ stated in parts (a) and (b) there also necessary, respectively? \item Is the considered range of values for $ p $, considered in part (b) of that proposition, maximal? \end{mylist_indent} We show in this section that the answers to those questions basically are affirmative, when accretive\xspace operators on Hilbert spaces are considered. \subsection{The converse results in case of noisy data\xspace} We start with a simple converse result which even holds in reflexive Banach spaces in fact. \begin{proposition} \label{th:converse_noisy_data_2} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $, and let $ u \in \ix $. If $ \Pdeltau \to 0 $ as $ \delta \to 0 $, then necessarily $ u \in \overline{\mathcal{R}(A)} $ holds. \end{proposition} \begin{proof} We consider the decomposition $ u = u_{R} + u_{N} $ with $ u_{R} \in \overline{\mathcal{R}(A)}, \, u_{N} \in \mathcal{N}(A) $, see Lemma \ref{th:r-n-decomp} in the appendix. From this and the consideration of $ f^\delta = A u $ in the definition of $ \Pdelta $ we obtain \begin{align} \Pdeltau \ge \inf_{\gamma>0} \norm{\fpara + 0} \ge \inf_{\gamma>0} \norm{-\hspace{-0.5mm}u_{N} + \fparau{u_{R}}} \ge \tfrac{1}{M} \norm{u_{N}} \label{eq:converse_noisy_data_2} \end{align} for each $ \delta > 0 $. The latter estimate in \refeq{converse_noisy_data_2} follows again by Lemma \ref{th:r-n-decomp} in the appendix and the fact that $ \eparaur \in \overline{\mathcal{R}(A)} $ holds for each $ \gamma > 0 $. Letting $ \delta \to 0 $ in \refeq{converse_noisy_data_2} shows $ u_{N} = 0 $ which completes the proof. \end{proof} \bigskip \noindent The following lemma serves as preparation for the converse and saturation results related with noisy data. \begin{lemma} \label{th:balance_parameter_choice Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $ with $ \mathcal{R}(A) \not = \ixh $, and let $ 0 \neq u \in \ixh $. Let the parameters $ \delta > 0 $ and $ \para > 0 $ be related by \begin{align} \label{eq:balance_parameter_choice \delta = \para^2 \norm{(A + \para I)^{-1} u }. \end{align} Then we have \begin{align} \label{eq:saturation_noisy_data_b \norm{\uparab-u} = \frac{\delta}{\para} \le \Pdeltau. \end{align} \end{lemma} \begin{proof} Due to Theorem \ref{th:pseuopt=orderopt} it is sufficient to show that $ \norm{\uparab-u} = \frac{\delta}{\para} \le \Rdeltau{\infty} $ holds. In fact, by monotonicity we have $ \norm{\uparab-u} \le \norm{\upara-u} $ for $ \gamma \ge \para $ (cf.~Lemma \ref{th:bias-monotone} in the appendix), and $ \frac{\delta}{\para} \le \frac{\delta}{\gamma} $ evidently holds for $ 0 < \gamma \le \para $. The identity in \refeq{saturation_noisy_data_b} is a direct consequence of the identity \refeq{balance_parameter_choice}. \end{proof} \begin{remark} \label{th:balance_parameter_choice_remark In the proofs of the following two theorems, Lemma \ref{th:balance_parameter_choice} is applied by choosing the noise level $ \delta $ as a function of the parameter $ \para $. This remark, however, considers the converse case where $ \para = \gamma_\delta $ is chosen as a function of $ \delta > 0 $, i.e., \begin{align*} \delta = \gamma_\delta^2 \norm{(A + \gamma_\delta I)^{-1} u }. \end{align*} It immediately follows from Lemma \ref{th:balance_parameter_choice} that this parameter choice strategy is orderoptimal\xspace. Note that this strategy is of theoretical interest only, and moreover note that the existence of $ \gamma_\delta $ follows from Corollary~\ref{th:bias-monotone-corollary} in the appendix. \quad $ \vartriangle $ \end{remark} We now present the main converse result related with noisy data. \begin{theorem} \label{th:converse_noisy_data_1 \ahuassump. If, for some $ 0 < p \le 1 $, we have $ \Pdeltau = \mathcal{O}(\delta^{p/(p+1)}) $ as $ \delta \to 0 $, then $ \norm{\upara-u} = \mathcal{O}(\gamma^p) $ as $ \gamma \to 0 $ holds. \end{theorem} \proof If $ \mathcal{R}(A) = \mathcal{H} $ holds, then the statement of the theorem follows immediately from Proposition~\ref{th:direct_exact_data}. We now assume that $ \mathcal{R}(A) \not = \mathcal{H} $, and without loss of generality we may also assume that $ u \not = 0 $ holds. Due to the usage of $ \gamma $ in \refeq{pdelta}, we change notation here and show $ \norm{\fpara[\parab]} = \mathcal{O}(\para^p) $ as $ \para \to 0 $. Let $ \para > 0 $ be arbitrary but fixed, and let $ \delta = \delta(\para) $ be given by \begin{align*} \delta = \para^2 \norm{(A + \para I)^{-1} u }, \end{align*} cf.~Lemma~\ref{th:balance_parameter_choice}. From that lemma we now obtain \begin{align*} \frac{\delta}{\para} \le \Pdeltau \le c \delta^{p/(p+1)} \end{align*} for some constant $ c $ which may be chosen independently from $ \delta $, and then $ \delta^{1/(p+1)} \le c \para $ and thus $ \delta^{p/(p+1)} \le c^p \para^p $ holds. This finally gives \begin{align*} \norm{\fpara[\parab]} = \frac{\delta}{\para} \le c \delta^{p/(p+1)} \le c^{p+1} \para^p, \end{align*} and this completes the proof of the theorem. \endproof \bigskip \noindent As an immediate consequence of Theorems \ref{th:converse_exact_data_2}, \ref{th:converse_exact_data_3} and \ref{th:converse_noisy_data_1} we obtain the following result. \begin{corollary} \label{th:converse_noisy_data_1b \ahuassump. \begin{myenumerate_indent} \item Let $ 0 < p < 1 $. If $ \Pdeltau = \mathcal{O}(\delta^{p/(p+1)}) $ as $ \delta \to 0 $, then $ u \in \mathcal{R}(A^q) $ for each $ 0 < q < p $. \item If $ \Pdeltau = \mathcal{O}(\delta^{1/2}) $ as $ \delta \to 0 $, then $ u \in \mathcal{R}(A) $. \end{myenumerate_indent} \end{corollary} \begin{remark} \label{th:closed_range_noisy_data} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a reflexive Banach space $ \ix $. Throughout this section we have not required that the range $ \mathcal{R}(A) $ is non-closed, in general. In case of a closed range, i.e., $ \mathcal{R}(A) = \overline{\mathcal{R}(A)} $, the results of Propositions \ref{th:direct_noisy_data} and \ref{th:converse_noisy_data_2} can be summarized as follows: \begin{align} \lim_{\delta \to 0} \Pdeltau = 0 \ \Longleftrightarrow \ u \in \mathcal{R}(A) \ \Longleftrightarrow \ \Pdeltau = \mathcal{O}(\delta^{1/2}) \textup{ as } \delta \to 0. \label{eq:closed_range_noisy_data_1} \end{align} The case $ 0 < p < 1 $ considered in Theorem \ref{th:converse_noisy_data_1} (in the Hilbert space setting in fact) is not relevant in the closed range case. If we have even $ \mathcal{R}(A) = \overline{\mathcal{R}(A)} $ and $ \mathcal{N}(A) = \{0\} $ (which in fact is equivalent to the identity $ \mathcal{R}(A) = \mathcal{X} $, cf.~Lemma~\ref{th:r-n-decomp} in the appendix), then $ \Pdeltau = \mathcal{O}(\delta) $ as $ \delta \to 0 $ holds for each $ u \in \mathcal{X} $. This follows from $ \max_{\gamma \ge 0} \norm{(A+\gamma I)^{-1}} < \infty $. \quad $ \vartriangle $ \end{remark} We have completed our considerations of converse results for Lavrentiev regularization\xspace in case of noisy data\xspace. Saturation will be considered in the next section. \section{Saturation in case of noisy data\xspace} \label{saturation_noisdat} We are now in a position to present a saturation result for Lavrentiev regularization\xspace in case of perturbed data. \begin{theorem} \label{th:saturation_noisy_data Let $ A: \ixh \to \ixh $ be an \accretive bounded linear operator on a Hilbert space $ \ixh $, with $ \mathcal{R}(A) \not = \ixh $, and let $ u \in \mathcal{H} $. If $ \Pdelta(u) = \mbox{\scriptsize $ \mathcal{O} $}(\delta^{1/2}) $ as $ \delta \to 0 $, then necessarily $ u = 0 $ holds. \end{theorem} \begin{proof} We prove the theorem by contradiction and assume that $ u \not = 0 $ holds. For any $ \para > 0 $ consider \begin{align} \delta = \delta(\para) := \para^2 \norm{(A + \para I)^{-1} u } > 0, \label{eq:balance_parameter_choice_b} \end{align} cf.~Lemma \ref{th:balance_parameter_choice}. From this lemma we then obtain \begin{align*} \frac{\delta}{\para} \le \Pdeltau = \mbox{\scriptsize $ \mathcal{O} $}(\delta)^{1/2} \quad \textup{as } \ \para \to 0, \end{align*} and thus $ \delta^{1/2} = \mbox{\scriptsize $ \mathcal{O} $}(\para) $ as $\para \to 0 $. Note that $ \delta = \delta(\para) > 0 $ for each $ \para > 0 $, and $ \delta \to 0 $ as $ \para \to 0 $. This finally gives \begin{align*} \norm{\uparab-u} = \frac{\delta}{\para} = \mbox{\scriptsize $ \mathcal{O} $}(\delta^{1/2}) = \mbox{\scriptsize $ \mathcal{O} $}(\para) \quad \textup{ as } \para \to 0. \end{align*} Theorem \ref{th:saturation_exact_data} now yields $ u = 0 $, a contradiction to the assumption made in the beginning of our proof. \end{proof} \begin{remark} We note that the assumptions in Theorem \ref{th:saturation_noisy_data} may be weakened without changing the conclusion of the theorem. We may in fact replace the condition $ \Pdelta(u) = \mbox{\scriptsize $ \mathcal{O} $}(\delta^{1/2}) $ as $ \delta \to 0 $ by $ \liminf_{\delta \to 0} \Pdelta(u)/\delta^{1/2} = 0 $ there. The only necessary modification in the proof of Theorem \ref{th:saturation_noisy_data} is that $ \para = \gamma_\delta $ in \refeq{balance_parameter_choice_b} is chosen as a function of $ \delta > 0 $ then, i.e., $ \delta = \gamma_\delta^2 \norm{(A + \gamma_\delta I)^{-1} u } $, and part \ref{liminf-weakening-exact-data_2} of Corollary \ref{th:liminf-weakening-exact-data} is also applied in this case. Further notes on $ \gamma_\delta $ are given in Remark~\ref{th:balance_parameter_choice_remark}. Note that we have $ \gamma_\delta \to 0 $ as $ \delta \to 0 $ which follows from Corollary~\ref{th:bias-monotone-corollary} in the appendix. The weakened version of Theorem \ref{th:saturation_noisy_data} implies that for given $ u \neq 0 $ and $ \delta_0 > 0 $, there exists a constant $ c > 0 $ such that \begin{align*} c \delta^{1/2} \le \Pdeltau \quad \textup{for } \ 0 < \delta \le \delta_0. \end{align*} We note that there exists a result for Tikhonov regularization which is similar to Theorem~\ref{th:saturation_noisy_data}. For Tikhonov regularization, however, a weakening like the one considered in the present remark is not possible. For a counterexample see Neubauer~\cite{Neubauer[97]}. \quad $ \vartriangle $ \end{remark} \begin{remark} \label{th:closed_range_noisy_data_2} Note that the assumption $ \mathcal{R}(A) \not = \mathcal{H} $ made in Theorem \ref{th:closed_range_noisy_data} includes the case $ \overline{\mathcal{R}(A)} = \mathcal{R}(A), \, \mathcal{N}(A) \not = \{0\} $ (closed range, nontrivial nullspace). Note moreover that the saturation level is different if $ \mathcal{R}(A) = \mathcal{H} $ holds. In this case we have (even for nonnegative\xspace operators on Banach spaces) $ \Pdeltau = \mathcal{O}(\delta) $ as $ \delta \to 0 $ for each $ u \in \mathcal{H} $, cf.~Remark \ref{th:closed_range_noisy_data}. \quad $ \vartriangle $ \end{remark} \section{Auxiliary results for nonnegative\xspace operators} \label{auxiliary} In this section we present some auxiliary results which are being used at several places in this paper. We start with a structural result on the range and nullspace of a nonnegative\xspace operator on a reflexive Banach space. \begin{lemma} \label{th:r-n-decomp For a \postype bounded linear\xspace operator $ A: \mathcal{X} \to \mathcal{X} $ on a reflexive Banach space $ \mathcal{X} $ we have $ \overline{\mathcal{R}(A)} \oplus \mathcal{N}(A) = \mathcal{X} $, where the symbol $ \oplus $ denotes direct sum. In addition, there holds $ \norm{ u_{N} } \le M \norm{ u_{R}+u_{N} } $ for each $ u_{R} \in \overline{\mathcal{R}(A)} $ and each $ u_{N} \in \mathcal{N}(A) $, where the constant $ M $ is taken from \refeq{postype}. \end{lemma} \begin{proof} See, e.g., \cite[Theorem 1.1.10]{Plato[95]}. \end{proof} \bigskip \noindent We now present results on the behaviour of the bias and the resolvent. \begin{lemma} \label{th:bias-monotone Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a Banach space $ \ix $, and let $ u \in \ix $. \begin{myenumerate_indent} \item The functional $ \gamma \mapsto \norm{\fpara} = \norm{\gamma (A + \gamma I)^{-1} u } $ is continuous on $ \mathbb{R}_+ $. \item We have $ \lim_{\gamma\to \infty} \norm{\fpara} = \norm{ u } $. \item If \refeq{postype} holds with $ M = 1 $, then $ \gamma \mapsto \norm{\fpara} $ is monotonically increasing on $ \mathbb{R}_+ $. \end{myenumerate_indent} \end{lemma} \begin{proof} Continuity of the mapping $ \gamma \mapsto \norm{\fpara} $ is obvious. The asymptotical behaviour of the bias considered in part (b) follows from the representation \begin{align} \norm{\fpara} = \norm{(\gamma^{-1} A + I)^{-1} u } = \norm{(I + \sigma A)^{-1} u} =: g(\sigma) \quad \textup{with } \sigma := \gamma^{-1} \label{eq:bias-monotone \end{align} and by letting $ \sigma \to 0 $ then. Next we consider monotonicity. For $ 0 \le \sigma_1 \le \sigma_2 $ we have $ (I + \sigma_2 A)^{-1} (I + \sigma_1 A) = \omega I + (1-\omega)(I + \sigma_2 A)^{-1} $ with $ 0 \le \omega := \tfrac{\sigma_1}{\sigma_2} \le 1 $. Therefore $ \norm{ (I + \sigma_2 A)^{-1} (I + \sigma_1 A) } \le 1 $ holds, and then $ \norm{ (I + \sigma_2 A)^{-1} u } \le \norm{ (I + \sigma_1 A)^{-1} u } $ easily follows. This means that the functional $ g $ in \refeq{bias-monotone} is monotonically decreasing on $ \mathbb{R}_+ $, and therefore the function $ \norm{\fpara} $ is monotonically increasing with respect to $ \gamma $. \end{proof} \bigskip \noindent As an immediate consequence of Lemma \ref{th:bias-monotone} we obtain the following result. \begin{corollary} Let $ A: \ix \to \ix $ be a \postype bounded linear operator on a Banach space $ \ix $. Then for each $ 0 \not = u \in \mathcal{X} $, the function \begin{align*} f(\gamma) = \gamma^2 \norm{(A + \gamma I)^{-1} u }, \quad \gamma > 0, \end{align*} is continuous on $ \mathbb{R}_+ $, and in addition $ \lim_{\gamma\to 0} f(\gamma) = 0 $ and $ \lim_{\gamma\to \infty} f(\gamma) = \infty $ holds. If \refeq{postype} holds with $ M = 1 $, then the function $ f $ is strictly increasing on $ \mathbb{R}_+ $. \label{th:bias-monotone-corollary} \end{corollary} In a Hilbert space setting we finally present a monotonicity result for the resolvent. \begin{lemma} \label{th:resolvent-monotone \ahassump. For $ u \in \mathcal{H} $ fixed, the functional $ r(\gamma) = \norm{ (A + \gamma I)^{-1} u }, \gamma > 0 $, is monotonically decreasing on $ \mathbb{R}_+ $. \end{lemma} \begin{proof} \begin{myenumerate} \item In a first step we assume that the operator $ A $ has a continuous inverse $ A^{-1}: \mathcal{H} \to \mathcal{H} $. The functional $ r $ then can be written in the form $ r(\gamma) = \norm{(I + \gamma A^{-1})^{-1} A^{-1}u } $ which according to the proof of Lemma \ref{th:bias-monotone} is monotonically decreasing, since the operator $ A^{-1} $ is accretive\xspace. \item We now proceed with the general case for $ A $ and consider the operator $ A_\varepsilon = A + \varepsilon I : \mathcal{H} \to \mathcal{H} $ which obviously is an accretive\xspace invertible operator. The first part of this proof shows that $ \gamma \mapsto \norm{ (A + (\gamma + \varepsilon) I)^{-1} u } $ is decreasing on $ \mathbb{R}_+ $ which means that $ \gamma \mapsto \norm{ (A + \gamma I)^{-1} u } $ is decreasing on the interval $ (\varepsilon, \infty) $. Letting $ \varepsilon \to 0 $ then yields the desired monotonicity result. \end{myenumerate} This completes the proof of the lemma. \end{proof} \bigskip \noindent \textbf{Acknowledgment.} The author would like to thank Bernd Hofmann (TU Chemnitz). Without his continued encouragement and advice this work would not have been possible.
be742c748bda70c92004fa6aeb295bf9980a9747
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Due to their ease of fabrication and integration as well as their large area and high quality factor of resonances \citep{Lee2012}, photonic crystal slabs with one or two dimensional periodicity \cite{Johnson1999,Fan2002,Zhou2014} have been widely used in many applications, such as filters \cite{Suh2004a}, lasers \cite{Hirose2014}, and sensors \cite{Chow2004}. For more efficient utilization of light, it is often desirable to achieve highly directional out-of-plane coupling of light from photonic crystal slabs, in which light predominantly radiates to only one side of the slab. This would eliminate the need of a back-reflection mirror in high-power photonic crystal surface emitting lasers (PCSELs) \cite{Hirose2014}, where fabrication uncertainties in the laser wavelength and mirror-cavity distance currently make reliably achieving high slope efficiency difficult. This could also lead to increased efficiency of grating couplers for silicon photonics and light detection and ranging (LIDAR) devices. Previous designs of grating couplers have achieved a top-down asymmetry ratio (defined as the ratio of power going to the top and to the bottom) of up to 50:1 \cite{Taillaert2004, Roncone1993, Vermeulen2010, Subbaraman2015}, but they typically make use of a substrate reflector or involve multiple layers and grooves \cite{Fan2007}, which complicate fabrication and could be difficult to scale to larger areas if desired. Asymmetric out-of-plane emission from photonic crystal defect cavities of 4:1 has also been demonstrated \cite{Ota2015}, but all these works were guided primarily by numerical optimization. It is thus important to gain an understanding of the fundamental bounds on asymmetric radiation, and use such bounds as a guide to design stronger asymmetries. Closely related to highly directional radiation is achieving perfect absorption of fields incident from a single side of a weakly-absorbing photonic structure. This can be viewed as the time-reversal partner of the single-sided radiation emission process. An increased absorption efficiency could be important for improving the performance of many devices, including modulators \cite{Liu2011}, photodetectors \cite{Xia2009}, solar cells \cite{Yu2010,Pospischil2014}. However, the single-pass absorption of a thin absorbing layer in air is at most 50$\%$ \cite{Hadley1947,Radi2013}. By combining electric and magnetic responses \textcolor{black}{or utilizing material anisotropy}, it is possible to design metamaterial perfect absorbers with near unity absorptance \cite{Radi2013,Landy2008,Watts2012,Baranov2015}, but such designs can be difficult to implement at optical frequencies. Recent work achieving perfect absorption in photonic crystal structures has either employed illumination from both sides and used the interference between the beams---analogous to a time-reversed laser---to achieve coherent perfect absorption \cite{Chong2010,Wan2011,Sun2014}, or employed a back-reflection mirror and critical coupling to resonances to approach perfect absorption \cite{Tischler2006,Lin2011,Piper2014,Liu2014,Zhu2015,Piper2016,Sturmberg2016}. Alternatively, specific surface textures can be designed to enhance light trapping and subsequent absorption\cite{Yu2010,Ganapati2014,Oskooi2014}, but a back-reflection mirror is still required to keep the photons inside the absorbing layer. In general, however, two-sided illumination can often be challenging to implement in realistic systems, while backing mirrors are often either lossy (e.g. metallic mirrors) or require additional fabrication efforts (e.g. distributed Bragg reflectors). Therefore, the possibility of achieving perfect absorption of fields incident from a single side, without the aid of backing mirrors, is highly attractive and could open up many engineering possibilities. \textcolor{black}{One recent approach achieving this is to utilize accidental degeneracies of critically coupled modes with opposite symmetries \cite{Piper2014a}, but such an approach requires aligning the frequencies and quality factors of multiple resonances.} Here, we first design structures with highly directional radiation, and then consider the time-reversal scenario at critical coupling to realize devices with high absorption efficiency. Moreover, away from the strongly-absorbing resonance frequency, light can be mostly transmitted through the designed devices, which could have important applications in multi-junction solar cells. Previous work \cite{Wang2013} has used a temporal coupled-mode theory (TCMT) formalism \cite{Suh2004,Fan2003} with a single resonance and two coupling ports (one on each side of the slab) to examine bounds of asymmetric radiation from photonic crystal slabs. There, they reached the conclusion that the asymmetry ratio is bounded by $(1+r)/(1-r)$, where $r$ is the background amplitude reflection coefficient. For index contrasts found in realistic materials and at optical frequencies, this bound limits the proportion of radiation going to one side of the photonic crystal slab to around $90\%$, even for the high index contrast between silicon and air. For a smaller index contrast, this will be even more significantly different from perfect directional radiation. We find, however, that in more general scenarios, the bounds in Ref. \cite{Wang2013} can be greatly surpassed. In a periodic photonic structure, the natural choice of mode basis is the momentum-conserving Bloch-wave basis \cite{Joannopoulos2011}. As shown in Fig.~\ref{fig:tcmt}(a), for general incident directions (nonzero in-plane momentum $\vec{k}_\parallel$) in asymmetric structures, the time-reversal operation relates the resonance at $\vec{k}_\parallel$ to the resonance at $-\vec{k}_\parallel$, suggesting that a two-resonance, four-port model is required to impose time-reversal constraints in the more general case. \textcolor{black}{Moreover, reciprocity automatically ensures that the two resonances share identical frequencies, eliminating the need for exquisite degenerate-frequency alignment to achieve multi-resonant responses.} Only when the system under consideration possesses certain symmetries---either in the structure \cite{Hsu2013a} ($C_2^z$, i.e. $180^\circ$ rotation around the out-of-plane axis) or in the incident field (normal incidence)---can we use the simplified model \cite{Wang2013} with only a single independent port on each side of the slab. In this article we show that the general two-resonance, four-port TCMT formalism, widely applicable to periodic structures with arbitrary geometry or in-plane momentum, enables bounds with significantly higher (sometimes even infinitely high) radiation asymmetry for realistic materials when $C_2^z$ symmetry of the structure is broken. As an example, we apply this formalism to inversion-symmetric ($P$-symmetric) structures without $C_2^z$ symmetry. Through numerical examples, we show that a top-down asymmetry ratio exceeding $10^4$ can be achieved by tuning the resonance frequency to coincide with the perfectly transmitting frequency on the Fabry-Perot background. The emission direction can also be rapidly switched from top to bottom by tuning the $\vec{k}_\parallel$ vector or frequency. These results provide important design principles for PCSELs, grating couplers, LIDARs, and many other applications that could benefit from directional emission and rapid tuning. In addition, we derive analytical expressions for the transmission spectrum and discuss features such as full transmission or reflection. We then show that such highly asymmetric coupling to the two sides of the photonic crystal slab can also be employed to achieve perfect absorption of light incident from one side of the slab, without the need of back-reflection mirrors as in previous designs. \section{Temporal Coupled-Mode Theory Formalism} We start by considering arbitrary photonic crystal slab structures embedded in a uniform medium (identical substrate and superstrate). We assume weak coupling, linearity, energy conservation and time-reversal symmetry in the system, and we consider frequencies below the diffraction limit so that higher-order diffractions are not present. A plane wave with in-plane momentum $\vec{k}_\parallel=(k_x,k_y)$ incident from the top [port 1, see Fig.~\ref{fig:tcmt}(a)] will only couple to resonances and outgoing waves with the same $\vec{k}_\parallel$ (conservation of Bloch momentum). We shall consider the typical case where there is a single resonance at $\vec{k}_\parallel$ near the frequencies of interest, with the transmission spectrum consisting of a Fabry-Perot background and sharp resonant features, as in Fig.~\ref{fig:tcmt}(b). To describe time-reversal symmetry constraints for general geometries and incident angles, we need to include the resonance at $-\vec{k}_\parallel$ in our description as well, resulting in a two-resonance, four-port model. Although conservation of Bloch momentum means that each input port only excites either the $\vec{k}_\parallel$ or $-\vec{k}_\parallel$ resonance, the two resonances still influence one another indirectly via the time-reversal symmetry constraints on the coupling matrices between resonances and ports ($K$ and $D$ in Eq. (\ref{eq:TCMT1})), as described below. Writing down expressions consistent with momentum conservation and time-reversal symmetry, we obtain the TCMT equations \begin{align} \frac{d\textbf{A}}{dt}&=\left(j\omega-\frac{1}{\tau}-\frac{1}{\tau_{nr}}\right)\textbf{A}+K^{\rm T}|s_+\rangle,\; |s_-\rangle=C|s_+\rangle+D\textbf{A},\label{eq:TCMT1}\\ C&=e^{j\phi}\begin{pmatrix} 0 & r & 0 & jt\\r & 0 & jt & 0\\0 & jt & 0 & r\\jt & 0 & r & 0\\ \end{pmatrix},\; D=K\sigma_x=\begin{pmatrix} 0 & d_1\\d_2 & 0\\0 & d_3\\d_4 & 0\\ \end{pmatrix},\label{eq:TCMTcoeff} \end{align} \begin{comment} \begin{align} \frac{d}{dt}\begin{pmatrix} a_1\\a_2 \end{pmatrix}&= (j\Omega-\Gamma) \begin{pmatrix} a_1\\a_2 \end{pmatrix} +\begin{pmatrix} \kappa_1 & 0 & \kappa_3 & 0\\ 0 & \kappa_2 & 0 & \kappa_4\\ \end{pmatrix} \begin{pmatrix} s_{1+}\\s_{2+}\\s_{3+}\\s_{4+} \end{pmatrix}\\ \begin{pmatrix} s_{1-}\\s_{2-}\\s_{3-}\\s_{4-} \end{pmatrix} &=e^{j\phi}\begin{pmatrix} 0 & r & 0 & jt\\r & 0 & jt & 0\\0 & jt & 0 & r\\jt & 0 & r & 0\\ \end{pmatrix},\begin{pmatrix} s_{1+}\\s_{2+}\\s_{3+}\\s_{4+} \end{pmatrix} +\begin{pmatrix} 0 & d_1\\d_2 & 0\\0 & d_3\\d_4 & 0\\ \end{pmatrix} \begin{pmatrix} a_1\\a_2 \end{pmatrix},\label{eq:TCMT3}\\ C&=e^{j\phi}\begin{pmatrix} 0 & r & 0 & jt\\r & 0 & jt & 0\\0 & jt & 0 & r\\jt & 0 & r & 0\\ \end{pmatrix}, D=K^T\sigma_x=\begin{pmatrix} 0 & d_1\\d_2 & 0\\0 & d_3\\d_4 & 0\\ \end{pmatrix} \end{align} \end{comment} where $\textbf{A}=(A_1,A_2)^{\rm T}$ are the amplitudes of the two resonances (at $\vec{k}_\parallel$ and $-\vec{k}_\parallel$ respectively), $\omega$ is the resonance frequency shared by both resonances, $\tau$ is the radiative $e^{-1}$-decay lifetime ($\omega$, $\tau$ are identical for the two resonances due to reciprocity), $\tau_{nr}$ is the $e^{-1}$-decay lifetime for nonradiative processes such as absorption, and $|s_{+}\rangle=(s_{1+},s_{2+},s_{3+},s_{4+})^{\rm T}$, $|s_{-}\rangle=(s_{1-},s_{2-},s_{3-},s_{4-})^{\rm T}$ are the amplitudes of the incoming and outgoing waves. $C$ is the scattering matrix for the direct (non-resonant) transmission and reflection through the slab (namely, the Fabry-Perot background). Energy conservation and reciprocity constrain $C$ to be unitary and symmetric. For identical substrates and superstrates, $C$ takes the form in Eq.~(\ref{eq:TCMTcoeff}), where $t$ and $r$ are real numbers satisfying $r^2+t^2=1$ that characterize the Fabry-Perot background, and the phase $\phi$ depends on the choice of reference plane position. $K$ and $D$ are the coupling matrices in and out of the resonances. Time reversal flips the two resonances, so instead of the usual relation $D=K$ \cite{Suh2004}, here we have $D=K\sigma_x$, where $\sigma_x$ is the $2\times 2$ $X$-Pauli matrix acting to flip the resonances. We note that an alternative (and equivalent) formalism is to adopt a basis in which the underlying modes are time-reversal invariant (for which the standard multimode treatment \cite{Suh2004} is adequate), by superimposing the resonance at $\vec{k}_\parallel$ with its time-reversal partner at $-\vec{k}_\parallel$. This is to be contrasted with a basis change for ports, as discussed in Ref. \cite{Ruan2012}. Detailed derivations of this and the following expressions are given in \href{link}{Supplement 1}. \begin{figure}[h!tb] \centering \fbox{\includegraphics[width=\linewidth]{setup_H2.pdf}} \caption{Temporal coupled-mode theory (TCMT) setup and transmission spectrum. (a) Schematic of our TCMT setup with four ports and two resonances related by the time reversal operation. This general setup is valid for structures with arbitrary shapes and incident angles as long as the assumption of four ports and two resonances is correct. (b) Typical transmission spectrum of an inversion-symmetric, $C_2^z$-symmetry-broken structure, with the Fano resonances exhibiting full transmission at certain frequencies as predicted by our TCMT formalism. Strong asymmetry is achieved when the Fano resonance is aligned with the frequencies where the background reaches full transmission (red circles).} \label{fig:tcmt} \end{figure} Energy conservation and time-reversal symmetry impose constraints on the coefficients. Energy conservation requires \begin{align}\label{eq:econs} |d_2|^2+|d_4|^2=|d_1|^2+|d_3|^2=\frac{2}{\tau}, \end{align} while time-reversal symmetry gives the constraint $D=K\sigma_x$ and the two independent equations \begin{align}\label{eq:fulld1} e^{j\phi}(rd_2^*+jtd_4^*)+d_1&=0,\\ e^{j\phi}(jtd_2^*+rd_4^*)+d_3&=0.\label{eq:fulld2} \end{align} In the following, we shall fix the phase $\phi$ to be 0 by appropriately choosing the location of our reference plane. Eqs.~(\ref{eq:econs}, \ref{eq:fulld1}, \ref{eq:fulld2}) impose constraints on the values and phases of the couplings, and hence constrain the transmission spectrum and set bounds on the asymmetric coupling ratios. From the preceding equations, we can derive an expression for the transmission spectrum \cite{Wang2013,Fan2003} that only depends on the frequencies and decay rates of the resonances and the transmission and reflection coefficients of the direct Fabry-Perot pathway. The full scattering matrix including the direct pathway and resonance pathway \cite{Suh2004} is given by Eq.~(\ref{eq:scatmat}) in \href{Link}{Supplement 1}. The power reflection and transmission coefficient for a wave incident from port 1 correspond to the amplitude squared of the (1,2), (1,4) element of the scattering matrix, given by \begin{align}\label{eq:R} R&=|S_{12}|^2=\left|e^{j\phi}r+\frac{d_1d_2}{j(\omega-\omega_0)+\frac{1}{\tau}+\frac{1}{\tau_{nr}}}\right|^2,\\\label{eq:T} T&=|S_{14}|^2=\left|e^{j\phi}jt+\frac{d_1d_4}{j(\omega-\omega_0)+\frac{1}{\tau}+\frac{1}{\tau_{nr}}}\right|^2, \end{align} and the power reflection coefficient can be rewritten in the lossless limit $\tau_{nr}\rightarrow\infty$ as \begin{align}\label{eq:nolossR} R&=\frac{\left[r(\omega-\omega_0)\pm\sqrt{\frac{4}{\tau_1\tau_2}-\frac{r^2}{\tau^2}-\frac{2}{\tau\sigma}-\frac{1}{\sigma^2 r^2}}\right]^2+\left(\frac{1}{\sigma r}\right)^2}{(\omega-\omega_0)^2+\frac{1}{\tau^2}}, \end{align} where we have written $\tau_i=2/|d_i|^2$, $1/\sigma=1/\tau_1-1/\tau_4$ to simplify the expression. This expression provides general conditions for reaching full transmission or reflection with the Fano resonance. As shown in \href{link}{Supplement 1}, full transmission $R=0$ can only occur when the coupling rates satisfy $\tau_1=\tau_4$, $\tau_2=\tau_3$ ($P$-symmetric coupling), consistent with the transmission spectrum shown in Fig.~\ref{fig:tcmt}(b). Full reflection $R=1$ can only occur when the coupling rates satisfy $\tau_1=\tau_2$, $\tau_3=\tau_4$ ($C_2^z$ symmetric coupling), consistent with the results in Ref.~\cite{Wang2013}. Note that for structures that do not have $P$ or $C_2^z$ symmetry, it is still possible for the coupling rates for resonances to be $P$ or $C_2^z$ symmetric, leading to full transmission/reflection features in the frequency spectrum (see for example \href{link}{Supplement 1}, Fig. \ref{fig:multispec}). \section{General Bounds on Asymmetric Coupling Rates} We now derive bounds on the achievable asymmetry of coupling to the top and bottom based on Eqs.~(\ref{eq:fulld1}, \ref{eq:fulld2}) derived from time-reversal symmetry. Denote $|d_4/d_2|=a_r$, $|d_3/d_1|=a_\ell$, and define the asymmetric coupling ratios on the right ($\vec{k}_\parallel$) and left ($-\vec{k}_\parallel$) directions of the resonator as $a_r^2$ and $a_\ell^2$ (the ratio of the power going to bottom and top). By taking the ratio of Eqs.~(\ref{eq:fulld1}, \ref{eq:fulld2}), we find \begin{align} a_\ell^2=\left|\frac{jt+ra_re^{j\theta}}{r+jta_re^{j\theta}}\right|^2=\frac{t^2+r^2a_r^2+2tra_r\sin\theta}{r^2+t^2a_r^2-2tra_r\sin\theta}, \end{align} where $\theta=\arg(d_2)-\arg(d_4)$ characterizes the phase difference between $d_2$ and $d_4$. This gives the bound \begin{align}\label{eq:fullbound} \left|\frac{t-ra_r}{r+ta_r}\right|\leq a_\ell\leq \left|\frac{t+ra_r}{r-ta_r}\right|. \end{align} \begin{figure}[htb] \centering \fbox{\includegraphics[width=\linewidth]{bound_H3.pdf}} \caption{Simulated structures and verification of TCMT bounds. (a) The $P$-symmetric structure we use in our numerical examples and its structural parameters. $a$: periodicity of photonic crystal, $h$: height of central slab, $w$: width of central slab, $n_0$: refractive index of central slab, $d$: height of additional pieces on the sides (the width of the additional pieces is $(a-w)/2$), $n_d$: refractive index of additional pieces on the sides; (b) Numerical verification of TCMT bounds on asymmetric radiation for $P$-symmetric structures. Red lines indicate the bound from Eq.~(\ref{eq:tbound}). Each blue cross indicates simulation results of the asymmetry for a given structure, optimized over in-plane momentum. The transmissivity $t$ is fitted from the Fabry-Perot background, and the asymmetric coupling ratio is calculated from the Poynting flux in the top and bottom directions. \label{fig:bound}} \end{figure} Therefore, the amount of achievable asymmetry to top and bottom on the left is bounded by that on the right, and vice versa. Note that in general the phase here can be tuned through a $2\pi$ cycle, so the bounds---even up to infinitely high asymmetry ratio---should be saturable for appropriate parameter choices. \textcolor{black}{The coefficients that enter the bounds are the transmission/reflection coefficients ($t,r$) of the direct process (Fabry-Perot background), as opposed to the total transmission/reflection including the resonant pathway.} If the structure has $C_2^z$ symmetry, the two channels on the top and bottom will be constrained to have the same coupling rates, so $d_1=d_2$, $d_3=d_4$, $a_\ell=a_r$. Plugging this into Eq.~(\ref{eq:fullbound}), we find the same bound as in Ref.~\cite{Wang2013}: $\frac{1-r}{1+r}\leq a_\ell^2=a_r^2\leq\frac{1+r}{1-r},$ which shows the consistency of our approach. In typical photonic crystal systems at optical frequencies, the index contrast between the slab and the background medium is limited to around 3, which constrains the interface reflection coefficient to be less than 0.5 for most incident angles. This results in the Fabry-Perot direct pathway reflection coefficient $r=\sqrt{1-t^2}$ being considerably smaller than 1, so strong asymmetry in the decay rates is difficult to achieve for $C_2^z$-symmetric structures without the use of an additional back-reflecting mirror. The general bound Eq.~(\ref{eq:fullbound}) suggests, however, that much stronger asymmetry can be achieved if we break the $C_2^z$ symmetry of the system. A simple example is when the structure possesses inversion symmetry $P$ but breaks $C_2^z$ symmetry, as shown in Fig.~\ref{fig:bound}(a). In this case, the decay rates must satisfy $d_1=d_4$, $d_2=d_3$, $a_\ell=1/a_r$, and Eq.~(\ref{eq:fullbound}) becomes \begin{align}\label{eq:tbound} \frac{1-t}{1+t}\leq a_\ell^2=\frac{1}{a_r^2}\leq \frac{1+t}{1-t}. \end{align} For any index contrast, due to the up-down symmetry of the background material, the Fabry-Perot background will always have frequencies with full transmission, as exemplified by the red circles in Fig.~\ref{fig:tcmt}(b). Therefore, by tuning the resonance frequency to such points, the lower and upper bounds of Eq.~(\ref{eq:tbound}) approach $0$ and $+\infty$. Moreover, the bound can be saturated for appropriate choices of structural parameters and wavevectors, yielding arbitrarily high asymmetric decay rates of the photonic structure to top and bottom directions. \textcolor{black}{We note that similar design principles of breaking $C_2^z$ symmetry to achieve higher radiation asymmetry have also been realized in Ref. \cite{Wade2015} using the different design intuition of destructive interference.} To verify these analytical results, we perform numerical simulations using the finite difference time domain (FDTD) method \cite{Taflove2013} with a freely available software package \cite{Oskooi2010}. We extract the coupling rate to top or bottom by monitoring the field amplitude at reference planes placed in the far-field, and determine the Fabry-Perot background transmissivity from plane wave excitation calculations. The results are given in Fig.~\ref{fig:bound}(b), where each data point in the figure (blue crosses) represents the maximal asymmetric coupling ratio searching over all $\vec{k}$ points in the Brillouin zone, for $P$-symmetric structural parameter choices (Fig.~\ref{fig:bound}(a)) with $h=1.5a$, $w=0.45a$, $n_0=1.45$, and varying $n_d$ and $d$ (see \href{Link}{Supplement 1} for more details). We can see that all data points obey the bound Eq.~(\ref{eq:tbound}) derived above (red solid lines). Moreover, this bound can be saturated for each value of the background transmission coefficient by appropriate optimization of the structural parameters and in-plane momentum. The blue crosses that do not saturate the bound are structures with very little perturbation from $C_2^z$ symmetry due to the choice of structural parameters. \section{Examples of Highly Asymmetric Radiation} In this section, we provide numerical examples of strong asymmetry that highlight two features of the extreme data points in Fig.~\ref{fig:bound} that are not obvious from the preceding data: it is possible to achieve high asymmetry even at the point of highest quality factor, and it is possible to achieve rapid tuning of the direction of asymmetry by slightly changing the frequency. As the form $(1+t)/(1-t)$ of the bound (for $P$-symmetric structures) suggests, strong asymmetry can be achieved when the resonance frequency coincides with locations of large transmissivity on the Fabry-Perot background, for any refraction index contrast. \begin{figure}[htb] \centering \fbox{\includegraphics[width=\linewidth]{asymexample_H1.pdf}} \caption{Examples of highly asymmetric radiation. (a) Plot of the asymmetry ratio and quality factor as a function of $k_x$, along the $k_y=0$ axis in momentum space. Strong asymmetric radiation occurs over a range of momenta, including the point of highest quality factor. \textcolor{black}{Inset: log scale plot of the $z$-component of the electric field amplitude at the highest asymmetry point;} (b) Similar plot for a different set of parameters, showing rapid switching of asymmetric direction by tuning the frequency or in-plane momentum. \label{fig:maxasym}} \end{figure} We optimize over the structural parameters shown in Fig.~\ref{fig:bound}(a) to find examples of high asymmetry in coupling to the top and bottom. This example consists of the second transverse electric (TE) polarization band (nonzero $E_z, E_x, H_y$, classified by mirror-symmetry with respect to the $x-z$ plane) of a 1D photonic crystal with structural parameters $h=1.5a$, $w=0.45a$, $n_0=n_d=1.45$, $d=0.3a$, as defined in Fig.~\ref{fig:bound}(a). See \href{link}{Supplement 1} for a plot of the band structure. The resulting asymmetry ratio and quality factor ($Q$) as a function of the in-plane $k_x$, along with the radiation field distribution at maximal asymmetry, are shown in Fig.~\ref{fig:maxasym}(a). The resonance frequency lies very close to a point of full transmission on the Fabry-Perot background, and exhibits an asymmetry exceeding $10^4$ at the $\vec{k}_\parallel$ point of largest asymmetry as well as an asymmetry over $300$ at the point of highest quality factor. It may therefore be possible to produce a laser that preferentially emits to the top or bottom using the principles discussed above. Another application of our results is the rapid steering of the direction of light emission by slight tuning of the frequency, which could be useful for LIDARs \cite{Yaacobi2014} or antennas. We design such a structure by perturbing a bound state in the continuum (BIC) \cite{Hsu2013a,Zhen2014,Hsu2013b,Hsu2016}. BICs are localized solutions embedded in the radiation continuum, where due to destructive interference of the amplitude for decay between outgoing wave channels, the quality factor of a resonance above the light line approaches infinity. In previous work \cite{Hsu2013a,Zhen2014,Hsu2013b}, the photonic structures were chosen to have both $P$ and $C_2^z$ symmetries. With a perturbation that breaks $C_2^z$ but preserves $P$, the peak quality factor will be finite but still very high \cite{Hsu2013b}. We expect that this symmetry breaking will also split the momenta where radiation towards the top and towards the bottom vanish, thereby creating strong asymmetry in the two directions, with the extrema separated only by a small $\vec{k}_\parallel$. We choose $h=1.5a$, $w=0.45a$, $n_0=1.45$, $d=0.1a$, $n_d=1.1$, again examining the second TE band. The resulting asymmetry ratio and frequency are shown in Fig.~\ref{fig:maxasym}(b). The asymmetric coupling flips from mostly radiating to the top to mostly radiating to the bottom (by a factor of $10^4$) when $k_x$ is changed by as small as $0.05\times 2\pi c/a$ or equivalently, when the frequency is changed by $3\times 10^{-4}\times 2\pi c/a$. The radiative quality factor of these resonances are on the order of $10^6$, so these two bands will be well-separated in emission. One can thereby envision rapid tuning of the emission direction by changing the frequency of radiation slightly. Moreover, the high $Q$ of these resonances will enable long propagation lengths for collimated emission from large areas that is important for LIDAR applications, complementing the low-$Q$ designs of conventional grating couplers. \begin{figure}[h!tb] \centering \fbox{\includegraphics[width=\linewidth]{absorb.pdf}} \caption{Perfect absorption with single-sided illumination and no backing mirror, for single-pass absorption less than $0.5\%$. (a) Schematic for perfect absorption at one incident angle and perfect transmission at the opposite incident angle. (b) Transmission, reflection and absorption spectra for no loss ($Q_{nr}=\infty$) and critical loss ($Q_{nr}=Q_r$), consistent with the theoretical results in (a). (c) Loss dependence of absorption, showing near-perfect absorption for critical coupling. \label{fig:absorb}} \end{figure} \section{Perfect Absorption with Single-Sided Illumination and No Backing Mirrors} We now discuss achieving perfect absorption in photonic crystal structures by combining the highly asymmetric coupling to different channels and matched radiative and non-radiative quality factors. Previous work on achieving perfect absorption has utilized metamaterial responses \cite{Radi2013,Landy2008,Watts2012}, interference between multiple incident directions \cite{Chong2010,Wan2011,Sun2014}, or a backing mirror to confine and trap light \cite{Tischler2006,Lin2011,Piper2014,Liu2014,Zhu2015,Piper2016,Sturmberg2016}. On the other hand, our results on achieving highly directional coupling suggest that it may be possible to achieve near-perfect absorption with single-sided illumination from the direction with strong coupling, only using dielectric structures and without the need of any backing mirrors. Intuitively, since the radiation coupling to one of the emission channels is strongly suppressed, there is only one direction to which the excited resonance can radiatively decay into. Appropriate tuning of the quality factor can then result in destructive interference towards this direction, thus achieving perfect absorption of incident waves. To incorporate material loss in our description, we include a non-radiative decay channel, setting $\tau_{nr}$ to be finite in Eq.~(\ref{eq:TCMT1}). We assume that the loss rate is small, and that the direct transmission pathway is not affected by the loss \cite{Piper2014}. We start from a $P$-symmetric structure and incident direction where the coupling rates have a large asymmetry ratio. The parameters are chosen such that $t$ is close to 1, and the asymmetry ratio saturates the bound $|d_2|/|d_1|=\sqrt{(1+t)/(1-t)}$. As shown in \href{Link}{Supplement 1}, $P$-symmetry and the time-reversal constraints Eqs.~(\ref{eq:fulld1}, \ref{eq:fulld2}) imply that $d_1=d_4=\sqrt{(1-t)/\tau}\exp(-3\pi j/4)$, $d_2=d_3=\sqrt{(1+t)/\tau}\exp(-\pi j/4)$. Plugging this into the expressions for the power transmission and reflection coefficients, we find that on resonance $\omega=\omega_0$ and at critical coupling $\tau=\tau_{nr}$, \begin{align} R_{14}&=R_{23}=\frac{r^2}{4},\quad T_{14}=\left(\frac{1+t}{2}\right)^2,\quad T_{23}=\left(\frac{1-t}{2}\right)^2,\\ A_{14}&=1-R_{14}-T_{14}=\frac{1-t}{2},\quad A_{23}=\frac{1+t}{2}. \end{align} In the limit where $t\rightarrow 1$, $r\rightarrow 0$, it follows that light incident from port 1 or 4 will be completely transmitted ($T_{14}=1$), while light incident from port 2 or 3 will be completely absorbed ($A_{23}=1-R_{23}-T_{23}=1$). A schematic of the resulting transmission and absorption characteristics is shown in Fig.~\ref{fig:absorb}(a). To verify these analytical results, we performed numerical simulations of the transmission and reflection spectrum with the rigorous coupled-wave analysis (RCWA) method using a freely available software package \cite{Liu2012}. The structural parameters are identical to the simulation in Fig.~\ref{fig:maxasym}(a), with the difference being the addition of loss in the system. There is a slight shift of the resonance location relative to the Fabry-Perot background due to the different discretization schemes used in the FDTD and RCWA methods. Fig.~\ref{fig:absorb} shows the simulation results. As shown in Fig.~\ref{fig:absorb}(b), when no loss is present (top panel), the transmissivity is close to 1 in the vicinity of the resonance, reaching full transmission $T=1$ at a single point on the Fano resonance as required by $P$-symmetry of the structure (see end of Section 2 for a discussion), and the reflectance is close to 0. With the addition of loss and with waves incident from the port with stronger coupling (incident direction is $\theta=3.5^\circ$ from normal, middle panel), the transmittance is reduced, and at critical coupling the transmittance drops to 0 for the resonance frequency, resulting in full-absorption of the incoming waves. On the other hand, for the same lossy structure and for the opposite incident direction (bottom panel), there is negligible absorption and most of the waves are transmitted. In Fig.~\ref{fig:absorb}(c), we show the maximum absorption for a given incident angle as a function of loss, clearly showing a peak of near-perfect absorption at critical coupling. \textcolor{black}{For other incident angles near that of maximum absorption, the $Q$-matching condition and asymmetric radiation condition are still approximately satisfied, giving rise to high absorption, and the absorption peak will shift to different frequencies following the band dispersion, as shown in \href{link}{Supplement 1}, Fig.~\ref{fig:varyangle}.} Numerically, we find that when light is incident from the port with stronger coupling, the absorption can be as high as $A_{23}=99.8\%$; when light is incident from the port with weaker coupling, the transmission is $T_{14}=99.4\%$ while absorption is only $A_{14}=0.5\%$ (this is roughly equal to the single-pass absorption rate of $0.4\%$ in our simulations). The numerical simulations show excellent agreement with our TCMT predictions for a background Fabry-Perot transmissivity of $t=0.996$ (Fig.~\ref{fig:absorb}(b)); the small difference comes from the contribution of absorption to the Fabry-Perot background. Further numerical optimization placing the resonance frequency closer to the frequency of full transmission on the Fabry-Perot background could further increase on-resonance absorption in the desired port. These results are widely applicable to many different absorbing materials. The wide range of achievable resonance quality factors, as discussed in the preceding section, implies that for both strong and weak absorbers, structures can be designed such that there is highly asymmetric coupling and critical coupling, yielding perfect absorption in the system. While our simulations were performed assuming a material with a spatially uniform absorption profile, the generality of the TCMT formalism ensures that it is also applicable to scenarios with only a thin active layer with absorption, such as with 2D materials \cite{Mak2010,Piper2014,Piper2016}. \section{Discussion and Conclusion} In conclusion, we developed a temporal coupled-mode theory formalism for general dielectric photonic crystal slab structures with arbitrary in-plane wavevectors, adequately taking into account the time-reversal-symmetry related pair of resonances and coupling channels. Using this formalism, we derived general bounds on the asymmetric radiation rates to the top and bottom of a photonic crystal slab. We then used the intuitions developed from these bounds to show examples of highly asymmetric radiation from inversion-symmetric photonic crystal slabs, demonstrating strong asymmetry, rapid tuning, and a variety of quality factors for different applications. Moreover, we showed how the highly asymmetric coupling to the top and bottom of photonic crystal slabs can be used to achieve perfect absorption for light incident from a single side, for a single-pass absorption rate of less than $0.5\%$, without the need of a back-reflection mirror as in conventional setups. The highly directional radiation could greatly benefit applications such as PCSELs, grating couplers, and LIDARs, while achieving perfect absorption without the need of back-reflection mirrors could increase the efficiency and simplify the design of photodetectors and solar cells. \textcolor{black}{While our numerical examples focused on a particular structural design, the general principle of breaking $C_2^z$ symmetry is applicable to a wide range of structures. We now briefly discuss how to implement such structures using readily available fabrication techniques. For example, gratings with slanted walls share the same structural symmetries as those in Fig.~\ref{fig:bound}(a), and thus can approach perfect single-sided radiation and absorption as well. Such gratings can be fabricated using focused ion beam milling \cite{Schrauwen2007}, angled-etching with Faraday cages \cite{Burek2012}, or inclined lithography \cite{DelCampo2008}. More generally, almost any of the techniques for fabricating blazed gratings \cite{Fujita1982,Palmer2005} or constructing 3D photonic crystals \cite{Joannopoulos2011} (e.g. layer-by-layer lithography or holographic lithography) could also be employed in a simplified form to make an asymmetric coating. A wide range of structures breaking $C_2^z$ symmetry and achieving high asymmetry can thus be easily realized with these different techniques.} Our work provides new design principles for achieving highly directional radiation and perfect absorption in photonics, and could be extended to systems where there are nonlinearities, gain and loss, different substrates and superstrates, and non-reciprocal structures with magneto-optical effects \cite{Wang2005}. Our work can also be generalized to other systems characterized by temporal coupled-mode theory, such as in-plane chiral meta-surfaces, asymmetric ring resonators, and scattering from nano-plasmonic structures. \section*{Funding and Acknowledgements} Institute for Soldier Nanotechnologies (W911NF-13-D0001). Solid State Solar Thermal Energy Center (DE-SC0001299). United States-Israel Binational Science Foundation (2013508). National Science Foundation (DMR-1307632). We thank Yong Liang, Ling Lu, Scott Skirlo, Yichen Shen, Aviram Massuda, Emma Regan, Francisco Machado and Nicholas Rivera for helpful discussions. See appendix for supporting material.
f1655213d3dd10a4d942c00b15957b0fe391fb55
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Two gravitational wave signals GW150914~\cite{Ligo} and GW151226~\cite{GW151226} generated by black hole merging systems were detected recently by the Laser Interferometer Gravitational-Wave Observatory. This confirms directly a prediction of Einstein based on his classical theory of general relativity~\cite{Einstein} regarding the existence of gravitational waves which are spacetime ripples propagating through the universe. Quantum mechanically, gravitational interaction is presumably mediated by gravitons when gravity is quantized. However, a full theory of quantum gravity is elusive at the present. Even though, general relativity as an effective field theory provides a framework to probe the low energy quantum gavity effects. In this respect, it has been found that there exists a quantum correction to the Newtonian force between two mass monopoles which behaves as $r^{-3}$~\cite{Donoghue}, and this is obtained by summing one-loop Feynman diagrams with off-shell gravitons. Recently, the quantum gravity correction to classical forces was extended to include the quadrupole-quadrupole interaction between a pair of polarizable objects from their induced quadrupole moments due to two-graviton exchange~\cite{Ford16}. This correction is computed by first finding out the normal field modes which keep the cycle in phase in which at the outset one object is polarized and radiates a gravitational field which polarizes the second and induces a quadrupole of which the gravitational field in return polarizes the first, and then summing up the zero-point energy of all these normal modes to get the interaction potential between two quadrupoles. This is in close analogy to that in the computation of the Casimir-Polder and van der Waals (vdW) forces between a pair of atoms from their induced dipole moments due to two photon exchange~\cite{Sernelius}. An advantage of this method is that the details of quantization of the gravitational metric are not needed. Although a full theory of quantum gravity is absent, one can still use linearized quantum gravity to find quantum gravitational corrections to classical physics which an ultimate quantum gravity theory must produce at low energies. One such example is the quantum light-cone fluctuations produced by gravitons propagating on a background spacetime~\cite{Yu99, Yu00, Yu09}. When gravity is quantized, there inevitably exist quantum gravitational fluctuations which induce quadrupoles in gravitationally polarizable objects, thus giving rise to a quantum correction to the classical interaction between polarizable objects. In this paper, we present in the framework of linearized quantum gravity a different field-theoretic approach to the computation of the quantum gravity correction to the classical gravitational interaction between a pair of polarizable objects. Our approach, which is parallel to that used by Casimir and Polder in studying the quantum electromagnetic vacuum fluctuation induced electric dipole-dipole interaction between two neutral atoms in a quantum theory of electromagnetism~\cite{Casimir}, is based upon the leading-order perturbation theory. Let us note that a rather simple perturbative calculation of both the retarded and instantaneous electromagnetic interaction can also be found in~\cite{Holstein2001}. The system we consider consists of two gravitationally polarizable objects in a bath of fluctuating quantum gravitational fields in vacuum. We assume that the two objects labeled as $A$ and $B$ can be treated as two-level harmonic oscillators. Then, the total Hamiltonian can be give by \bea{H} H=H_{F} + H_{A}+H_{B}+H_{AF} +H_{BF}\ , \end{eqnarray} where $H_{F}$ is the Hamiltonian of gravitational fields, the Hamiltonian of the object $H_{A(B)}$ takes the form \bea{} H_{A(B)}=E_{A(B)}^{0} |0_{A(B)}\rangle\langle 0_{A(B)}|+E_{A(B)}^{1} |1_{A(B)}\rangle\langle 1_{A(B)}|\ , \end{eqnarray} and $H_{A(B)F}$ represents the interactions between the objects and gravitational fields \bea{HA} {H}_{A(B)F}=-\frac{1}{2} {Q}_{A(B)}^{ij} {E}_{ij}\;. \end{eqnarray} Here the Einstein convention is assumed for repeated indices, Latin indices run from 1 to 3, $ {Q}_{ij}$ is the gravitational vacuum fluctuation induced quadrupole moment of the object and $E_{ij}=R_{0i0j}$ is the gravito-electric tensor with $R_{\mu\nu\alpha\beta}$ being the Riemann tensor. $E_{ij}$, which is defined by an analogy between the linearized Einstein field equations and the Maxwell equations~\cite{Campbell}, determines the tidal gravitational acceleration between two nearby test particles in the classical Newtonian gravity. For a flat background spacetime with a linearized perturbation $h_{\mu\nu}$ propagating upon it, the metric can be expanded as $g_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}$. From the definition of $E_{ij}$ one can obtain \bea{Eij} E_{ij}=\frac{1}{2}\ddot{h}_{ij}\;, \end{eqnarray} where a dot denotes a derivative with respect to time $t$. In the transverse tracefree (TT) gauge, the gravitational field can be quantized as~\cite{Yu99} \bea{hij} h_{ij}({\bf r}, t)=\sum_{{\bf k}, \lambda}[a_{\lambda}(\omega, {\bf r}) e_{ij}({\bf k}, \lambda) f_{{\bf k}}+H. c.]\;, \end{eqnarray} where $H.c.$ denotes the Hermitian conjugate, $a_{\lambda}(\omega, {\bf r}) $ is the gravitational field operator, $\lambda$ labels the polarization states, $e_{ij}({\bf k}, \lambda)$ are polarization tensors, and \bea{fk} f_{{\bf k}}=\frac{1}{\sqrt{2\omega (2\pi)^{3}}} e^{i ({\bf k} \cdot {\bf r}- \omega t)} \end{eqnarray} is the field mode with \bea{wk} \omega=|{\bf k}|=(k_{x}^{2}+ k_{y}^{2} +k_{z}^{2})^{1/2}\ . \end{eqnarray} Here we choose the units in which $32\pi G=1$ and $\hbar=c=1$, where $G$ is the Newton's gravitational constant. The vacuum state of gravitational fields is defined as \bea{} a_{\lambda}(\omega, {\bf r}) | \{0\}\rangle=0\ , \end{eqnarray} and the single- and two-graviton excited states are \bea{} a^{\dagger}_{\lambda_{{\alpha}}}(\omega_{\alpha}, {\bf r}_{\alpha}) |\{0\}\rangle&=&|1^{(\alpha)}\rangle\ , \\ \frac{1}{\sqrt{2}}a^{\dagger}_{\lambda_{{\alpha}}}(\omega_{\alpha}, {\bf r}_{\alpha}) a^{\dagger}_{\lambda_{{\beta}}}(\omega_{\beta}, {\bf r}_{\beta}) |\{0\}\rangle&=& |1^{(\alpha)}, 1^{(\beta)}\rangle \ . \end{eqnarray} Since the object-gravitational field coupling is linear in the object and the field variables, each object must interact with the field at least two times and return to its ground state. Thus, the position-dependent shift of the ground-state energy between two objects arises from fourth-order perturbations~\cite{Casimir, Buhmann, Safari} in the leading-order perturbation theory \bea{ES} \Delta E_{AB}=&-&{\sum_{\mathrm{I,II,III}}}' \frac{\langle 0| \hat{H}_{AF}+ \hat{H}_{BF}| \mathrm{I} |\langle\mathrm{I} | \hat{H}_{AF}+ \hat{H}_{BF}| \mathrm{II} \rangle}{(E_{\mathrm{I}}-E_{0})(E_{\mathrm{II}}-E_{0})} \nonumber \\ && \qquad \times \frac { \langle\mathrm{II} | \hat{H}_{AF}+ \hat{H}_{BF}| \mathrm{III} \rangle \langle\mathrm{III} | \hat{H}_{AF}+ \hat{H}_{BF}| 0\rangle}{(E_{\mathrm{III}}-E_{0})}\;, \end{eqnarray} where the primed sum means that the ground state of the whole system $|0\rangle= |0_{A}\rangle |0_{B}\rangle|\{0\}\rangle$ is omitted and the summation includes position and frequency integrals. During each interaction, a graviton may either be emitted or absorbed by an object. The intermediate states $| \mathrm{I} \rangle$ and $| \mathrm{III} \rangle$ must consist of an excited object and a single graviton. While for the intermediate state $| \mathrm{II} \rangle$ there are three possibilities, i.e., both objects in the ground state with two gravitons, both objects excited with no graviton or both objects excited with two gravitons. Therefore, there are ten possible combinations of intermediate states, which are listed in Table.~(\ref{Tab1}). For the case (1) in Table.~(\ref{Tab1}), substituting $|\mathrm I\rangle$, $|\mathrm{ II}\rangle$ and $|\mathrm{III}\rangle$ into Eq.~(\ref{ES}) yields \bea{} \Delta E_{AB(1)}({\bf r}_{A}, {\bf r}_{B}) &=& -\frac{1}{16} \int_{0}^{\infty}d\omega \int_{0}^{\infty}d\omega' \bigg(\frac{1}{D_{ \mathrm i}} +\frac{1}{D_{\mathrm {ii}}} \bigg) \nonumber \\ && \times \tilde{Q}_{A}^{ij} \ \tilde{Q}_{A}^{*kl}\ \tilde{Q}_{B}^{ab}\ \tilde{Q}_{B}^{*cd}\ G_{ijab}(\omega, {\bf r}_{A}, {\bf r}_{B}) G_{klcd}(\omega', {\bf r}_{A}, {\bf r}_{B})\ , \end{eqnarray} where $\tilde{Q}_{A(B)}^{ij}= \langle 0_{A(B)}| {Q}_{A(B)}^{ij} | 1_{A(B)}\rangle$, $\tilde{Q}_{A(B)}^{*ij}= \langle 1_{A(B)}| {Q}_{A(B)}^{ij} | 0_{A(B)}\rangle$, and \bea{fijkl} G_{klcd}(\omega, {\bf r}_{A}, {\bf r}_{B})=\langle 0| {E}_{kl}(\omega, {\bf r}_{A}) {E}_{cd}(\omega, {\bf r}_{B}) |0\rangle \end{eqnarray} is the two-point correlation function of gravito-electric fields. The expressions of $D_{\mathrm i}$ and $D_{\mathrm {ii}}$ are given in Tab.~(\ref{Tab1}) in which $\omega_{A(B)}=(\omega_{A(B)}^{1}-\omega_{A(B)}^{0})$ with $\omega_{A(B)}^{1}=E_{A(B)}^{1}$ and $\omega_{A(B)}^{0}=E_{A(B)}^{0}$, which represents the transition frequency of the object. The contributions of other cases in Table.~(\ref{Tab1}) to $\Delta E_{AB}$ can be calculated similarly. Summing up all possible intermediate states, we obtain that \bea{} \Delta E_{AB}({\bf r}_{A}, {\bf r}_{B}) &=& -\frac{1}{16} \int_{0}^{\infty}d\omega \int_{0}^{\infty}d\omega' \sum_{n=\mathrm i}^{\mathrm {xii} } \frac{1}{D_{n}} \nonumber \\ && \times \tilde{Q}_{A}^{ij} \ \tilde{Q}_{A}^{*kl}\ \tilde{Q}_{B}^{ab}\ \tilde{Q}_{B}^{*cd}\ G_{ijab}(\omega, {\bf r}_{A}, {\bf r}_{B}) G_{klcd}(\omega', {\bf r}_{A}, {\bf r}_{B})\ . \end{eqnarray} It is easy to show \bea{} \sum_{n=\mathrm i}^{\mathrm { xii}} \frac{1}{D_{n}}= \frac{4(\omega_{A}+\omega_{B}+\omega)}{(\omega_{A}+\omega_{B})(\omega_{A}+\omega)(\omega_{B}+\omega)} \bigg(\frac{1}{\omega+\omega'}-\frac{1}{\omega-\omega'} \bigg)\ . \end{eqnarray} Since the shift of the ground-state energy is just the vdW-like potential $U_{AB}({\bf r}_{A}, {\bf r}_{B})$ between two polarizable objects, one has from the above two equations, \bea{UAB} U_{AB}({\bf r}_{A}, {\bf r}_{B}) &=& -\frac{1}{4} \int_{0}^{\infty}d\omega \int_{0}^{\infty}d\omega' \frac{\omega_{A}+\omega_{B}+\omega}{(\omega_{A}+\omega_{B})(\omega_{A}+\omega)(\omega_{B}+\omega)} \bigg(\frac{1}{\omega+\omega'}-\frac{1}{\omega-\omega'} \bigg) \nonumber \\ &&\qquad \qquad \times \tilde{Q}_{A}^{ij} \ \tilde{Q}_{A}^{*kl}\ \tilde{Q}_{B}^{ab}\ \tilde{Q}_{B}^{*cd}\ G_{ijab}(\omega, {\bf r}_{A}, {\bf r}_{B}) G_{klcd}(\omega', {\bf r}_{A}, {\bf r}_{B})\ . \end{eqnarray} Assuming that the objects A and B are isotropically polarizable , we have \bea{Q} \tilde{Q}_{A(B)}^{ij} \tilde{Q}_{A(B)}^{*kl} = 2 |\tilde{Q}_{A(B)}^{ij}|^{2} \delta_{ik}\delta_{jl}\equiv \frac{1}{2} \tilde{\alpha}_{A(B)} \delta_{ik}\delta_{jl}\ . \end{eqnarray} Here $\tilde{\alpha} (\omega)\equiv \frac{1}{4}|\tilde{Q}_{ij}(\omega) |^{2}$. Then, Eq.~(\ref{UAB}) can be simplified as \bea{UAB0} U_{AB}({\bf r}_{A}, {\bf r}_{B}) &=& -\frac{1 }{16(\omega_{A}+\omega_{B})} \int_{0}^{\infty}d\omega \int_{0}^{\infty}d\omega' \;\frac{ \tilde{\alpha}_{A} \tilde{\alpha}_{B}(\omega_{A}+\omega_{B}+\omega)}{(\omega_{A}+\omega)(\omega_{B}+\omega)} \bigg(\frac{1}{\omega+\omega'}-\frac{1}{\omega-\omega'} \bigg) \nonumber \\ &&\qquad \qquad \times G_{ijab}(\omega, {\bf r}_{A}, {\bf r}_{B}) G_{ij ab }(\omega', {\bf r}_{A}, {\bf r}_{B})\ . \end{eqnarray} The two-point correlation function $G_{ijab}(\omega, {\bf r}_{A}, {\bf r}_{B}) $ can be obtained from $G_{ijab}( {\bf r}_{A}, {\bf r}_{B}, {t_{A}}, t_{B})$ by Fourier transform. From Eqs.~(\ref{Eij}), (\ref{hij}), and (\ref{fijkl}), one finds \bea{} G_{ijkl}({\bf r}, {\bf r}', t, t') &=& \frac{1}{4}\langle 0| \ddot {h}_{ij}( {\bf r}, t) \ddot{h}_{kl}( {\bf r}', t') |0\rangle \nonumber \\ &=& \frac{1}{4 (2\pi)^{3}}\int d^{3} {\bf k} \sum_{\lambda} e_{ij}({\bf k}, \lambda) e_{kl}({\bf k}, \lambda) \frac{\omega^{3}}{2} e^{i{\bf k}\cdot ({\bf r-r'})- i\omega(t-t')} \ .\end{eqnarray} The summation of polarization tensors in the TT gauge gives~\cite{Yu99} \bea{eijekl} \sum_{\lambda}\, e_{ij} ({{\bf k}, \lambda}) e_{kl} ({{\bf k}, \lambda})&=&\delta_{ik}\delta_{jl} +\delta_{il}\delta_{jk}-\delta_{ij}\delta_{kl} +\hat k_i\hat k_j \hat k_k\hat k_l+\hat k_i \hat k_j \delta_{kl} \nonumber\\ &&+\hat k_k \hat k_l \delta_{ij}-\hat k_i \hat k_l \delta_{jk} -\hat k_i \hat k_k \delta_{jl}-\hat k_j \hat k_l \delta_{ik}-\hat k_j \hat k_k \delta_{il}\,, \end{eqnarray} where \begin{equation} \hat k_i=\frac{ k_i}{ k}\,. \end{equation} Transforming to the spherical coordinate: $\hat{k}_{1}=\sin \theta \cos \varphi$, $\hat{k}_{2}=\sin \theta \sin \varphi$ and $\hat{k}_{3}=\cos \theta $, and letting \bea{eijekl} \sum_{\lambda}\, e_{ij} ({{\bf k}, \lambda}) e_{kl} ({{\bf k}, \lambda})=g_{ijkl}(\theta, \varphi)\,, \end{eqnarray} we arrive at \bea{AG1} G_{ijkl}(r, \Delta t) = \frac{1}{8(2\pi)^{3} }\int_{0}^{\infty} \omega^{5} d \omega \int_{0}^{\pi} \sin\theta d\theta \int_{0}^{2\pi} d \varphi \ g_{ijkl}(\theta, \varphi) e^{i \omega (r \cos\theta- \Delta t)} \ ,\end{eqnarray} where $r=\Delta {\bf r}= |{\bf r}_{A}- {\bf r}_{B}|$ is the distance between objects A and B. Fourier transforming the above expression yields \bea{G16} G_{ijkl}(\omega, {\bf r}_{A}, {\bf r}_{B}) & =& \frac{1}{2\pi} \int_{-\infty}^{\infty} d \Delta t e^{i\omega \Delta t} G_{ijkl}( { r}, {\Delta t}) \nonumber \\ & = & \frac{\omega^{5}}{8(2\pi)^{3} } \int_{0}^{\pi} \sin\theta d\theta \int_{0}^{2\pi} d \varphi \ g_{ijkl}(\theta, \varphi) e^{i w r \cos\theta }\ , \end{eqnarray} Substituting Eq.~(\ref{G16}) into Eq.~(\ref{UAB0}) and integrating over ($\theta, \varphi, \theta', \varphi'$), we obtain \bea{UAB2} U_{AB}(r) &=& -\frac{1 }{2^{9} \pi^{4}(\omega_{A}+\omega_{B}) r^{10}} \int_{0}^{\infty}d\omega \int_{0}^{\infty}d\omega' \;\frac{ \tilde{ \alpha}_{A} \tilde{\alpha}_{B}(\omega_{A}+\omega_{B}+\omega) }{(\omega_{A}+\omega)(\omega_{B}+\omega)} \bigg(\frac{1}{\omega+\omega'}-\frac{1}{\omega-\omega'} \bigg) \nonumber \\ &&\quad\quad \times [A_{1}(\omega r, \omega' r)\cos(\omega' r)+B_{1}(\omega r, \omega' r)\sin(\omega' r)]\ , \end{eqnarray} where \bea{A1} A_{1}(x, x') &=& x x' (315 +8 x^{2}x'^{2}-30x^{2}- 30x'^{2})\cos x \\ \nonumber &+& x'(-315+ 135x^2+30 x'^{2} - 18 x^{2} x'^{2}- 3x^{4} + 2 x^{4} x'^{2})\sin x \ , \end{eqnarray} and \bea{B1} B_{1}(x, x') &=& x (-315+ 135x'^2+30 x^{2} - 18 x^{2} x'^{2}- 3x'^{4} + 2 x^{2} x'^{4})\cos x \\ \nonumber & +& (315 -135 x^{2} +3 x^{4}- 135 x'^{2}+ 63 x^{2} x'^{2}-3 x^4 x'^2 + 3 x'^4 - 3 x^2 x'^4 + x^4 x'^4)\sin x \ . \end{eqnarray} Since $A_{1}(x, -x')=-A_{1}(x, x') $ and $B_{1}(x, -x')=B_{1}(x, x') $, Eq.~(\ref{UAB2}) can be re-expressed as \bea{UAB3} U_{AB}(r) &=& -\frac{ 1 }{2^{9} \pi^{4}(\omega_{A}+\omega_{B}) r^{10}} \int_{0}^{\infty}d\omega\; \frac{ \tilde{ \alpha}_{A} \tilde{\alpha}_{B} (\omega_{A}+\omega_{B}+\omega) }{(\omega_{A}+\omega)(\omega_{B}+\omega)} \\ \nonumber && \times \int_{-\infty}^{\infty} d\omega' \bigg( \frac{1}{2(\omega+\omega')}+\frac{1}{2(-\omega+\omega')} \bigg) [A_{1}(\omega r, \omega' r)-iB_{1}(\omega r, \omega' r)]e^{i\omega' r} . \end{eqnarray} Performing the principle value integral on $\omega'$, one has \bea{UAB3} U_{AB}(r) &=& -\frac{1}{ 32^{2}\pi^{3} (\omega_{A}+\omega_{B}) r^{10}} \int_{0}^{\infty}d\omega\; \frac{ \tilde{ \alpha}_{A} \tilde{\alpha}_{B} (\omega_{A}+\omega_{B}+\omega) }{(\omega_{A}+\omega)(\omega_{B}+\omega)} \\ \nonumber &&\times [A_{2}(\omega r) \cos(2\omega r)+B_{2}(\omega r) \sin(2\omega r)]\ , \end{eqnarray} where \bea{A2} A_{2}(x)&=&-630 x+330 x^{3}-42 x^{5}+4x^{7} \ ,\nonumber \\ \nonumber B_{2}(x)&=&315 -585 x^{2}+ 129 x^{4} - 14 x^{6}+ x^{8} \ . \end{eqnarray} Eq.~(\ref{UAB3}) can be re-written as \bea{} U_{AB}(r) &=& -\frac{1 }{ 32^{2} \pi^{3} (\omega_{A}+\omega_{B}) r^{10}} \bigg(\int_{0}^{\infty}d\omega\; \frac{ \tilde{\alpha}_{A} \tilde{ \alpha}_{B} (\omega_{A}+\omega_{B}+\omega) }{(\omega_{A}+\omega)(\omega_{B}+\omega)} \bigg [\frac{A_{2}(\omega r)}{2} +\frac{B_{2}(\omega r)}{2i} \bigg] e^{2i\omega r} \nonumber \\ &&+\int_{0}^{-\infty}d\omega\; \frac{ \tilde{\alpha}_{A} \tilde{\alpha}_{B} (\omega_{A}+\omega_{B}-\omega) }{(\omega_{A}-\omega)(\omega_{B}-\omega)} \bigg [\frac{A_{2}(\omega r)}{2} +\frac{B_{2}(\omega r)}{2i} \bigg] e^{2i\omega r} \bigg) \ . \end{eqnarray} Simplifying the above equation by contour-integral techniques, we obtain \bea{UAB31} U_{AB}(r) &=& -\frac{ 1 }{32^{2} \pi^{3} r^{10}} \int_{0}^{\infty}d u \ {\alpha}_{A}(iu) {\alpha}_{B}(iu)[i A_{2}(i u r) + B_{2}(i u r)] e^{-2 u r} \nonumber \\ &=& -\frac{ 1 }{32^{2} \pi^{3} r^{10}} \int_{0}^{\infty}d u \ {\alpha}_{A}(iu) {\alpha}_{B}(iu) S(ur) e^{-2 u r} \ . \end{eqnarray} where \bea{} {\alpha}_{A(B)}(\omega)=\lim _{\epsilon \rightarrow0^{+}}\frac{ \tilde{\alpha}_{A(B)} \omega_{A(B)}}{\omega_{A(B)}^{2}-\omega^{2}-i \epsilon \omega}=\lim _{\epsilon \rightarrow0^{+}} \frac{1}{4}\frac{ |\tilde{Q}^{ij}_{A(B)}|^{2} \omega_{A(B)}}{\omega_{A(B)}^{2}-\omega^{2}-i \epsilon \omega}\ , \end{eqnarray} is the object's ground-state polarizability, which is defined in analogy to the electric polarizability of atoms~\cite{Buhmann1} and satisfies \bea{}Q_{ij}(\omega)= {\alpha}(\omega) E_{ij} (\omega, \bf r)\ ,\end{eqnarray} and \bea{} S(x)=315 + 630 x + 585 x^2 + 330 x^3 + 129 x^4 + 42 x^5 + 14 x^6 + 4 x^7 + x^8 \ . \end{eqnarray} In the far regime $r\gg \omega_{A(B)}^{-1}$, since there is an exponential in the integrand in Eq.~(\ref{UAB31}), small values of $u$ provide the dominating contribution. Thus, we can use approximately the static polarizability $\tilde{\alpha}_{A(B)}(0)$ and obtain \bea{FL} U_{AB}(r) &=& -\frac{3987}{ 4 \pi (32\pi)^{2} r^{11}} {\alpha}_{A}(0) {\alpha}_{B}(0) \nonumber \\ &=& -\frac{3987 \hbar c G^{2}}{ 4\pi r^{11}} {\alpha}_{A}(0) {\alpha}_{B}(0)\ . \end{eqnarray} In the near regime $r\ll \omega_{A(B)}^{-1}$, the integral in Eq.~(\ref{UAB31}) is effectively limited to a region where $e^{-2ur}\simeq 1$ and so all terms in $S(x)$ dependent on $x$ can be neglected. Thus, the potential becomes \bea{NL} U_{AB}(r) &=& -\frac{ 315}{32^{2} \pi^{3} r^{10}} \int_{0}^{\infty}d u \ {\alpha}_{A}(iu) {\alpha}_{B}(iu) \nonumber \\ &=& -\frac{ 315 \hbar G^{2}}{ \pi r^{10}} \int_{0}^{\infty}d u \ {\alpha}_{A}(iu) {\alpha}_{B}(iu) \ . \end{eqnarray} In the second line of Eqs.~(\ref{FL}) and (\ref{NL}), we return to the SI units. These results agree with that obtained with a normal mode evaluation~\cite{Ford16} and a two-graviton exchange calculation~\cite{Holstein2016} It is interesting to note here that the qualitative behavior of the effective interaction between two gravitationally polarizable objects can be inferred from a simple dimensional analysis. Dimensionally, the object's polarizability scales as $1/M^5$. Since the retarded potential energy must go as $M$ and is quadratic in the polarizability, the $r$ dependence must be $ r^{-11}$. In the instantaneous case there is an additional factor of the excitation energy involved, so the scaling is as $ r^{-10}$. In conclusion, based upon the quantum theory of linearized gravity and the leading-order perturbation theory, we have studied the quantum gravitational vacuum fluctuation induced quadrupole-quadrupole quantum correction to the classical Newtonian force between a pair of polarizable objects. We find that the interaction potential decays as $r^{-11}$ in the retarded region and $r^{-10}$ in the near region. Our approach parallels that of Casimir and Polder in the investigation of the quantum electromagnetic vacuum fluctuation induced dipole-dipole interaction between two neutral atoms in a quantum theory of electromagnetism and suggests that it may be robust to use linearized quantum gravity to study quantum gravitational effects at low energies, such as those associated with quantized gravitational waves in the same way as one usually does with quantized electromagnetic waves. \acknowledgments This work was supported by the National Natural Science Foundation of China under Grants No. 11435006, No. 11375092 and No. 11447022; and the Zhejiang Provincial Natural Science Foundation of China under Grant No. LQ15A050001. \begin{table}[ht] \begin{center} \begin{tabular}{cllll} \hline Case & $|\mathrm I\rangle$ & \hspace{1ex} $|\mathrm{II}\rangle$ & \hspace{1ex} $\hspace{-1ex} |\mathrm{III}\rangle$ & Denominator\\ \hline ($1$) & $|1_A,0_B\rangle |1^{(1)}\rangle$ &\hspace{1ex} $|0_A,0_B\rangle |1^{(2)},1^{(3)}\rangle$ & \hspace{1ex}$|0_A,1_B\rangle |1^{(4)}\rangle$ & $D_{\mathrm {i}}=(\omega_A+\omega') (\omega'+\omega)(\omega_B+\omega')$, \\ {} & ${}$ & ${}$ & ${}$ & $D_{\mathrm {ii}}=(\omega_A+\omega') (\omega'+\omega)(\omega_B+\omega)$ \\ ($2$) &$|1_A,0_B\rangle |1^{(1)}\rangle$ & \hspace{2ex}$|1_A,1_B\rangle |\{0\}\rangle$ & \hspace{1ex}$|0_A,1_B\rangle |1^{(2)}\rangle$ & $D_{\mathrm {iii}}=(\omega_A+\omega') (\omega_A+\omega_B) (\omega_B+\omega)$\\ ($3$) &$ |1_A,0_B\rangle |1^{(1)}\rangle$ & \hspace{2ex} $|1_A,1_B\rangle |\{0\}\rangle$ &\hspace{1ex}$|1_A,0_B\rangle |1^{(2)}\rangle$ & $D_{\mathrm {iv}}=(\omega_A+\omega') (\omega_A+\omega_B) (\omega_A+\omega)$\\ ($4$) & $|1_A,0_B\rangle |1^{(1)}\rangle$ & \hspace{2ex}$|1_A,1_B\rangle |1^{(2)},1^{(3)}\rangle$ & \hspace{1ex}$|0_A,1_B\rangle |1^{(4)}\rangle$ & $D_{\mathrm {v}}=(\omega_A+\omega') (\omega_A+\omega_B+\omega'+ \omega) (\omega_B+\omega')$\\ ($5$) & $|1_A,0_B\rangle |1^{(1)}\rangle$ & \hspace{2ex} $|1_A,1_B\rangle |1^{(2)}, 1^{(3)}\rangle$ & \hspace{1ex}$|1_A,0_B\rangle |1^{(4)}\rangle$ & $D_{\mathrm {vi}}=(\omega_A+\omega') (\omega_A+\omega_B+ \omega'+\omega) (\omega_A+\omega)$\\ ($6$) & $|0_A,1_B\rangle |1^{(1)}\rangle$ & \hspace{2ex}$|0_A,0_B\rangle |1^{(2)},1^{(3)}\rangle$ & \hspace{1ex}$|1_A,0_B\rangle |1^{(4)}\rangle$ & $D_{\mathrm {vii}}=(\omega_B+\omega') (\omega'+\omega)(\omega_A+\omega')$, \\ {} & ${}$ & ${}$ & ${}$ & $D_{\mathrm{viii}}=(\omega_B+\omega') (\omega'+\omega)(\omega_A+\omega)$ \\ ($7$) & $|0_A,1_B\rangle |1^{(1)}\rangle$ &\hspace{1ex} $|1_A,1_B\rangle |\{0\}\rangle$ & \hspace{1ex}$|1_A,0_B\rangle |1^{(2)}\rangle$ & $D_{\mathrm {ix}}=(\omega_B+\omega') (\omega_A+\omega_B) (\omega_A+\omega)$\\ ($8$) & $|0_A,1_B\rangle |1^{(1)}\rangle$ & \hspace{2ex} $|1_A,1_B\rangle |\{0\}\rangle$ & \hspace{1ex}$|0_A,1_B\rangle |1^{(2)}\rangle$ & $D_{\mathrm {x}}=(\omega_B+\omega') (\omega_A+\omega_B) (\omega_B+\omega)$\\ ($9$) & $|0_A,1_B\rangle |1^{(1)}\rangle$ & \hspace{2ex}$|1_A,1_B\rangle |1^{(2)},1^{(3)}\rangle$ & \hspace{1ex}$|1_A,0_B\rangle |1^{(4)}\rangle$ & $D_{\mathrm {xi}}=(\omega_B+\omega') (\omega_A+\omega_B+\omega'+\omega) (\omega_A+\omega')$\\ ($10$) & $|0_A,1_B\rangle |1^{(1)}\rangle$ & \hspace{2ex} $|1_A,1_B\rangle |1^{(2)}, 1^{(3)}\rangle$ & \hspace{1ex}$|0_A,1_B\rangle |1^{(4)}\rangle$ & $D_{\mathrm {xii}}=(\omega_B+\omega') (\omega_A+\omega_B+\omega'+ \omega) (\omega_B+\omega)$\\ \hline \end{tabular} \caption{ \label{Tab1} Intermediate states contributing to the two-objects potential and corresponding denominators. } \end{center} \end{table}
5fadb3d7db6aeb74774979309120004d2bbcaf90
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:intro} The discovery of a light Higgs boson~\cite{higgs,discovery} has opened a major new avenue in experimental and theoretical particle physics: comprehensive tests of a possible non-minimal fundamental scalar sector, for which there exists a plethora of motivations. While there has been a lot of progress in developing combined Higgs and gauge analysis strategies for the LHC Run~II~\cite{barca,legacy1,legacy2,other}, there exists no general and proven analysis framework even for a Higgs portal model~\cite{portal}. The announcement of an excess seen in the di-photon spectrum by both ATLAS and CMS~\cite{750ex,S750_1000,S750_2100}, if confirmed by future data, suggests such an extended scalar sector. The anomaly has led to an excessive number of publications, so we feel that adding one more, and hopefully useful publication can be justified somehow\footnote{Beyond the di-photon anomaly we present the first full \textsc{SFitter} analysis of a Higgs portal allowing for higher-dimensional operators.}. Early studies of the anomaly in an effective theory framework can be found in Ref.~\cite{not_eft}. Intriguingly, an additional scalar is not sufficient to explain the signal in complete models. The new scalar's sizable couplings to photons and gluons need to be induced by relatively light new particles~\cite{not_vectors}. For example in supersymmetric models, vector-like matter added to the MSSM or non-trivial signatures in the NMSSM are necessary for a successful explanation of the excess~\cite{not_susy}. In models in which these new states are connected to the SUSY breaking sector the new scalar can be identified with the sgoldstino, implying a very low SUSY breaking scale~\cite{not_sgoldstino}. Other extended spacetime symmetries give rise to dilaton~\cite{not_dilaton} and radion interpretations~\cite{not_radion}, which imply similarly unintended consequences, such as low ultraviolet (UV) scales or a very large curvature of the extra dimension. Extra dimensional scalars not directly related to the compactification circumvent this problem~\cite{not_extrad1} and can explain the localization of extra dimensional fermions, which makes the new scalar a localizer field~\cite{not_extrad2}. Related models, which consider the electroweak scale (or the TeV scale) arising from composite dynamics are less constrained than the MSSM, due to the large number of potential scalar resonances and fermionic quark partners~\cite{not_composite}. The possibility of the new resonance to be a spin 2 particle, associated with a higher dimensional theory of gravity is strongly constrained by dilepton searches and just like the radion implies sizable curvature terms~\cite{not_spin2}. The large width of $\Gamma_S=45$~GeV, as reported by ATLAS, can be addressed in some models~\cite{not_interference1,not_wide}; while such a large width only slightly increases the statistical significance, if it is true, background interference effects are important~\cite{not_interference2}. In this case, it is well motivated to assume that the new scalar provides a portal to a dark sector, inducing a sizable width through invisible decays~\cite{not_dm}. Alternatively, it might be the sign of cascade decays or other explanations not based on a single scalar resonance, which lead to cusps and endpoint structures that can fake a large width~\cite{not_exotics}. The very minimal, yet not perturbatively realizable assumption of photon fusion induced production can not explain a large width~\cite{not_photon}. The new resonance could also be related to the various, persistent flavor anomalies~\cite{not_flavor}, to the mechanism behind the electroweak phase transition~\cite{not_phase}, the strong CP problem~\cite{not_strong}, or an underlying string theory~\cite{not_strings}. Finally, a variety of models, motivated by different extensions of the Standard Model (SM) not fitting in the above categories, and further measurements testing the properties of the new resonance have been proposed~\cite{not_stuff}.\medskip In spite of all these considerations, the most obvious question is whether such an additional, likely scalar resonance can be part of an extended Higgs sector~\cite{not_higgs,Dawson:2016ugw}; in other words, if the new scalar can form a Higgs portal, possibly to a new sector. To answer this question we will remain agnostic about the underlying physics, but assume that a resonantly produced narrow scalar singlet is responsible for the excess. We couple the new scalar to the SM through an effective Lagrangian. This assumption exactly corresponds to recent developments on how to describe deviations from the Standard Model Higgs sector at the LHC~\cite{eftfoundations,eftorig,higgsreview,barca,legacy1,legacy2,other,Grzadkowski:2010es}. The combined Higgs portal Lagrangian is organized by the field content, the symmetry structure, and the mass dimension. This way we can contrast the apparent absence of dimension--six effects in the range $\Lambda \approx 300~...~500$~GeV for the SM-like Higgs and gauge sector~\cite{legacy2} with the need for higher-dimensional operators coupling to the new scalar with $\Lambda \lesssim 1$~TeV. \subsection{Theoretical framework} \label{sec:lag} The most general linear effective Lagrangian up to dimension six and built from Standard Model particles and a new scalar singlet reads \begin{align} \ensuremath{\mathcal{L}} = \ensuremath{\mathcal{L}}_\text{SM} + \ensuremath{\mathcal{L}}_{\text{dim-6}}^H + \ensuremath{\mathcal{L}}_{\text{dim}\leq 5}^S+\ensuremath{\mathcal{L}}_{\text{dim-6}}^S \; . \label{eq:lag1} \end{align} Here, $\ensuremath{\mathcal{L}}_\text{SM}$ stands for the renormalizable SM Lagrangian, while $\ensuremath{\mathcal{L}}_{\text{dim-6}}^H$ contains the dimension--six operators made out of SM fields. Adopting the basis of our set of Higgs legacy papers~\cite{barca,legacy1,legacy2} it reads \begin{align} \ensuremath{\mathcal{L}}_\text{dim-6}^H = & \frac{f_{BB}}{\Lambda^2} \phi^{\dagger} \hat{B}_{\mu \nu} \hat{B}^{\mu \nu} \phi + \frac{f_{WW}}{\Lambda^2} \phi^{\dagger} \hat{W}_{\mu \nu} \hat{W}^{\mu \nu} \phi - \frac{\alpha_s }{8 \pi} \frac{f_{GG}}{\Lambda^2} \phi^\dagger \phi G_{\mu\nu}^a G^{a\mu\nu} + \frac{f_{WWW}}{\Lambda^2} \text{tr} \left( \hat{W}_{\mu \nu} \hat{W}^{\nu \rho} \hat{W}_\rho{}^\mu \right) \notag \\ & + \frac{f_B}{\Lambda^2} (D_{\mu} \phi)^{\dagger} \hat{B}^{\mu \nu} (D_{\nu} \phi) + \frac{f_W}{\Lambda^2} (D_{\mu} \phi)^{\dagger} \hat{W}^{\mu \nu} (D_{\nu} \phi) + \frac{f_{\phi,2}}{\Lambda^2} \frac{1}{2} \partial^\mu ( \phi^\dagger \phi ) \partial_\mu ( \phi^\dagger \phi ) \notag \\ & + \left(\frac{f_\tau m_\tau}{v \Lambda^2} (\phi^\dagger\phi)(\bar L_3 \phi e_{R,3}) + \frac{f_b m_b}{v \Lambda^2} (\phi^\dagger\phi)(\bar Q_3 \phi d_{R,3}) + \frac{f_t m_t}{v \Lambda^2} (\phi^\dagger\phi)(\bar Q_3 \tilde \phi u_{R,3})+\text{h.c.}\right) \; . \label{eq:ourleff} \end{align} The Higgs covariant derivative is $D_\mu\phi= \left(\partial_\mu+ i g' B_\mu/2 + i g \sigma_a W^a_\mu/2 \right)\phi $, and the field strengths are $\hat{B}_{\mu \nu} = i g' B_{\mu \nu}/2$ and $\hat{W}_{\mu\nu} = i g\sigma^a W^a_{\mu\nu}/2$ in terms of the Pauli matrices $\sigma^a$. The $SU(2)_L$ and $U(1)_Y$ gauge couplings are $g$ and $g^\prime$, respectively. While the minimum independent set consists of 59 baryon number conserving operators, barring flavor structure and Hermitian conjugation~\cite{Grzadkowski:2010es}, we follow the definition of the relevant operator basis describing Higgs coupling and triple gauge boson vertex (TGV) modifications at the LHC in Ref.~\cite{barca}. In our construction we assume a narrow, CP-even Higgs, focusing on the minimal, Yukawa-like, couplings to the heavy fermions. We use the equations of motion to rotate to a basis where there are no blind directions linked to electroweak precision data. That way, we neglect all operators contributing to electroweak precision observables at tree level in our LHC analysis. For the Standard Model fit~\cite{barca,legacy1,legacy2} we omit the operator $(\phi^\dagger \phi )^3$, which only contributes to the rather poorly measured triple Higgs coupling. In the appendix we argue why even in the presence of an additional, mixing scalar, this operator will not add any extra relevant features to the fit.\medskip Moving to the new scalar Lagrangian terms, we assume in the following that the additional singlet does not develop a VEV, or that the Lagrangian can be re-defined such that the VEV vanishes~\cite{Dawson:2016ugw}. The effective Lagrangian of such an additional singlet scalar can be divided into two pieces. Following Refs.~\cite{Gripaios:2016xuo,Kamenik:2016tuv,Bauer:2016lbe,Franceschini:2016gxv} we first write down a set of non-redundant, independent operators up to dimension five, \begin{align} \ensuremath{\mathcal{L}}_{\text{dim}\leq 5}^S =& \frac{1}{2}\partial_\mu S \, \partial^\mu S-a_1 S-\frac{M_S^2}{2}S^2-a_3 S^3-a_4 S^4 -\frac{f_5^S}{\Lambda} S^5\notag\\ &-\mu_S S\phi^\dagger\phi -\frac{\lambda_{SH}}{2}S^2\phi^\dagger\phi -\frac{f_1^S}{\Lambda}S ( \phi^\dagger\phi )^2 -\frac{f_3^S}{\Lambda} S^3 \phi^\dagger\phi \notag\\ &+\frac{\alpha_s}{4\pi}\frac{f_{GG}^S}{\Lambda}SG_{\mu\nu}^aG^{a\;\mu\nu} +\frac{\alpha}{4\pi c_w^2}\frac{f_{BB}^S}{\Lambda}SB_{\mu\nu}B^{\mu\nu} +\frac{\alpha}{4\pi s_w^2}\frac{f_{WW}^S}{\Lambda}SW_{\mu\nu}^aW^{a\;\mu\nu} \notag \\ &\left(-\frac{f_d^S}{\Lambda}S\bar{Q}_L\phi d_R-\frac{f_u^S}{\Lambda}S\bar{Q}_L\tilde\phi u_R -\frac{f_\ell^S}{\Lambda}S\bar{L}_L\phi \ell_R+\text{h.c.}\right)\;\;\;. \label{eq:lagd5} \end{align} To be fully consistent with the Standard Model Lagrangian we could then add all dimension--six operators including at least one power of the new singlet scalar. The corresponding set of additional operators can be written as~\cite{Gripaios:2016xuo} \begin{align} \ensuremath{\mathcal{L}}_{\text{dim-6}}^S =& \frac{f^{SS}_\phi}{\Lambda^2} \phi^\dagger\phi \partial_\mu S \partial^\mu S -\frac{f_6^S}{\Lambda^2} S^6 -\frac{f_4^S}{\Lambda^2} S^4 \phi^\dagger\phi -\frac{f_2^S}{\Lambda^2} S^2 (\phi^\dagger\phi )^2 \notag \\ &+\frac{f_{GG}^{SS}}{\Lambda^2}S^2 G_{\mu\nu}^aG^{a\;\mu\nu} +\frac{f_{BB}^{SS}}{\Lambda^2}S^2B_{\mu\nu}B^{\mu\nu} +\frac{f_{WW}^{SS}}{\Lambda^2}S^2W_{\mu\nu}^aW^{a\;\mu\nu} \notag \\ &\left(-\frac{f_d^{SS}}{\Lambda^2}S^2\bar{Q}_L\phi d_R-\frac{f_u^{SS}}{\Lambda^2}S^2\bar{Q}_L\tilde\phi u_R -\frac{f_\ell^{SS}}{\Lambda^2}S^2\bar{L}_L\phi \ell_R+\text{h.c.}\right)\;\;. \label{eq:lagd6} \end{align} Nevertheless, given the singlet nature of the new scalar and neglecting lepton number violation, all dimension--six operators including the singlet are quadratic in the field $S$. Consequently, their phenomenological effects will be contributions to the mass terms ($f_2^S/\Lambda^2$), re-definitions of the $S$ field to recover canonical kinetic terms ($f^{SS}_\phi/\Lambda^2$), and the contributions to several vertices including two or more heavy scalars. After scalar-Higgs mixing, the two operators $f^{SS}_\phi/\Lambda^2$ and $f^S_2/\Lambda^2$ will contribute to the $SHH$ interaction as well. However, all these phenomenology features are already taken into account in our analysis by the free parameters in the dimension--five Lagrangian. Therefore, we neglect for the time being the explicit features induced by Eq.\eqref{eq:lagd6}. We give more details on the effective Lagrangian and the Higgs portal mixing in the Appendix. \subsection{Analysis framework} \label{sec:sfitter} The set of analyses presented here are derived using the \textsc{SFitter} framework. \textsc{SFitter} allows us to study multi-dimensional parameter spaces in the Higgs sector~\cite{sfitter_higgs,legacy1}, the gauge sector~\cite{legacy2} and in new physics models like supersymmetry~\cite{sfitter_susy}. The fit procedure uses Markov chains to create an exclusive, multidimensional log-likelihood map, based on the available measurements and including all the relevant uncertainties and correlations. The construction of a profile likelihood with flat theory uncertainties leads to the \textsc{RFit} scheme~\cite{rfit}. The statistic uncertainties on the measurements, both for event rates and kinematic distributions, follow Poisson statistics, as do the background uncertainties. All systematic uncertainties are described by Gaussian distributions and can be correlated between the relevant channels. We show log-likelihood projections on two-dimensional planes after profiling over all other parameters. Here, red-yellow regions will illustrate points within $\Delta (-2\log \mathcal{L})=2.3$ of the best fit point log-likelihood ($1\sigma$ in the Gaussian approximation), green regions indicate $\Delta (-2\log {\cal L})=6.18$ ($2\sigma$ in the Gaussian limit), and black dots imply the $\Delta (-2\log \mathcal{L})=5.99$ exclusion limits (95\%~CL in the Gaussian case).\medskip \begin{table}[b!] \begin{tabular}{llc} \hline Channel & Dataset & Reference \\ \hline $ S \rightarrow \gamma \gamma$ & ATLAS 8 TeV & \cite{S750_1000} \\ $ S \rightarrow \gamma \gamma$ & ATLAS 13 TeV & \cite{S750_1000} \\ $ S \rightarrow \gamma \gamma$ & CMS 8 TeV & \cite{S750_2000} \\ $ S \rightarrow \gamma \gamma$ & CMS 13 TeV & \cite{S750_2100} \\ $ S \rightarrow W W$ & ATLAS 8 TeV & \cite{S750_1010} \\ $ S \rightarrow W W$ & ATLAS 13 TeV & \cite{S750_1110} \\ $ S \rightarrow Z Z$ & ATLAS 8 TeV & \cite{S750_1020} \\ $ S \rightarrow Z Z$ & ATLAS 13 TeV & \cite{S750_1120} \\ $ S \rightarrow Z Z$ & ATLAS 13 TeV & \cite{S750_1121} \\ $ S \rightarrow Z \gamma$ & ATLAS 13 TeV & \cite{S750_1130} \\ $ S \rightarrow Z \gamma$ & CMS 13 TeV & \cite{S750_2130} \\ $ S \rightarrow Z \gamma$ & ATLAS 8 TeV & \cite{S750_1030} \\ $ S \rightarrow t\bar{t}$ & ATLAS 8 TeV & \cite{S750_1040} \\ $ S \rightarrow j j$ & CMS 8 TeV & \cite{S750_2050} \\ $ S \rightarrow h h$ & ATLAS 13 TeV & \cite{S750_1160} \\ $ S \rightarrow h h$ & CMS 8 TeV & \cite{S750_2060} \\ $ S \rightarrow \tau \bar{\tau}$ & CMS 8 TeV & \cite{S750_2070} \\ \hline \end{tabular} \caption{Experimental data on the heavy resonance included in our fit.} \label{tab:750_data} \end{table} The implementation of experimental results in the \textsc{SFitter} framework is described in Ref.~\cite{legacy1} for the Higgs measurements and in Ref.~\cite{legacy2} for anomalous triple gauge boson coupling measurements. For the triple gauge boson vertex (TGV) analyses\footnote{Note that pair production of weak bosons at the LHC is a crucial ingredient to a Higgs fit based on an effective Lagrangian assuming a linear realization of electroweak symmetry breaking. Without taking these measurements into account the qualitative and quantitative outcome of the fit will be wrong~\cite{legacy2}.} the correlation of the theory uncertainties between the different bins of a given kinematic distribution is taken into account by flat profiled nuisance parameters~\cite{legacy2}, while for the different Higgs channels the theory uncertainties are considered uncorrelated without a sizable impact on the shown results~\cite{legacy1}. For the Higgs portal analysis we take into account the constraints on a possible new resonance based on the data listed in Tab.~\ref{tab:750_data}. For the new resonance we only implement inclusive measurements assuming a narrow width. \section{Higgs portal analysis} \label{sec:ana} In the following we will use the \textsc{SFitter} effective Lagrangian framework to analyze a new gluon-fusion produced resonance in combination with the electroweak gauge and Higgs sectors at the weak scale. In other words, we ask the question whether such a new particle could be part of an extended Higgs sector and what the allowed parameter space is. In Sec.~\ref{sec:ana_heavy} we only include the dimension--five operators given in Eq.\eqref{eq:lagd5}, restricting the analysis to the data in Tab.~\ref{tab:750_data}. In Sec.~\ref{sec:ana_combined} we combine this analysis with the Higgs-electroweak measurements and the \textsc{SFitter} results induced by the dimension--six Lagrangian in Eq.\eqref{eq:ourleff}. Finally we link the size of different operators to a common origin in Sec.~\ref{sec:ana_magic}. \subsection{Heavy scalar fit} \label{sec:ana_heavy} \begin{figure}[t] \includegraphics[width=0.32\textwidth]{Figures/run2_chi2_unprim1.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run1_chi2_unprim5.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run1_chi2_unprim2.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}}\\ \includegraphics[width=0.32\textwidth]{Figures/run1_chi2_unprim6.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run1_chi2_unprim8.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run1_chi2_unprim10.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}} \caption{Two-dimensional profile log-likelihoods for the analysis of the heavy scalar sector alone spanning $f^S_{WW}$, $f^S_{BB}$, $f^S_{GG}$, $\sin\alpha$, and $c_{SHH}$. The black points indicate $\Delta(-2\log L)=5.99$.} \label{fig:5param_750} \end{figure} As a first step we analyze only the measurements for the heavy scalar, as listed in Tab.~\ref{tab:750_data}. In Fig.~\ref{fig:5param_750} we use this data to determine the five parameters \begin{align} \{ \, f^S_{WW}, f^S_{BB}, f^S_{GG}, \sin\alpha, c_{SHH} \, \} \; . \end{align} In our parametrization $c_{SHH}$ accounts for the independent contributions to the $SHH$ vertex from the dimension--five Lagrangian terms beyond the terms generating the mixing, as discussed in the Appendix. The best fit point for this analysis has $-2\log L=8.9$, while the SM point leads to $-2\log L=28.2$, within a $3.1\sigma$ range for a 5-parameter study (in the Gaussian limit). In the upper left panel of Fig.~\ref{fig:5param_750} we can see that within the displayed range of parameters both $f^S_{WW}$ and $f^S_{BB}$ are strongly correlated, and they present a flat direction. The correlation reflects the fact that they are the only Wilson coefficients contributing to the di-photon decay of the new scalar at tree level. Because $f^S_{WW}$ is constrained through the decay $S \rightarrow WW$, the di-photon excess cannot be accommodated through this coupling only. Due to that we find that $|f^S_{BB}/\Lambda| > 2~{\ensuremath\rm TeV^{-1}}$ in the upper-center panel. This is caused by the fact that $f^S_{BB}$ does not contribute to the $SWW$ vertex, and in addition its contribution to the $SZZ$ vertex is suppressed by the weak mixing angle. This allows us to explain the observed excess without getting into conflict with the exclusion bounds, what makes $f^S_{WW}=0$ compatible with the best fit point, as shown in the upper-right panel. Moving on to the mixing angle, we find $\sin \alpha < 0.15$ at 95\%~CL and for the displayed ranges of $f^S_{BB}$ and $f^S_{WW}$ in the lower-left panel. This bound comes from the absence of a heavy scalar signal in $WW$ and $ZZ$, but also in di-jet, $t\bar{t}$, $\tau\bar{\tau}$, and $hh$ decay channels. It is linked to maximum assumed values for $f^S_{BB}$ and $f^S_{WW}$, because a larger mixing angle can be partially compensated by larger Wilson coefficients $f^S_{BB}$+$f^S_{WW}$. For large values the di-photon branching ratio of the heavy scalar can exceed 50\%, while the remaining decay channel modes are suppressed, allowing $\sin \alpha$ to increase without conflicting with data. If we allow for extreme values of $f^S_{BB}/\Lambda+f^S_{WW}/\Lambda \sim 250~{\ensuremath\rm TeV^{-1}}$, the upper bound on $\sin \alpha$ goes up to $0.3$. In the lower-center panel we again observe two distinct regions in $f^S_{GG}$. The vertical region with $f^S_{GG}/\Lambda < 1.5$ TeV$^{-1}$ is characterized by a large branching ratio for $S\rightarrow \gamma \gamma$, linked to large values of $f^S_{BB}+f^S_{WW}$. The horizontal region with $f^S_{GG}/\Lambda = 1.5~...~10$ TeV$^{-1}$ is characterized by a large production rate for the new scalar and a total decay width driven by $f^S_{GG}$. The upper limit on $f^S_{GG}$ is set by di-jet searches, and the mixing in this regime has to be small to respect the limits from other decay channels. Finally, in the lower-right panel we show the correlation between the mixing angle and $c_{SHH}$ from the limit on the decay $S\rightarrow HH$. Fixing $c_{SHH} = 0$ and generating the $SHH$ interaction through the mixing angle alone has no effect on any of the other correlations.\medskip We proceed with an analysis allowing the new scalar to couple to the two fermions for which there are direct searches available. The analysis now includes \begin{align} \{ \, f^S_{WW}, f^S_{BB}, f^S_{GG}, \sin\alpha, c_{SHH}, f^S_{t}, f^S_{\tau} \, \} \; . \end{align} A selection of results is shown in Fig.~\ref{fig:7param_750}. The fermionic Wilson coefficients $f_t^S$ and $f_\tau^S$ are constrained by $t\bar{t}$ and $\tau^+\tau^-$ resonance searches, as well as from an upper limit $\Gamma_S < 25$~GeV which we assume throughout our analysis and which sets hard limits on $f_t^S$ and $f_\tau^S$. The best fit point of this run is only mildly better than before, $-2\log L=8.3$. The limits on these two fermion couplings are stronger for smaller $f_{BB}^S$, as illustrated for $f_t^S$ in the upper-left panel, and $f_{\tau}^S$ in the upper-center one. The reason is that in those regions the partial decay width to photons becomes small, and the required di-photon branching ratio translates into small fermionic couplings. Conversely, larger fermionic Wilson coefficients now allow for best fit regions with large $f^S_{GG}$ and $f^S_{BB}$ at the same time, as shown in the upper-right panel of Fig.~\ref{fig:7param_750}. This is the main difference with respect to the reduced analysis shown in Fig.~\ref{fig:5param_750}. The rest of correlations remain qualitatively unchanged, as can be seen in the lower panels of Fig.~\ref{fig:7param_750}. In particular the upper 95\% CL limit on the mixing angle is still $\sin \alpha < 0.15$.\medskip \begin{figure}[t] \includegraphics[width=0.32\textwidth]{Figures/run3_2.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run3_3.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run3_1.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}}\\ \includegraphics[width=0.32\textwidth]{Figures/run3_4.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run3_5.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run3_0.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}}\\ \caption{Two-dimensional profile log-likelihoods for the analysis of the heavy scalar sector alone. In contrast to Fig.~\ref{fig:5param_750} we now include fermion couplings in our set of Wilson coefficients $f^S_{WW}$, $f^S_{BB}$, $f^S_{GG}$, $\sin\alpha$, $c_{SHH}$, $f^S_{t}$ and $f^S_{\tau}$. The black points indicate $\Delta(-2\log L)=5.99$.} \label{fig:7param_750} \end{figure} In passing we note that all the results shown so far have been derived assuming a CP-even new scalar. Nevertheless, for the analysis up to this point the results remain unchanged when instead we assume a heavy CP-odd scalar. \subsection{Combined Higgs portal fit} \label{sec:ana_combined} \begin{figure}[t] \includegraphics[width=0.32\textwidth]{Figures/run7_0.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run7_1.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run7_2.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}} \caption{Two-dimensional profile log-likelihoods for the combined Higgs, TGV, and heavy scalar sectors. The black points indicate $\Delta(-2\log L)=5.99$.} \label{fig:17param_750} \end{figure} Next, we discuss the results for the general scenario, where we constrain the 17 parameters \begin{align} \{ \, f^S_{WW}, f^S_{BB}, f^S_{GG}, \sin\alpha, f^S_t, f^S_b, f^S_\tau, f_{WW}, f_{BB}, f_{GG}, f_{W}, f_B, f_{\phi,2}, f_{WWW}, f_t, f_b, f_\tau \, \} \end{align} from the combined measurements in the electroweak-Higgs, and the heavy scalar sector. We have fixed $c_{SHH}=0$ given its minor impact on the fit results. In this case the best fit value has a likelihood of $-2\log L=242.0$, for an analysis containing $252$ measurements, while the Standard Model point leads to $-2\log L=273.9$. In Fig.~\ref{fig:17param_750} we show a reduced selection of correlations between Wilson coefficients. When adding the heavy scalar to the combined Higgs and gauge boson analysis, the potentially largest change in the results appears for $f_W$ and $f_B$. The twofold reason is illustrated in detail in the Appendix. First, focusing on the electroweak-Higgs phenomenology, while the contribution of $f_W$ and $f_B$ to the Higgs vertices is now weighted by the cosine of the mixing angle, their contribution to the triple gauge boson vertex is not. This generates a different pattern of Higgs-TGV correlations once we add the new scalar. Second, the mixing of the Higgs boson with the heavy scalar allows $f_W$ and $f_B$ to generate genuinely new Lorentz structure contributions to the $SWW$, $SZZ$ and $SZ\gamma$ vertices, on top of the contributions from the rest of dimension--five and dimension--six operators. The first effect turns out to be negligible, and given the small allowed size for the mixing angle, the electroweak-Higgs measurements are not precise enough to be sensitive to the scalar mixing contributions. Conversely, the second effect is more important. The mild preference for non-zero $f_W$ values from the electroweak-Higgs measurements~\cite{legacy2} causes the best fit regions to generate the new contribution to the decays $S \to WW, ZZ, Z\gamma$. These channels can be then better fit suppressing them further with a smaller mixing angle. The addition of the dimension--six operator causes then the upper bound on the scalar mixing to be mildly reduced with respect to the results in the previous section: now $\sin \alpha < 0.10$ at 95\% CL. This can be observed in the left panel of Fig.~\ref{fig:17param_750}. Apart from this effect, the small mixing angle causes a lack of sizable correlations between both the new scalar sector and the electroweak-Higgs sector. Consequently, the results and two-dimensional planes involving dimension--five operators are very similar to the ones shown in Fig.~\ref{fig:7param_750}. The planes involving dimension--six operators remain unchanged with respect to the results shown in Ref.~\cite{legacy2}, something that we illustrate in the center and right panels of Fig.~\ref{fig:17param_750} for two of the dimension--six correlations. \subsection{A common origin of operators} \label{sec:ana_magic} When we split a common scalar potential for two mixing states into a set of dimension--five and dimension--six operators, the question becomes how different the higher-dimensional effects in the light and heavy scalar couplings can really be. In this section we assume that the set of heavy scalar couplings are directly tied to their Higgs-like counter parts, \begin{alignat}{9} \frac{f_{GG}}{\Lambda^2} &= -2 \frac{f_{GG}^S}{\Lambda} \left|\frac{f_{GG}^S}{\Lambda}\right| \qquad \qquad \qquad & \frac{f_{f}}{\Lambda^2} &=-\frac{v}{m_f}\frac{f_{f}^S}{\Lambda}\left|\frac{f_{f}^S}{\Lambda}\right| \notag\\ \frac{f_{BB}}{\Lambda^2} & = -\frac{1}{4\pi^2}\frac{f_{BB}^S}{\Lambda}\left|\frac{f_{BB}^S}{\Lambda}\right| & \frac{f_{WW}}{\Lambda^2} & = -\frac{1}{4\pi^2}\frac{f_{WW}^S}{\Lambda}\left|\frac{f_{WW}^S}{\Lambda}\right| \; , \label{eq:prior1} \end{alignat} for $f = b,t,\tau$. The relative signs and pre-factors ensure that the underlying new physics scales are consistent, as defined in Eq.\eqref{eq:lagd5}. For the fermion case, this is motivated by the need to have a minimal flavor violating structure in both dimension--five and dimension--six operators to avoid large flavor changing neutral currents~\cite{Goertz:2014qia}. In a Bayesian language this approach would correspond to a Dirichlet prior, for example employed in the dark matter fit of Ref.~\cite{tim}, with an exponent parameter $\alpha\gg1$.\medskip After imposing the relations in Eq.\eqref{eq:prior1}, we proceed to perform the combined Higgs, triple gauge boson vertex and heavy scalar analysis spanning the 11 free parameters \begin{align} \{ \, f^S_{WW}, f^S_{BB}, f^S_{GG}, f^S_t, f^S_b, f^S_\tau, \sin\alpha, f_{W}, f_B, f_{\phi,2}, f_{WWW} \, \} \; . \end{align} We have again fixed $c_{SHH}=0$, while $f_{WW}$, $f_{BB}$, $f_{GG}$, $f_t$, $f_b$ and $f_\tau$ are set from Eq.\eqref{eq:prior1}. Interestingly, the best fit point is $-2\log L=242.6$, {i.\,e.}\ within the analysis precision very close to the best fit point of the previous general scenario. This illustrates one of the most important conclusions: when dimension--five and dimension--six operators of a similar type are imposed to be related, there are still regions in the new physics parameter space which can accommodate the di-photon anomaly while respecting the constraints from the electroweak-Higgs measurements.\medskip \begin{figure}[t] \includegraphics[width=0.32\textwidth]{Figures/run14_3.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run14_5.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run14_0.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}}\\ \includegraphics[width=0.32\textwidth]{Figures/run14_4.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run14_1.pdf} \hspace{-0.3cm} \includegraphics[width=0.32\textwidth]{Figures/run14_2.pdf} \hspace{-0.2cm}\raisebox{20pt}{\includegraphics[width=0.047\textwidth]{Figures/barcode.pdf}} \caption{Two-dimensional profile log-likelihoods for the combined Higgs, TGV, and heavy scalar fit, but assuming a common origin of operators as defined in Eq.\eqref{eq:prior1}. The black points indicate $\Delta(-2\log L)=5.99$.} \label{fig:11param_750} \end{figure} In Fig.~\ref{fig:11param_750} we again show a selection of two-dimensional correlations. In the upper-left panel we start with tight constraints on $f_{BB}^S$ and also on $f_{GG}^S$. Now $f_{BB}^S/\Lambda$ no longer presents an unconstrained direction, as the reduced allowed region for values around $-10~{\ensuremath\rm TeV^{-1}}$ is limited from the constraint that $f_{BB}$ and hence $f_{BB}^S$ is constrained by the Higgs measurements. Because of the minus signs in Eq.\eqref{eq:prior1} the region of allowed values for both $f_{BB}^S$ and $f_{WW}^S$ corresponds to the solution that flips the sign of the $H\gamma\gamma$ vertex while respecting its measured size~\cite{legacy2}, as seen in the upper-center panel. In the case of $f_{GG}^S$ and $f_{GG}$ the several best fit regions are due to the measurement of a SM--like Higgs boson in gluon fusion production, the interference between $f_{GG}$ and $f_t$~\cite{legacy1}, and the heavy scalar anomaly that excludes $f_{GG}$ null values. As seen in the upper-right panel, the stronger constraints on $f_{BB}^S$ are directly translated into a stringent 95$\%$ CL bound on the mixing angle, $\sin\alpha<0.02$. In the lower-left panel we show the impact of Eq.\eqref{eq:prior1} on $f_t^S$. The fact that in this analysis $f_{BB}^S$ is more constrained than in the general scenario implies that $f_t^S$ is constrained to order one values, as expected from the $f_{BB}^S$ vs $f_t^S$ correlation in Fig.~\ref{fig:7param_750}. The solution for $f_t$ that flips the sign of the Higgs-Yukawa present in Ref.~\cite{legacy1} is excluded through $f_t^S$. This reduces the number of allowed regions for $f_{GG}$, as compared to the electroweak-Higgs fit~\cite{legacy1}. In the case of $f_\tau^S$ and $f_b^S$, the allowed regions are limited by the $H\tau^+\tau^-$ and $Hb\bar{b}$ measurements. The $v/m_f$ factors in Eq.\eqref{eq:prior1} lead to reduced allowed ranges for $f_\tau^S$ in comparison to the previous general scenario. We illustrate in the lower-center panel the allowed region for two of the dimension--six operators not involved in the simplifications of Eq.\eqref{eq:prior1}, $f_W$ and $f_B$. They remain unaltered with respect to the general analysis or the electroweak-Higgs results~\cite{legacy2}. Conversely, in the lower-right panel of Fig.~\ref{fig:11param_750} we illustrate the two parameter regions $f_{BB}$ vs. $f_{WW}$. There we see how the SM solution observed in the electroweak-Higgs analysis is now disfavored with respect to positive values for the Wilson coefficients.\medskip In this section we have illustrated the results of a constrained scenario imposing hard relations between the heavy scalar and Higgs operators in Eq.\eqref{eq:prior1}. Realistically, we would expect such relations to not be as strict. We therefore checked that relaxing Eq.\eqref{eq:prior1} and allowing for order-one variations does not qualitatively change our conclusions. Numerically, the bound on the mixing angle $\sin \alpha$ becomes weaker once the relation $f_{BB}\propto f_{BB}^S |f_{BB}^S|$ is relaxed. \section{Epitaph} \label{sec:summary} We have developed the framework to perform a combined analysis of the electroweak-Higgs sector extended with a new scalar to test Higgs portal scenarios. The theoretical description we have studied is that of a linear effective Lagrangian extended with the addition of a singlet scalar. The key question we face is the test of the portal structure hypothesis for an extended scalar sector. With that purpose we include a large set of Higgs event rates and kinematic distributions, combined with the recently implemented LHC triple gauge boson vertex distributions~\cite{legacy2}. As a test of a Higgs portal scenario we study the possibility that a di-photon anomaly recently observed at the LHC~\cite{750ex,S750_1000,S750_2100} could be part of an extended Higgs sector. For that we include a selection of relevant experimental searches for heavy resonances as listed in Tab.~\ref{tab:750_data}.\bigskip Analyzing first the new scalar sector only, we recover the result that a non-zero value for a reduced set of singlet scalar effective operators ($f_{GG}^S$, $f_{BB}^S+f_{WW}^S$) fits the observed anomaly in the di-photon channel, without conflicting with the lack of other positive observations, see Fig.~\ref{fig:5param_750}. The mixing angle of the new singlet state with the Higgs boson can be sizable, the upper bound we find in the analysis is $\sin(\alpha)<0.15$ at the 95\% CL. The addition of fermionic dimension--five operators increases the allowed parameter space regions for the bosonic operators. However it has no impact on the maximum allowed mixing angle value, see Fig.~\ref{fig:7param_750}. We then extend the analysis combining the new scalar sector with the electroweak-Higgs sector, using the Lagrangian description based on the dimension--six operators in Eq.\eqref{eq:ourleff}. In this extended scenario the upper bound on the mixing angle is further reduced in order to suppress the new dimension-six contributions to the heavy scalar non-observed decays. The upper bound is now $\sin(\alpha)<0.1$ at 95\% CL, with a size still compatible with Higgs portal hypothesis. Beyond this change, the maximum allowed mixing angle reduces the correlations between the Higgs-electroweak phenomenology and the hypothetical heavy scalar interactions. This leads to results that in the most general scenario are very similar to the ones of the individual Higgs-electroweak~\cite{legacy2}, and heavy scalar analysis, respectively. Motivated by a scalar portal scenario we define and test a hypothesis for a unique origin of the dimension--five and dimension--six operators studied in the analysis. Imposing Eq.\eqref{eq:prior1} we find new physics regions of parameters that fit the di-photon anomaly while being consistent with the lack of deviations measured on the electroweak-Higgs measurements. The upper bound on the mixing angle is reduced in this case to $\sin(\alpha)<0.02$, due to the strong constraints on the operators modifying $h\rightarrow \gamma\gamma$. \subsubsection*{Acknowledgments} MB, AB, JG-F and TP would like to thank the MITP for the hospitality while this paper was finalized. They also acknowledge support from the German Research Foundation (DFG) through the Forschergruppe `New Physics at the Large Hadron Collider' (FOR 2239), AB also through the Heidelberg Graduate School for Fundamental Physics and the Graduiertenkolleg `Particle physics beyond the Standard Model' (GRK 1940).
b48121aa2001194028a484c9382b4613d322e19a
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction}\quad Determining the properties and phases of cold and dense nuclear matter, as found for example in the interior of neutron stars, is a difficult task, because neither lattice methods nor perturbation theory can be applied in this regime of QCD. Holographic methods allow the study of strongly-coupled QCD-like theories as a function of temperature and baryon chemical potential, to hopefully provide some useful qualitative information about the rich variety of phases and transitions that might be expected in QCD. Pioneering studies \cite{Kim:2006gp,Bergman:2007wp,Rozali:2007rx} on cold and dense nuclear matter in holographic QCD have used the Sakai-Sugimoto model \cite{Sakai:2004cn,Sakai:2005yt}, as this is the preeminent example of a string theory realization of a holographic QCD-like theory. The two flavour version has a low energy description as a five-dimensional $U(2)$ gauge theory with a Chern-Simons term, in which baryons correspond to topological solitons in the bulk. As baryon number is identified with the instanton number of the gauge field, the study of baryons requires the construction of Yang-Mills-Chern-Simons solitons in curved space with a prescribed instanton number. It is a significant challenge to calculate the required solitons, even numerically, and to date only the single baryon solution has been computed \cite{Bolognesi:2013nja}. The computation in the baryon number one sector is facilitated by the application of continuous symmetries that are not present for higher baryon numbers, or at finite baryon number density. As a consequence, there are currently no numerical solutions available at finite density, either in the full Sakai-Sugimoto model or its five-dimensional version. This absence of solutions has motivated the use of a number of approximate methods \cite{Rozali:2007rx,Kim:2007vd,Kaplunovsky:2012gb,Li:2015uea} to study the Sakai-Sugimoto model at finite density. Often this involves the use of point-like instantons, but below we discuss some alternative approaches that share some common features with the effective kink model to be introduced in the present paper. An obvious way to simplify the problem at hand is to assume spatial homogeneity in the non-holographic spatial directions, so that the construction of the required finite density static solitons in the five-dimensional theory reduces to an effective one-dimensional problem. However, it is easy to show \cite{Rozali:2007rx} that there are no smooth spatially homogeneous gauge potentials that yield a non-zero baryon number density and have finite energy per baryon. To circumvent this problem, the approach in \cite{Rozali:2007rx} imposes homogeneity through the use of singular configurations. This reduces the problem to the determination of a profile function that specifies the singluar gauge potential as a function of the holographic coordinate. An alternative way to obtain spatial homogeneity is to replace the three-dimensional Euclidean spatial slice by a three-sphere \cite{Kim:2007vd}. As there are smooth finite energy spatially homogeneous gauge potentials on the three-sphere, this yields a finite baryon number density that can be varied by changing the radius of the three-sphere. However, in the studies in \cite{Kim:2007vd} the dependence on the holographic direction is then assumed to be a fixed self-dual form, so an effective one-dimensional kink model is not obtained and the possibility of different types of behaviour in the holographic direction is excluded. The approach taken in the present paper may be viewed as a hybrid of the above two methods. We apply a homogeneous approximation at the level of the field strength, rather than at the level of the gauge potential, in a way that produces an expression for the field strength that is very similar to that obtained on the three-sphere. However, we retain an arbitrary dependence on the holographic coordinate that leads to the derivation of an effective kink theory in the holographic direction. As the baryon chemical potential is increased this kink theory yields a first order phase transition, just below the baryon mass, to lightly bound nuclear matter, as found in QCD. We explain why our homogeneous approximation should be viewed as a smeared crystal that is expected to provide a lower bound on the energy, as it is an unattainable idealization that distributes the energy perfectly equally in space, which a true crystal is unable to match. We provide a justification for this expectation based on the simpler case of the Skyrme crystal. Homogeneous approximations are often employed to study field theory at very high density and usually overestimate the true energy because of the existence of symmetry breaking negative modes. It is important to note that our smearing approximation is in contrast to this standard application of homogeneity, as smearing lowers the energy to provide an underestimate that is expected to give valid predictions even for the small to moderate densities at which bound nuclear matter is formed. As baryon number density is increased it is natural to expect that the bulk soliton explores more of the holographic direction and this has been interpreted \cite{Kaplunovsky:2012gb} as a holographic realization of the quarkyonic phase \cite{McLerran:2007qj}, where there is a quark Fermi sea with a baryonic Fermi surface. It has also been proposed \cite{deBoer:2012ij} that this provides a mechanism for approximate chiral symmetry restoration. However, the details of the way in which the bulk soliton expands into the holographic direction are unknown and several different possibilities have been suggested. One proposal is a dyonic salt phase \cite{Rho:2009ym}, in which the bulk soliton splits into constituents at high density. This suggestion is motivated by a point particle approximation and the analogy with calorons, which are flat space self-dual instantons that can split into monopole constituents if a periodic direction is smaller than the size of an instanton. We find that our effective kink model provides an explicit realization of the dyonic salt phase, as the kink splits into a pair of half-kink constituents as the density is increased. Furthermore, there is a simple explanation for this splitting into constituents in terms of the novel potential that appears in the effective kink model and its explicit dependence on the holographic coordinate. A second proposal, based on the use of approximations involving flat space calorons and dilute instantons, is a baryonic popcorn phase \cite{Kaplunovsky:2012gb,Kaplunovsky:2015zsa}, in which a sequence of transitions takes place where the soliton crystal develops additional layers in the holographic direction with increasing density. The kink model is also able to capture the phenomenon of baryonic popcorn and displays a first order phase transition that generates an additional soliton layer in the holographic direction that is energetically preferred over the dyonic salt phase. The fact that the phase transition of the first pop into baryonic popcorn occurs before the dyonic salt phase agrees with the results of full numerical simulations of low-dimensional toy models \cite{Bolognesi:2013jba,Elliot-Ripley:2015cma}, based on sigma model instantons rather than Yang-Mills intantons. However, we find that the kink model predicts only one pop, rather than the sequence of pops suggested by previous point particle and flat space approximations \cite{Kaplunovsky:2012gb,Kaplunovsky:2015zsa}. Once again, the form of the potential that appears in the effective kink model provides a simple explanation of the absence of additional pops once the first pop has taken place. In the kink model, the two layers produced by the single pop form the surface of a soliton bag that increases in size as the baryon chemical potential is increased. The interior of the bag is filled with abelian electric potential and the instanton charge density is localized on the surface of the bag. Our soliton bag is therefore very similar to the magnetic bag \cite{Bolognesi:2005rk} that approximates a large number of coincident non-abelian BPS monopoles. A magnetic bag also has the topological charge density localized on its surface, with the interior and the exterior of the bag corresponding to different values of the modulus of the Higgs field. In our soliton bag the abelian electric potential plays the role of the modulus of the Higgs field, taking different values inside and outside the bag. Note that our soliton bag is qualitatively different from the previously proposed instanton bag \cite{Bolognesi:2014dja}, in which monopole walls are embedded into the Sakai-Sugimoto model to produce a bag where the interior is filled with instanton charge density, rather than this density being localized on the surface of the bag. \section{An effective holographic kink theory}\quad In our study of the Sakai-Sugimoto model we choose to work with the Yang-Mills theory that results from the non-abelian Dirac-Born-Infeld action at leading order in $\alpha'$. This simplification has the advantage that we do not need to choose a particular proposal for the non-abelian form of the Dirac-Born-Infeld action. The five-dimensional $U(2)$ Yang-Mills gauge theory version of the Sakai-Sugimoto model involves a spacetime with a warped metric of the form \begin{equation} \label{metric} ds^2 = H(z) \,dx_{\mu} dx^{\mu} + \frac{1}{H(z)}dz^2, \end{equation} where $x_\mu,$ with $\mu=0,1,2,3,$ are the coordinates of four-dimensional Minkowski spacetime and $z$ is the spatial coordinate in the additional holographic direction. The warp factor in this expression is $H(z)=(1+z^2)^{2/3}$ and the signature of the metric is $(-,+,+,+,+)$. In this section we follow the notation of \cite{Bolognesi:2013nja}, in which the hermitian $U(2)$ gauge potential is decomposed into a non-abelian $SU(2)$ component $A$ and an abelian $U(1)$ component $\hat A$, with lowercase latin indices (excluding $z$) running over the three non-holographic spatial directions $i=1,2,3$, whilst uppercase latin indices run over all four spatial directions, $I=1,2,3,z.$ We immediately restrict to the case of time independent fields, where the appropriate static ansatz is \begin{equation} {A}_0=0, \qquad A_I={A}_I(x_{J}), \qquad \hat{A}_0=\hat{A}_0(x_{J}), \qquad \hat{A}_{I} = 0. \end{equation} Then, in units of $N_c\lambda/(216\pi^3)$, where $N_c$ is the number of colours and $\lambda$ is the 't Hooft coupling, the static energy of the five-dimensional Sakai-Sugimoto model is given by \begin{equation} E=\int\bigg\{ -\frac{(\partial_i\hat A_0)^2}{2H^{1/2}}-\frac{H^{3/2}}{2}(\partial_z\hat A_0)^2 +\frac{\mbox{Tr}(F_{ij}^2)}{2H^{1/2}}+H^{3/2}\mbox{Tr}(F_{iz}^2) +\frac{4}{\Lambda}\hat A_0\epsilon_{ijk}\mbox{Tr}(F_{iz}F_{jk})\bigg\}\,d^3x\,dz, \label{energy} \end{equation} where $\Lambda=8\lambda/(27\pi)$ is the rescaled 't Hooft coupling. The final term in the above expression arises from the Chern-Simons term. Baryon number is identified with the $SU(2)$ instanton number of the soliton \begin{equation} N=\frac{1}{8\pi^2}\int\epsilon_{ijk}\mbox{Tr}(F_{iz}F_{jk})\,d^3x\,dz, \label{charge} \end{equation} and this provides the energy bound $E\ge 8\pi^2 |N|.$ From the final term in (\ref{energy}) we see that the instanton charge density sources the abelian electric field. In the gauge $A_z=0$ the holonomy that carries the pion degrees of freedom \begin{equation} U({\bf x})={\cal P}\mbox{exp}\bigg(i\int_{-\infty}^\infty A_z({\bf x},z)\,dz\bigg), \end{equation} appears in the boundary condition for $A_i$, namely, \begin{equation} A_i({\bf x},z)\to R_i({\bf x}) \quad \mbox{ as }\quad z\to\infty, \qquad \mbox{ where } \quad R_i({\bf x})=i(\partial_iU)U^{-1}. \label{bc} \end{equation} The holonomy $U({\bf x}):{\mathbb R}^3\to SU(2)$ carries the topology of the field configuration due to the equality $N=\mbox{deg}\,U\in\mathbb{Z}=\pi_3(SU(2))$, where the compactification of $\mathbb{R}^3$ is a consequence of the fact that $U({\bf x})$ tends to the identity matrix as $|{\bf x}|\to \infty.$ Note that the pure gauge currents, $R_i({\bf x})$, satisfy the zero curvature condition \begin{equation} \partial_iR_j-\partial_jR_i+i[R_i,R_j]=0. \label{zerocurvature} \end{equation} The first step in deriving our effective kink model is to work in the gauge $A_z=0$ and apply the separable approximation \begin{equation} A_i({\bf x},z)=R_i({\bf x})\psi(z), \label{sep}\end{equation} where the real function $\psi(z)$ satisfies the boundary conditions $\psi(-\infty)=0$ and $\psi(\infty)=1$, in order to reproduce the correct boundary condition (\ref{bc}). Applying the separable approximation (\ref{sep}), and making use of the zero curvature relation, yields the gauge field strength \begin{equation} F_{ij}=-i[R_i,R_j]\psi(1-\psi), \qquad F_{iz}=-R_i\psi'. \label{homog} \end{equation} Our second step is to make a homogeneous approximation by assuming that the gauge field strength is independent of ${\bf x}$. Explicitly, we set $R_i=-\beta\sigma_i/2$, where $\sigma_i$ denote the Pauli matrices and $\beta$ is a real constant that we take to be positive, as the sign of $\beta$ will turn out to be equal to the sign of the baryon number density. It is crucial that we apply the homogeneous approximation at the level of the field strength and not at the level of the gauge potential, since there are no homogeneous gauge potentials that yield the formulae (\ref{homog}). This is because the zero curvature condition (\ref{zerocurvature}) is obviously violated by restricting to constant non-commuting currents. By imposing homogeneity of the field strength, rather than the gauge potential, we have been able to incorporate the zero curvature relation before restricting to homogeneous fields that violate it. If a homogeneous approximation is applied directly at the level of the gauge potential then the zero curvature relation (\ref{zerocurvature}) implies that the field is topologically trivial. The only way to reintroduce instanton charge is to incorporate it via a discontinuous gauge potential \cite{Rozali:2007rx}, with the discontinuity being the source of the instanton charge. Essentially, the winding of the field is then moved from infinity onto a discontinuity in the bulk. The advantage of our continuous approach is that we avoid the need to deal directly with the non-abelian gauge potential, by restricting attention to the continuous physical fields. In summary, our homogeneous approximation is \begin{equation} F_{ij}=\frac{1}{2}\beta^2\psi(1-\psi)\epsilon_{ijk}\sigma_k, \qquad F_{iz}=\frac{1}{2}\beta \psi'\sigma_i, \qquad \hat A_0=\omega(z). \label{ansatz} \end{equation} These expressions are very similar to those that appear if the spatial slice $\mathbb{R}^3$ is replaced by $S^3$ with a finite radius, as studied in \cite{Kim:2007vd}. In the case of the three-sphere, smooth spatially homogeneous gauge potentials exist and directly generate relations analogous to (\ref{ansatz}), with $\beta$ related to the inverse of the radius of the three-sphere. However, in the work in \cite{Kim:2007vd} the analogue of the kink function $\psi(z)$ is taken to be a fixed self-dual form, so an effective kink model is not obtained. Taking a fixed form for the kink function prevents a study of the way in which the bulk soliton explores the holographic direction with increasing density and the associated different phases. This is the main purpose of the present paper. It might be interesting to repeat the analysis presented here for the case of the three-sphere with an unfrozen kink field, where genuine homogeneous gauge potentials underpin the approximation. To begin with, we work in the canonical ensemble with fixed baryon number density and no explicit baryon chemical potential. The boundary condition on the real function $\omega(z)$ is therefore $\omega(\pm\infty)=0,$ because in holographic QCD the boundary value of $\hat A_0$ is proportional to the baryon chemical potential. At finite density, the true spatial distribution of the fields in the non-holographic directions is expected to form a soliton crystal. Our homogeneous approximation may be viewed as a smeared version of the crystal, and we expect that this approximation provides a lower bound on the true crystal energy, as homogeneity is an unattainable idealization. A justification for this expectation is provided in the appendix, where we consider the related, though simpler, case of the Skyrme crystal. Substituting the expressions (\ref{ansatz}) into (\ref{energy}) and (\ref{charge}) gives the energy per unit 3-volume \begin{equation} {\cal E}=\frac{1}{2}\int_{-\infty}^\infty\bigg\{ 3\beta^2H^{3/2}\psi'^2+\frac{3}{H^{1/2}}\beta^4\psi^2(1-\psi)^2-H^{3/2}\omega'^2 +\frac{24}{\Lambda}\beta^3\omega\psi(1-\psi)\psi'\bigg\}\,dz \label{henergy} \end{equation} and the baryon number density (ie. instanton number per unit 3-volume) \begin{equation} \rho=\int_{-\infty}^\infty\frac{3\beta^3}{8\pi^2}\psi(1-\psi)\psi'\,dz =\frac{\beta^3}{16\pi^2}. \label{density} \end{equation} The field equation for $\omega$ that follows from the variation of (\ref{henergy}) is \begin{equation} \big(H^{3/2}\omega'\big)'=\frac{12\beta^3}{\Lambda}\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2\bigg)'. \label{omegaeqn} \end{equation} \begin{figure}[ht] \begin{center} \includegraphics[width=9cm]{evsb.png} \caption{The ratio ${\cal E}/(8\pi^2\rho)$ of the energy to the energy bound as a function of baryon number density $\rho$. The blue curve is a single layer and the red curve is a double layer.} \label{fig-energy} \end{center} \end{figure} As energy is minimized by a bulk configuration that is localized around $z=0$, we impose the symmetry that both the instanton charge density and $\omega$ are even functions of $z$, that is, $\omega(-z)=\omega(z)$ and $\psi(-z)=1-\psi(z)$. We can therefore restrict our discussion to the region $z\ge 0$, together with the boundary conditions $\omega'(0)=0$ and $\psi(0)=\frac{1}{2}$. Integrating (\ref{omegaeqn}) once and applying these boundary conditions at $z=0$ yields \begin{equation} H^{3/2}\omega'=\frac{12\beta^3}{\Lambda}\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2+\frac{1}{12}\bigg). \label{wpp} \end{equation} Taking the limit of this equation as $z\to\infty$ gives \begin{equation} \lim_{z\to\infty}z^2\omega'=-\frac{\beta^3}{\Lambda}=-\frac{16\pi^2\rho}{\Lambda}.\end{equation} We therefore obtain the usual holographic result, relating the coefficient of the asymptotic behaviour to the baryon number density, \begin{equation} \omega=\frac{16\pi^2\rho}{\Lambda z}+o\bigg(\frac{1}{z}\bigg). \label{asymptotic} \end{equation} We can rewrite the final term in (\ref{henergy}) by applying an integration by parts to obtain the identity \begin{equation} \int_{-\infty}^\infty\omega\psi(1-\psi)\psi'\,dz =\int_{-\infty}^\infty\omega'\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2+\frac{1}{12}\bigg)\,dz. \end{equation} Using this result, together with (\ref{wpp}), we obtain \begin{equation} {\cal E}=\int_0^\infty\bigg\{ 3\beta^2H^{3/2}\psi'^2+\frac{3}{H^{1/2}}\beta^4\psi^2(1-\psi)^2 +\frac{144\beta^6}{\Lambda^2H^{3/2}}\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2+\frac{1}{12}\bigg)^2 \bigg\}\,dz, \label{psionly} \end{equation} where all reference to $\omega$ has been eliminated. This is the energy of our effective kink model. Note that imposing the flat space self-duality equation $F_{IJ}=\frac{1}{2}\epsilon_{IJKL}F_{KL}$ on the homogeneous fields (\ref{ansatz}) yields the first order equation \begin{equation}\psi'=\beta\psi(1-\psi),\end{equation} with kink solution \begin{equation} \psi=\frac{1}{1+e^{-\beta z}}. \label{selfdual} \end{equation} This is the self-dual form of the kink function that is assumed to hold for all densities in \cite{Kim:2007vd}. In the flat space limit $H=1$ with no Chern-Simons term ($\Lambda=\infty$), the energy (\ref{psionly}) of the self-dual solution (\ref{selfdual}) is ${\cal E}=\beta^3/2=8\pi^2\rho$, so the BPS behaviour of flat space instantons is recovered within our homogeneous approximation. \begin{figure}[ht] \begin{center} \includegraphics[width=9cm]{freeenergy.png} \caption{The free energy $\Omega$ as a function of the chemical potential $\mu$ for a single layer (blue curve) and a double layer (red curve).} \label{fig-freeenergy} \end{center} \end{figure} \begin{figure}[ht] \begin{center} \includegraphics[width=9cm]{muvsrho.png} \caption{The baryon number density $\rho$ as a function of the chemical potential $\mu.$ The blue curve corresponds to a single layer and shows a first order phase transition just below the baryon mass. The red curve corresponds to a double layer and shows the first order phase transition of baryonic popcorn.} \label{fig-chempot} \end{center} \end{figure} The field equation that follows from the variation of (\ref{psionly}) is \begin{equation} \psi''+\frac{3H'}{2H}\psi' +\frac{\beta^2}{H^2}\psi(1-\psi)(2\psi-1) +\frac{48\beta^4}{\Lambda^2H^{3}}\psi(1-\psi)\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2+\frac{1}{12}\bigg)=0. \label{psieqn} \end{equation} We solve this equation numerically using a gradient flow algorithm and the change of variable $z=\tan u,$ to map the infinite range of $z$ onto a finite range of $u.$ All the numerical results presented in this paper are computed with the fiducial value $\Lambda=10.$ In Figure~\ref{fig-energy} the blue curve shows a plot of the ratio of the energy to the lower bound, that is ${\cal E}/(8\pi^2\rho)$, as a function of the baryon number density $\rho$. We see that there is a non-zero optimum density $\rho_\star= 0.027$, corresponding to the analogue of the nuclear matter density in QCD. As described below, this optimal density is associated with a critical value of the baryon chemical potential at which there is a first order phase transition to an equilibrium density of baryons. \begin{figure}[ht] \begin{center} \includegraphics[width=6.5cm]{1a.png} \includegraphics[width=6.5cm]{1aa.png} \includegraphics[width=6.5cm]{1b.png} \includegraphics[width=6.5cm]{1bb.png} \includegraphics[width=6.5cm]{1c.png} \includegraphics[width=6.5cm]{1cc.png} \caption{The single layer profile function $\psi(z)$ (top row), the instanton charge density per unit 3-volume (middle row) and the abelian electric potential $\omega(z)$ (bottom row) for baryon number densities $\rho=0.1$ (left column) and $\rho=2$ (right column). At high density the kink splits into a pair of half-kink constituents.} \label{fig-psi} \end{center} \end{figure} In the above we have worked in the canonical ensemble with fixed baryon number density, but an alternative is to work in the grand canonical ensemble with fixed baryon chemical potential $\mu$. Applying the standard holographic dictionary, in the grand canonical ensemble a baryon chemical potential corresponds to a non-zero boundary value for $\hat A_0$. With our chosen normalizations, a vectorial chemical potential $\mu$ corresponds to the boundary condition \begin{equation} \omega(\pm\infty)= -\frac{\mu \Lambda}{32\pi^2}. \label{chempot} \end{equation} This is because a shift in $\omega$ by this constant value returns us to the previous boundary condition $\omega(\pm\infty)=0$, but from (\ref{henergy}) we see that the energy then transforms into the grand potential (also known as the Landau free energy) $\Omega={\cal E}-\mu\rho.$ Minimizing $\Omega$ at fixed $\mu$ we obtain the function $\Omega(\mu)$ displayed as the blue curve in Figure~\ref{fig-freeenergy}. The associated relation between the chemical potential $\mu$ and the baryon number density $\rho$ is displayed as the blue curve in Figure~\ref{fig-chempot}. These curves show that a non-zero value for $\rho$ is obtained for $\mu\ge\mu_\star,$ where $\mu_\star=98$ (for $\Lambda=10$) is the critical value of the chemical potential at which the density is indeed equal to $\rho_\star$. For attraction between baryons, the critical value of the chemical potential $\mu_\star$ must be less than the baryon mass $M_B.$ In the units we are using, the self-dual single instanton approximation to the baryon yields the formula \cite{Bolognesi:2013nja} \begin{equation} M_B=2\pi^2\bigg(4+\frac{32}{\Lambda}\sqrt{\frac{2}{15}}\bigg), \label{baryonmass} \end{equation} where terms of order $1/\Lambda^2$ have been neglected. Substituting our fiducial value $\Lambda=10$ into (\ref{baryonmass}) gives $M_B=102,$ so indeed our numerically computed value $\mu_\star=98$ is less then $M_B$ and baryons form bound states. The percentage binding energy per nucleon is given by \begin{equation} \Delta = (M_B-\mu_\star)/M_B\times 100\%, \end{equation} yielding $\Delta=4\%$ for the chosen value $\Lambda=10.$ In the limit $\Lambda\to\infty$ the BPS result is recovered, $M_B\to 8\pi^2$ and $\mu_\star\to M_B$, giving zero binding energy. Although we expect holographic QCD to provide only qualitative information about QCD with three colours, the above results imply that there must be a value of $\Lambda$ greater than 10 at which a realistic nuclear binding energy $\Delta \sim 0.9\%$ is obtained, and indeed we compute that the appropriate value is $\Lambda\sim 18$. Returning to the representative value $\Lambda=10$, in the top row in Figure~\ref{fig-psi} we display the kink profile function $\psi(z)$ for the densities $\rho=0.1$ (left column) and $\rho=2$ (right column). In the middle row we plot the corresponding instanton number densities per unit 3-volume, and in the final row we display the abelian electric potential $\omega(z)$. We see that at high density the kink splits into two half-kink constituents. This splitting of the soliton layer into a pair of constituents is the predicted dyonic salt phenomenon, corresponding to the fact that a periodic instanton (a caloron) splits into monopole constituents at high density. Note that a half-kink cannot exist in isolation as a finite energy configuration because the profile function $\psi(z)$ of a half-kink interpolates between values that differ by $\frac{1}{2}$, but for finite energy this difference must be equal to $\pm1$ or $0$. This mirrors the caloron situation, where the caloron splits into monopole constituents that are not finite energy configurations in isolation but together combine to form a well-defined periodic instanton. The form of the energy (\ref{psionly}) of the effective kink model provides a simple explanation for the split into half-kink constituents, as follows. The kink model contains a derivative term plus two potential terms, which are novel in a kink theory due to the form of the explicit dependence on the holographic spatial coordinate. The first potential term has an explicit spatial dependence that grows like $z^{-2/3}$ for large $z$. Such a term does not decay fast enough to be integrable, so the associated multiplying factor must tend to zero as $z\to\pm\infty.$ This enforces the boundary conditions $\psi(\pm\infty)\in\{0,1\}.$ The second potential term has an explicit spatial dependence that grows like $z^{-2}$, which decays fast enough that finite energy considerations impose no conditions on the boundary values of $\psi(z)$ from this term. However, as this term has a coefficient of $\beta^6$ then at high density it is, at least locally, of more relevance than the first potential term, which has a coefficient of only $\beta^4$. The second potential term is minimized by the value $\psi=\frac{1}{2}$, as it vanishes at this value. Putting all this toegther we see that as the density increases the second potential term induces the kink field $\psi(z)$ to remain close to the value $\frac{1}{2}$ over an increasingly large range of $z$, although the finite energy requirement of the first potential term always forces the kink boundary conditions to ultimately be attained. This is the simple reason why the kink splits into a pair of half-kinks. Previous results on low-dimensional models \cite{Bolognesi:2013jba,Elliot-Ripley:2015cma} suggest that a baryonic popcorn transition \cite{Kaplunovsky:2012gb} appears before the formation of dyonic salt. In other words, at densities high enough to split the kink into a pair of half-kink constituents, we expect that there is a competing solution that describes a double layer and has lower energy than the single layer considered in this section. In the following section we examine this possibility by constructing a double layer configuration and calculating its energy as a function of baryon number density. \begin{figure}[ht] \begin{center} \includegraphics[width=6.5cm]{2a.png} \includegraphics[width=6.5cm]{2aa.png} \includegraphics[width=6.5cm]{2b.png} \includegraphics[width=6.5cm]{2bb.png} \includegraphics[width=6.5cm]{2c.png} \includegraphics[width=6.5cm]{2cc.png} \caption{The double layer profile function $\psi(z)$ (top row), the instanton charge density per unit 3-volume (middle row) and the abelian electric potential $\omega(z)$ (bottom row) for baryon number densities $\rho=2$ (left column) and $\rho=10$ (right column).} \label{fig-psi2} \end{center} \end{figure} \section{Baryonic popcorn and soliton bags}\quad The effective kink model, with energy given by (\ref{psionly}), has an anti-kink solution with boundary conditions $\psi(-\infty)=1$ and $\psi(\infty)=0.$ It is obtained from the kink solution by the transformation $\psi\mapsto 1-\psi$ and has the same energy per unit volume as the kink but has a negative baryon number density. However, if the ansatz (\ref{ansatz}) is modified by the replacement $F_{ij}\mapsto -F_{ij}$ then the anti-kink now has a positive baryon number density, and this provides an equivalent anti-kink formulation of the single layer described in the previous section in terms of a kink. We can construct a double layer configuration by gluing together two single layers in a continuous manner. Explicitly, we apply the approximation (\ref{ansatz}) for $z\ge 0$ with kink boundary conditions on the half-line, $\psi(0)=0$ and $\psi(\infty)=1.$ For $z\le 0$ we take the same approximation (\ref{ansatz}) after the replacement $F_{ij}\to -F_{ij}$ with anti-kink boundary conditions on this half-line, $\psi(-\infty)=1$ and $\psi(0)=0.$ This corresponds to taking both $\psi(z)$ and $\omega(z)$ to be even functions of $z$ and gives a continuous instanton charge density. The instanton charge density vanishes at the join at $z=0$, and although it is continuous at this point it is not smooth there. Note that the parity of $\omega$ is as before, in agreement with the requirement that we consider a vectorial baryon chemical potential. The baryon number density for this two layer configuration is \begin{equation} \rho=\int_{0}^\infty\frac{6\beta^3}{8\pi^2}\psi(1-\psi)\psi'\,dz =\frac{\beta^3}{8\pi^2}, \end{equation} which is twice the value (\ref{density}) for the single layer. Integrating the field equation (\ref{omegaeqn}) for $\omega$, and imposing the new boundary condition $\psi(0)=0$ and $\omega'(0)=0$, gives for $z\ge 0$ \begin{equation} H^{3/2}\omega'=\frac{12\beta^3}{\Lambda}\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2\bigg). \label{wpp2} \end{equation} Taking the limit of this equation as $z\to\infty$ provides the relation \begin{equation} \lim_{z\to\infty}z^2\omega'=-\frac{2\beta^3}{\Lambda}=-\frac{16\pi^2\rho}{\Lambda},\end{equation} again reproducing the correct asymptotic behaviour (\ref{asymptotic}). As in the single layer case, an integration by parts yields the following expression for the double layer energy, \begin{equation} {\cal E}=\int_0^\infty\bigg\{ 3\beta^2H^{3/2}\psi'^2+\frac{3}{H^{1/2}}\beta^4\psi^2(1-\psi)^2 +\frac{144\beta^6}{\Lambda^2H^{3/2}}\bigg(\frac{1}{3}\psi^3-\frac{1}{2}\psi^2\bigg)^2 \bigg\}\,dz, \label{psionly2} \end{equation} as a function of $\psi$ only. We obtain the double layer solution by numerically solving the static field equation for $\psi(z)$ that follows from the variation of this energy. The red curve in Figure~\ref{fig-energy} is a plot of the ratio of the energy to the lower bound, as a function of the baryon number density, for the double layer solution. From this figure we see that the double layer has a lower energy than the single layer beyond the critical density $\rho_2= 0.22.$ This is the critical density for a homogeneous baryonic popcorn transition to a double layer, and is well below the density at which dyonic salt appears. This result is therefore in agreement with the previous low-dimensional studies mentioned earlier, where a popcorn transition also appears before the dyonic salt phase. In the top row in Figure~\ref{fig-psi2} we display the double layer profile function $\psi(z)$ for the densities $\rho=2$ (left column) and $\rho=10$ (right column). In the middle row we plot the corresponding instanton number densities per unit 3-volume, and in the final row we display the abelian electric potential $\omega(z)$. Although there are some qualitative similarities between a double layer and a single layer that has split into two half-layer constituents, the crucial distinction is that a half-layer cannot exist in isolation as a finite energy configuration. There are, of course, significant quantitative differences, including the fact that the double layer configuration has a much lower energy at high densities. The free energy of the double layer as a function of the chemical potential is shown as the red curve in Figure~\ref{fig-freeenergy}. This reveals that within the grand canonical ensemble there is a critical value of the chemical potential, given by $\mu\ge\mu_2=185$ for $\Lambda=10$, above which the free energy of the double layer is less than that of a single layer. This results in the baryonic popcorn first order phase transition seen in Figure~\ref{fig-chempot} at $\mu=\mu_2$, where the configuration pops from the single layer (given by the blue curve) to the double layer (given by the red curve). Note that in Figure 3 we only plot the single layer and double layer portions of the curve for the range of chemical potentials at which each is the global minimum of the free energy. The curves extend beyond these segments to physically irrelevant regions in which they are no longer global minima. Given the results in the previous section, where a single layer splits into half-layer constituents at high density, one might naively expect a similar phenomenon to take place for a double layer, with each layer splitting into half-layer constituents at high density. However, the right column in Figure~\ref{fig-psi2} demonstrates that the double layer does not split, even at very high densities. This has a simple explanation, again obtained by examining the form of the novel potential terms in the effective energy (\ref{psionly2}). This time, we see that the final potential term in this effective energy is minimal only at $\psi=0$ (for $\psi\in[0,1]$). Therefore, unlike the case of a single kink, the final potential term does not induce the kink or anti-kink to stay close to any new value as the density is increased. Hence the double layer does not split, but rather the two layers simply increase their separation as the density increases. As the density increases and the two layers move further apart, it would be reasonable to surmise that further pops would occur, since there now appears to be available space around $z=0$ to generate new layers. However, this simple view ignores the fact that the region between the two layers is not empty but instead is full of abelian electric potential, as demonstrated by the plots in the bottom row of Figure~\ref{fig-psi2}. The two layers form the surface of a soliton bag, with the interior of the bag associated with an approximately constant non-zero value of the abelian electric potential. The abelian electric potential decays to zero outside the bag, with the surface of the bag being the transition region where both the instanton charge density and the electric field are localized. This is similar to the magnetic bag description of large charge non-abelian monopoles \cite{Bolognesi:2005rk}, where the surface of the bag separates regions of zero and non-zero values for the modulus of the Higgs field. As the system under consideration is periodic in three spatial directions (approximated by homogeneity) then the surface of the bag is not a single connected component, like the magnetic bag in three-dimensional Euclidean space, but instead consists of two disconnected components corresponding to the top and bottom of the bag. This is why the soliton bag requires two layers. The creation of more layers through further baryonic popcorn transitions would produce a kind of multi-layer bag. In the monopole context, the possibility of multi-layer magnetic bags has been investigated in \cite{Manton:2011vm} with the conclusion that these typically consist of only a single bag surrounded by layers of isolated unit charge monopoles. Any attempt to create a multi-bag configuration automatically rules out an interior bag that carries any significant fraction of the total magnetic charge. The soliton bag description therefore suggests the absence of additional baryonic popcorn transitions. Further support for this view is obtained by extending our anti-kink plus kink double layer approximation, denoted $\bar K K$, to additional layers by the inclusion of more anti-kinks/kinks. For example, a 4-layer $\bar K K\bar K K$ approximation has been studied where all contiguous anti-kinks and kinks are joined in the same continuous manner as in the double layer approximation. The location of the joins, together with the fraction of instanton charge carried by each layer, are allowed to vary and the resulting energy minimizing configurations computed. As expected from the similar monopole magnetic bag story, these computations yield only signficiant instanton charge density in the outer layer and the inner layer is irrelevant. A recent study \cite{Preis:2016fsp}, applying a different approximation, also reached the same conclusion that a double layer is the preferred configuration at high density and additional layers are not generated. Finally, it is important to note that our soliton bag is qualitatively different from the instanton bag proposed in \cite{Bolognesi:2014dja} as a description of the high density phase in the Sakai-Sugimoto model. The instanton bag is obtained via an initial compactification of one of the spatial directions to allow the embedding of a monopole wall. A pair of monopole walls are then patched together to form the surface of the instanton bag, with the result that the interior is filled with instanton charge density. This contrasts with our soliton bag, where the instanton charge density is localized on the surface of the bag. Our kink approximation, with an appropriate shape for the kink profile function, could produce a bag filled with instanton charge density, but such a shape does not appear when the profile function is obtained by minimizing the energy of the effective kink model. \section{Conclusion}\quad It is an open problem to understand the phases of cold and dense holographic nuclear matter as a function of baryon number density. The distribution of baryonic matter in both the non-holographic and holographic directions is unknown and this has led to a number of different approximate descriptions and suggestions for phenomena that might occur. In this paper we have assumed homogeneity in the non-holographic spatial directions to investigate the distribution in the holographic direction. As the holographic coordinate corresponds to an energy scale in the boundary theory, then understanding this aspect is likely to be the key to a holographic description of the baryonic Fermi surface of a quarkyonic phase. Our homogeneous holographic nuclear matter is described by an effective holographic kink theory, which we have shown is capable of a simultaneous realization of a number of previously suggested phases, including dyonic salt, baryonic popcorn and soliton bags. An advantage of this unifying description is that we can compare the various phases and determine which phase is preferred as the baryon chemical potential varies. The effective kink model also successfully reproduces the QCD behaviour of a first order phase transition to lightly bound nuclear matter, at a value of the baryon chemical potential that is just below the baryon mass, in agreement with QCD. In holographic QCD the spectrum of fluctuations in the holographic direction determines the masses of the vector and axial vector mesons. As the background solution changes with increasing baryon number density then so does the spectrum, and this has been proposed \cite{deBoer:2012ij} as a mechanism for approximate chiral symmetry restoration, as the masses of vector and axial vector mesons could converge with increasing density. However, as the details of the background solution are unavailable, to date the investigations of this issue have been mostly qualitative. It would be interesting to study this aspect within our effective kink model and to compute some quantitative data. Even within our homogeneous approximation, this is still a non-trivial calculation, because of the mixing of various modes, but it should be a tractable problem that we hope to address in the near future. Other avenues for future research include extending our approach to the finite temperature regime and investigating how kinky holographic nuclear matter responds to external electric and magnetic fields by introducing new boundary conditions for $\hat A_\mu$, as in \cite{Bergman:2008sg}. Given that our soliton bag has some features in common with the magnetic bag description of a large number of coincident $SU(2)$ magnetic monopoles, it could be interesting to study both problems for higher rank gauge groups, as new features certainly emerge for non-abelian monopoles beyond $SU(2)$. In this paper we have restricted our investigations to the effective five-dimensional Yang-Mills-Chern-Simons version of the Sakai-Sugimoto model. However, it would be a simple matter to employ our ansatz directly in the full string theory version of the Sakai-Sugimoto model, with the usual caveat that a prescription must be employed to deal with the non-abelian Dirac-Born-Infeld action. We expect the same kind of behaviour as in the five-dimensional Yang-Mills-Chern-Simons version of the theory. \section*{Appendix}\renewcommand{\thesection}{A}\setcounter{equation}{0}\quad In this appendix we consider the Skyrme model and explain how the homogeneous approximation may be thought of as a smeared version of the Skyrme crystal. In particular, we show that the homogeneous approximation is an unattainable idealization that provides a lower bound on the energy of the true Skyrme crystal. In terms of the hermitian currents $R_i=i(\partial_i U)U^{-1}$ the static energy of the Skyrme model (in Skyrme units) is \begin{equation} E=\int\bigg( \frac{1}{2}\mbox{Tr}(R_iR_i)-\frac{1}{16}\mbox{Tr}([R_i,R_j][R_i,R_j])\bigg)\,d^3x \label{skyrmeenergy} \end{equation} and the baryon number is \begin{equation} B=\frac{i}{24\pi^2}\int \epsilon_{ijk} \mbox{Tr}(R_iR_jR_k) \,d^3x. \label{skyrmebaryon} \end{equation} The Faddeev-Bogomolny bound is $E\ge 12\pi^2 B$, but for non-zero $B$ this bound cannot be attained, as it requires that the Skyrme field $U({\bf x})$ is an isometry from $\mathbb{R}^3$ to $SU(2)$, which are two spaces that are not isometric. The solution of the Skyrme model that is closest to the bound is the triply periodic Skyrme crystal with $E/B=12\pi^2\times 1.04.$ This is a cubic lattice that contains four Skyrmions within a cube of side length $L=4.7$, and hence has a baryon number density ${\cal B}=4/L^3=0.04.$ Substituting our idealized homogeneous approximation $R_i=-\beta\sigma_i/2$ into (\ref{skyrmeenergy}) and (\ref{skyrmebaryon}) gives the energy per unit volume ${\cal E}$ and the baryon number density ${\cal B}$ to be \begin{equation} {\cal E}=\frac{3}{4}\beta^2+\frac{3}{16}\beta^4, \qquad\qquad {\cal B}=\frac{\beta^3}{16\pi^2}. \end{equation} Using these expressions, we recover the energy bound by the simple manipulation \begin{equation} {\cal E}=\frac{3}{4}\bigg(\beta-\frac{1}{2}\beta^2\bigg)^2+\frac{3}{4}\beta^3\ge \frac{3}{4}\beta^3=12\pi^2{\cal B}. \end{equation} Within the homogeneous approximation the bound is attained by $\beta=2$ with a corresponding baryon number density ${\cal B}=1/(2\pi^2)=0.05.$ These values provide a good estimate of both the energy and the baryon number density of the Skyrme crystal. The homogeneous approximation generates a lower bound for the true energy of the Skyrme crystal because there are no Skyrme fields that generate the idealized homogeneous currents required to attain the bound. \section*{Acknowledgements} This work is funded by the STFC grant ST/J000426/1 and an STFC PhD Studentship. We thank Andreas Schmitt for useful discussions.
f99681534bdabea5fb852d29e931cf29d07fb71d
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The standard model (SM) has been tested at an unprecedented level of precision through various experiments. However, some excesses have not yet been completely resolved. The first case is the muon anomalous magnetic moment {(muon $g-2$)}, where the discrepancy between experimental data and the SM prediction is currently $\Delta a_\mu = a^{\rm exp}_\mu - a^{\rm SM}_\mu =(28.8\pm 8.0)\times 10^{-10}$~\cite{PDG}. The second case is the angular observable $P'_5$ of $B\to K^* \mu^+ \mu^-$~\cite{DescotesGenon:2012zf}, where a $3.4\sigma$ deviation, resulting from the integrated luminosity of 3.0 fb$^{-1}$ at the LHCb~\cite{Aaij:2015oid}, recently confirmed an earlier result with $3.7\sigma$ deviations~\cite{Aaij:2013qta}. Moreover, the same measurement with $2.1\sigma$ deviations was reported by Belle~\cite{Abdesselam:2016llu}. Also, the other relevant $P_i$ observables are defined in Ref.~\cite{Matias:2012xw}. Various possible resolutions to this excess have been widely studied~\cite{Descotes-Genon:2013wba,Gauld:2013qja,Datta:2013kja,Hurth:2013ssa,Descotes-Genon:2014uoa,Altmannshofer:2014rta,Crivellin:2015mga,Sahoo:2015wya,Straub:2015ica,Crivellin:2015era,Lee:2015qra,Alonso:2015sja,Sahoo:2015qha,Sahoo:2015pzk,Chiang:2016qov,Belanger:2015nma,Dorsner:2016wpm,Becirevic:2015asa,Descotes-Genon:2015uva,Boucenna:2016wpr,Hiller:2016kry}. The third case is the ratio $R_K = BR(B^+\to K^+\mu^+\mu^-)/BR(B^+\to K^+ e^+e^-)$, where $BR(B^+ \to K^+ \ell^+ \ell^-)$ is the branching ratio (BR) of the decay $B^+ \to K^+ \ell^+ \ell^-$; and the LHCb measurement shows a $2.6\sigma$ deviation from the SM result~\cite{Aaij:2014ora}. In order to explain the deviation, various mechanisms have been proposed~\cite{Crivellin:2015mga,Hiller:2014yaa,Hurth:2014vma,Gripaios:2014tna,Glashow:2014iga,Sahoo:2015fla,Bauer:2015knc,Das:2016vkr,Li:2016vvp,Becirevic:2016yqi,Sahoo:2016pet,Bhattacharya:2016mcc,Duraisamy:2016gsd}. In addition to the excesses mentioned above, the LHC with energetic $pp$ collisions can also be a good place to test the SM and provide possible excess signals. For instance, a hint of resonance with a mass of around 750 GeV in the diphoton invariant mass spectrum was indicated by the ATLAS~\cite{Aaboud:2016tru} and CMS~\cite{Khachatryan:2016hje} experiments. Due to the results, various proposals have been broadly proposed and studied~\cite{Harigaya:2015ezk,Backovic:2015fnp,Angelescu:2015uiz,Nakai:2015ptz,Buttazzo:2015txu,DiChiara:2015vdm,Knapen:2015dap,Pilaftsis:2015ycr,Franceschini:2015kwy,Ellis:2015oso,Gupta:2015zzs,Kobakhidze:2015ldh,Falkowski:2015swt,Benbrik:2015fyz,Wang:2015kuj,Dev:2015isx,Allanach:2015ixl,Wang:2015omi,Chiang:2015tqz,Huang:2015svl,Ko:2016wce,Nomura:2016seu,Kanemura:2015bli,Cheung:2015cug,Nomura:2016fzs,Ko:2016lai,Belanger:2016ywb,Dorsner:2016ypw,Han:2016bvl,Mambrini:2015wyu}. Although it turns out that the resonance has not been confirmed by the updating data of ATLAS~\cite{ATLAS:2016eeo} and CMS~\cite{CMS:2016crm} and has been shown to be more like a statistical fluctuation, the search for the new exotic events in the LHC still continue and is an essential mission. To resolve the excesses in a specific framework, we propose the extension of the SM by including leptoquarks (LQs), where the LQs are colored scalars that simultaneously couple to the leptons and quarks. Hence, the $b\to s \ell^+ \ell^-$ decays can arise from the tree-level LQ-mediated Feynman diagrams when the muon $g-2$ is induced from LQ loops. In addition to the decays $B\to K^{(*)} \ell^+ \ell^-$, the effective interactions for $b\to s \ell^+ \ell^-$ can also contribute to $B_s \to \mu^+ \mu^-$, where the BR, measured by LHCb and CMS~\cite{CMS:2014xfa}, is given as: \begin{equation} \text{BR}(B_s \to \mu^+ \mu^-)^{\text exp} = (2.8^{+0.7}_{-0.6}) \times 10^{-9} \quad \quad \text{(LHCb-CMS)}\,. \end{equation} We note that the dominant effective couplings for $b\to s \ell^+ \ell^-$ processes are denoted by the Wilson coefficients $C_{9,10}$. Usually, both $C_9$ and $C_{10}$ are strongly correlated. Since this experimental result is consistent with the SM prediction of $\text{BR}(B_s \to \mu^+ \mu^-)^{\text SM} \approx 3.65\times 10^{-9}$~\cite{Bobeth:2013uxa}, in order to accommodate the anomalies of $P'_5$ and $R_K$ to the measurement of $B_s \to \mu^+ \mu^-$, we introduce two LQs with different representations of $SU(2)_L$ into the model. Thus, the correlation between $C_{9}$ and $C_{10}$ is diminished. It is found that when the $C_{10}$ is constrained by $B_s\to \mu^+ \mu^-$, the $C_9$ then can satisfy the requirements from the global analysis of $B\to K^* \mu^+ \mu^-$, and can also explain the anomaly of $R_K$, and the muon $g-2$ can fit the current data. The colored scalar LQs can couple to the SM Higgs in the scalar potential; thus, the LQ effects can influence the SM Higgs production and decays. The Higgs measurements have approached the precision level since the SM Higgs was discovered. Any sizable deviations from the SM predictions will indicate new physics. In this study, we analyze the LQ-loop contributions to the diphoton Higgs decay. It is worth mentioning that the introduced LQs can significantly enhance the production cross section of a heavy scalar boson if such heavy scalar is probed at the LHC in the future. The relevant studies on the heavy scalar production via LQ couplings can be found in Refs.~\cite{DiIura:2016wbx,Dey:2016sht,Deppisch:2016qqd,Hati:2016thk,Chao:2015nac,Murphy:2015kag,Bauer:2015boy}. The paper is organized as follows. We introduce the model and discuss the relevant couplings in Sec.~II. In Sec.~III, we study the phenomena: the SM Higgs diphoton decay, LFV processes, Wilson coefficients of $C_{9,10}$ from LQ contributions for $b\to s\ell^+ \ell^-$ decays, and the implication of $B_s \to \mu^+ \mu^-$. The conclusion is given in Sec. IV. \section{ Couplings to the leptoquarks} In this section, we briefly introduce the model and relevant interactions with the LQs. To reconcile the measurements of $B_s\to \mu ^+ \mu^-$ and $B\to K^* \ell^+ \ell^-$, we extend the SM by adding two different representations of LQ, which are $\Phi_{7/6}=(3,2)_{7/6}$ and $\Delta_{1/3}=(\bar 3,3)_{1/3}$ under $(SU(3)_C, SU(2)_L)_{U(1)_Y}$ SM gauge symmetry. The gauge-invariant Yukawa interactions of the SM fermions and LQs are written as: \begin{align} \label{eq:int1} L_{LQ} = k_{ij} \bar{Q}_i \Phi_{7/6 } \ell_{Rj} + \tilde{k}_{ij} \bar L_i \tilde \Phi_{7/6} u_{Rj} + y_{ij} \bar{Q}_i^c i \sigma_2 \Delta_{1/3} L_j + h.c.\,, \end{align} where the subscripts $i\,, j$ are the flavor indices; $L^T_i = (\nu_i, \ell_i^-)$ and $Q^T_i = (u_i, d_i)$ are the $SU(2)_L$ lepton and quark doublets, $\tilde \Phi_{7/6} =i \sigma_2 \Phi^*_{7/6}$, and $(k_{ij}, \tilde k_{ij}, y_{ij})$ are the Yukawa couplings. Since we do not study the CP violating effects, hereafter, we take all new Yukawa couplings as real numbers. We use the representations of the LQs as: \begin{align} \Phi_{7/6} = \begin{pmatrix} \phi^{5/3} \\ \phi^{2/3} \end{pmatrix}\,, \ \Delta_{1/3} = \begin{pmatrix} \delta^{1/3}/\sqrt{2} & \delta^{4/3} \\ \delta^{-2/3} & - \delta^{1/3}/\sqrt{2} \end{pmatrix}\,, \end{align} where the superscripts are the electric charges of the particles. The interactions in Eq.~(\ref{eq:int1}) are then expressed as: \begin{align} \label{eq:int2} L_{LQ} =& k_{ij} \left[ \bar u_{Li} \, \ell_{Rj} \phi^{5/3} + \bar d_{Li} \, \ell_{R j} \phi^{2/3} \right] + \tilde{k}_{ij} \left[ \bar \ell_{L i} \, u_{Rj} \phi^{-5/3} - \bar \nu_{L i}\, u_{R j} \phi^{-2/3} \right] \nonumber \\ &+ y_{ij} \left[ \bar u^c_{Li} \, \nu_{Lj} \delta^{-2/3} - \frac{1}{\sqrt{2}} \bar u^c_{L i} \, \ell_{L j} \delta^{1/3} - \frac{1}{\sqrt{2}} \bar d^c_{L i} \, \nu_{L j} \delta^{1/3} - \bar d^c_{L i} \, \ell_{L j} \delta^{4/3} \right] + h.c. \end{align} Since the LQs are colored scalar bosons, they can couple to the SM Higgs $H$ via the scalar potential. In order to get the Higgs couplings to the LQs, we write the gauge-invariant scalar potential as: \begin{align} \label{eq:potential} V & = \mu^2 H^\dagger H + \lambda (H^\dagger H)^2+ M_\Phi^2 \left(\Phi_{7/6}^\dagger \Phi_{7/6} \right) + M_{\Delta}^2 Tr \left( \Delta_{1/3}^\dagger \Delta_{1/3} \right) + \lambda_\Phi \left(\Phi_{7/6}^\dagger \Phi_{7/6} \right)^2 \nonumber \\ % & + \lambda_\Delta \left [Tr \left( \Delta_{1/3}^\dagger \Delta_{1/3}\right) \right]^2 + \lambda'_\Delta Tr \left( \left[ \Delta_{1/3}^\dagger \Delta_{1/3} \right]^2 \right) + \lambda_{H \Phi} (H^\dagger H)\left(\Phi_{7/6}^\dagger \Phi_{7/6} \right) \nonumber \\ &+ \lambda_{H \Delta} (H^\dagger H) Tr\left( \Delta_{1/3}^\dagger \Delta_{1/3} \right) + \lambda_{\Phi \Delta} \left(\Phi_{7/6}^\dagger \Phi_{7/6} \right) Tr \left( \Delta_{1/3}^\dagger \Delta_{1/3} \right)\,. \end{align} As usual, we adopt the representations of the Higgs doublet $H$ as: \begin{equation} H = \begin{pmatrix} G^+ \\ \frac{1}{\sqrt{2}} ( v+ \phi + iG^0) \end{pmatrix}\,, \end{equation} where $G^+$ and $G^0$ are the Goldstone bosons; $\phi$ is the SM Higgs field, and $v$ is the vacuum expectation value (VEV) of $H$. It is known that the VEV of scalar field is dictated by the scalar potential. \section{Phenomenological Analysis} Based on the introduced new interactions, in this section, we study the implications of the Higgs diphoton decay, $\ell_i \to \ell_j \gamma$, the muon $g-2$, $h\to \tau \mu$, $B\to K^* \ell^+ \ell^-$, and $R_K$. Since each of these processes has its own unique characteristics, we discuss these phenomena one by one below. \subsection{Higgs diphoton decay} The Higgs measurement is usually described by the signal strength parameter, which is defined as the ratio of observation to the SM prediction and expressed as: \begin{equation} \mu^f_i =\frac{\sigma(pp\to h) }{\sigma(pp\to h)_{\rm SM} } \cdot \frac{ \text{BR}(h\to f)}{ \text{BR}(h\to f)_{\rm SM}} \equiv \mu_i \cdot \mu_f \,, \label{eq:muf} \end{equation} where $f$ stands for the possible channels, and $\mu_i (\mu_f)$ denotes the signal strength of production (decay). Although vector-boson fusion (VBF) can also produce the SM Higgs, we only consider the gluon-gluon fusion (ggF) process because it is the most dominant. The diphoton Higgs decay approached the precision measurement since the 125 GeV Higgs was observed. Therefore, any significant deviation from the SM prediction (i.e., $\mu^f_i \neq 1$) can imply the new physical effects. As stated earlier, the SM Higgs can couple to the LQs via the scalar potential. From Eq.~(\ref{eq:potential}), it can be seen that after spontaneous symmetry breaking (SSB), the quartic terms $H^\dagger H\Phi^\dagger_{7/6} \Phi_{7/6}$ and $H^\dagger H Tr(\Delta^\dagger_{1/3} \Delta_{1/3})$ can lead to trilinear couplings of Higgs to LQs as: \begin{align} {\cal L} & \supset \mu_{h \Phi} h \left(\phi^{-5/3} \phi^{5/3} + \phi^{-2/3} \phi^{2/3} \right) + \mu_{h \Delta}h \left( \delta^{-1/3} \delta^{1/3} + \delta^{-2/3} \delta^{2/3} + \delta^{-4/3} \delta^{4/3} \right)\,, \label{eq:hLQ} \end{align} where $\mu_{h\Phi} = \lambda_{H\Phi} v$ and $\mu_{h\Delta} = \lambda_{H\Delta} v$. With the couplings in Eq.~(\ref{eq:hLQ}), the effective Lagrangian for $hgg$ by LQ-loop can be formulated as: % \begin{equation} \Delta {\cal L}_{hgg} = \frac{\alpha_s}{8 \pi} \left( \frac{ \mu_{h \Phi} }{m_\Phi^2 } A_0(\xi_\Phi) + \frac{3 \mu_{h \Delta}}{2 m_\Delta^2} A_0(\xi_\Delta) \right) h G^{a \mu \nu} G^a_{\mu \nu}\,, \label{eq:hgg} \end{equation} where $\xi_X = 4 m_{X}^2/m_h^2$ and the loop function is given by: \begin{equation} A_0 (x) = x (1 -x f(x)) \end{equation} with $f(x) = \left[\sin^{-1} (1/\sqrt{x})\right]^2 $ for $x > 1$. Accordingly, the signal strength of the Higgs production and decay to diphoton can be respectively obtained as: \begin{align} \mu_i & = \left| 1 + \frac{v}{A_{1/2}(\xi_t)}\sum_{X=\Phi,\Delta} \frac{n_X \mu_{hX} }{m^2_X}A_0(\xi_X) \right|^2\,, \nonumber \\ \mu_{\gamma \gamma } & = \left| 1 + \frac{v N_c}{2} \frac{\sum_{X=\Phi, \Delta} Q^2_X A_0(\xi_X) \mu_{hX}/m^2_X} {A_1(\xi_W) + Q^2_t N_c A_{1/2}(\xi_t)} \right|^2, \end{align} where $n_{\Phi(\Delta)}=2(3)$, $N_c=3$ is the number of colors; $Q^2_\Phi = 29/9$ and $Q^2_\Delta = 21/9$, and the functions for vector-boson and fermion loops are given by \begin{align} A_{1/2}(x) & = -2[x + (1-x) f(x))]\,, \nonumber \\ A_1(x) & = 2 + 3 x + 3(2x -x^2)f(x)\,. \end{align} Since the effects of the doublet and triplet LQs are similar, for simplicity, we set $\mu_{h\Phi}=\mu_{h \Delta} = \mu_{LQ}$ and $m_\Phi = m_\Delta = m_{LQ}$. The $\mu^{\gamma\gamma}_i$ as a function of $m_{LQ}$ is presented in Fig.~\ref{fig:muif}(a), and that of $\mu_{LQ}$ is shown in Fig.~\ref{fig:muif}(b), where the curves in plot (a) are $\mu_{LQ}=0.1,0.5,1$ TeV, and those in plot (b) are $m_{LQ}=0.8, 0.9, 1.0$ TeV. For comparison, we also show the results of ATLAS~\cite{Aad:2015gba} and CMS~\cite{CMS:2014ega} with $1\sigma$ errors in the plots. From the plots, it can be clearly seen that with $\mu_{LQ}$ of ${\cal O}(100)$ GeV, the LQ contributions can significantly shift the $\mu^{\gamma\gamma}_i$ away from the SM prediction and that the results are consistent with the current data. On the contrary, the $\mu^{\gamma\gamma}_i$ approaches the SM result when $\mu_{LQ}$ is of the order of GeV. \begin{figure}[hpbt] \includegraphics[width=70mm]{hdecay1.eps} \includegraphics[width=70mm]{hdecay2.eps} \caption{ Diphoton signal strength parameter $\mu^{\gamma\gamma}_i$ as a function of (a) $m_{LQ}$ and (b) $\mu_{LQ}$, where the curves in plots (a) and (b) denote $\mu_{LQ}=(0.1, 0.5,1)$ TeV and $m_{LQ}=(0.8,0.9,1.0)$ TeV, respectively. } \label{fig:muif} \end{figure} \subsection{ Radiative and Higgs LFV processes, muon $g-2$, and $B \to K^{(*)} \ell^+ \ell'^-$ decays} In the following analysis, we study the rare lepton-flavor violating processes, e.g. $\ell_i \to \ell_j \gamma$ and $h\to \bar\tau \mu + \bar\mu \tau$, muon $g-2$ $\Delta a_\mu$, and the FCNC process $B\to K^* \ell^+ \ell^-$. We first discuss the radiative LFV processes for $\ell_i \to \ell_j \gamma$. With the couplings in Eq.~(\ref{eq:int2}), the LQ-loop induced decay amplitude for $\ell_i \to \ell_j \gamma$ can be written as: \begin{eqnarray} {\cal L}_{\ell_i \to \ell_j \gamma} = \frac{e}{2} \bar \ell_j \sigma_{\mu\nu} \left[ (c_L)_{ji} P_L + (c_R)_{ji} P_R \right] \ell_i F^{\mu \nu} \,,\label{eq:Lllga} % \end{eqnarray} where the coefficient $(c_{R})_{ji}$ is expressed as: \begin{align} (c_R)_{ji} & \approx \frac{m_t}{(4 \pi)^2} ( k^\dagger)_{i 3} \tilde k_{3 j} \int d[X] \left( \frac{5}{\Delta(m_{t}, m_\Phi) } - \frac{2(1-x)}{\Delta(m_\Phi, m_{t}) }\right) \,, \nonumber \\ \Delta(m_1,m_2) & = x m_1^2 +(y+z) m_2^2 \,, \nonumber \\ \int [dX] & = \int dx dy dz \delta(1-x-y-z)\,, \label{eq:cRji} \end{align} $(c_L)_{ji}$ can be obtained from $(c_R)_{ji}$ by exchanging $k_{ab}$ and $\tilde k_{ab}$. In order to balance the chirality of the leptons, it is found that the contributions from $k^\dagger_{iq} k_{q j}$, $\tilde{k}^\dagger_{iq} \tilde{k}_{q j}$, $y^\dagger_{iq} k_{qj}$, and $y^\dagger_{iq} y_{qj}$ are suppressed by the lepton masses. Since the LQ $\phi^{5/3}$ can couple to left-handed and right-handed up-type quarks, the chirality flip by the mass insertion in the propagator of the up-type quark can lead to freeing of the lepton masses in the Feynman diagrams, which are associated with $k_{qi}$ and $\tilde k_{qi}$. In addition, the top-quark is much heavier than the $u$- and $c$-quarks; therefore, we only present the top-quark contribution in $(c_R)_{ji}$. Straightforwardly, the BR for $\ell_i \to \ell_j \gamma$ can be expressed as: \begin{equation} \text{BR}(\ell_i \to \ell_j \gamma) = \frac{48 \pi^3 \alpha \eta_i}{G_F^2 m_{\ell_i}^2} \left( \left| (c_R)_{ji} \right|^2 + \left| (c_L)_{ji} \right|^2 \right)\,, \end{equation} where $\eta_i \simeq (1,1/5)$ for $i = (\mu, \tau)$ and the BRs for $\ell_i \to \ell_j \bar \nu_j \nu_i$ in the SM have been applied. The current experimental upper limits are shown in Table~\ref{tab:Cif}. According to Eq.~(\ref{eq:Lllga}), muon $g-2$ can be easily obtained by setting $j=i=\mu$ and found as: \begin{equation} \Delta a_\mu \simeq - \frac{m_\mu}{2} (c_L + c_R)_{\mu\mu}. \label{eq:mug2} \end{equation} \begin{table}[hpbt] \begin{tabular}{c|c|c} \hline Process & $(i,j)$ & Experimental bounds ($90\%$ CL) \\ \hline $\mu^{-} \to e^{-} \gamma$ & $(2,1)$ & $\text{BR}(\mu \to e\gamma) < 5.7 \times 10^{-13}$ \\ $\tau^{-} \to e^{-} \gamma$ & $(3,1)$ & $\text{BR}(\tau \to e\gamma) < 3.3 \times 10^{-8}$ \\ $\tau^{-} \to \mu^{-} \gamma$ & $(3,2)$ & $\text{BR}(\tau \to \mu\gamma) < 4.4 \times 10^{-8}$ \\ \hline \end{tabular} \caption{ Current upper bounds on the BRs for the decays $\ell_i \to \ell_j \gamma$~\cite{Adam:2013mnn}.} \label{tab:Cif} \end{table} If the photon in $\ell_i \to \ell_j \gamma$ is replaced by the Higgs, similar Feynman diagrams can contribute to $h\to \bar \ell_j \ell_i + \bar \ell_i \ell_j \equiv \ell_j \ell_i$. Since the upper limit of $\text{BR}(\mu\to e\gamma)$ is of ${\cal O}(10^{-13})$ and can give strong constraints on the parameters $k^\dagger_{23}\tilde k_{31}$ and $\tilde{k}^\dagger_{23} k_{31}$, if we set $k_{31}$ and $\tilde k_{31}$ to be small, then it is apparent that $h\to e \mu$ and $h\to e \tau$ are much smaller than current upper limits. Hence, we just study the decay $h\to \mu \tau$. The one-loop induced effective couplings for $h \mu \tau$ are written as: \begin{equation} \mathcal{L} = h \bar \mu (C_R P_R + C_L P_L) \tau + h.c.\,, \end{equation} where $C_{L}$ is expressed as~\cite{Baek:2015mea,Baek:2016kud}: \begin{align} C_L =& \frac{(k^\dagger)_{23} \tilde k_{33} N_c m_t}{(4 \pi)^2 v} \left[ A\left(\frac{m^2_t}{m^2_\Phi}, \frac{m^2_h}{m^2_\Phi} \right) + B\left(\frac{m^2_t}{m^2_\Phi}, \frac{m^2_h}{m^2_\Phi} \right) \right] \nonumber \\ & + \frac{ N_c \mu_{h \Phi} }{(4\pi)^2} \sum_{i=1\text{-}3, q=u,d} \left[ m_\mu (k^\dagger)_{2 i} k_{i 3} G(m_{q_i},m_\Phi )+ m_\tau \tilde k^\dagger_{2 i} \tilde k_{i 3} \tilde G(m_{q_i},m_\Phi) \right] \nonumber \\ & + \frac{N_c m_\tau \mu_{h \Delta}}{(4 \pi)^2} \sum_{i=1\text{-}3,q=u,d} (y^\dagger)_{2 i} y_{i 3} \tilde G(m_{q_i},m_\Delta) \,, \label{eq:CL} \end{align} $C_R$ can be obtained from $C_L$ by exchanging $k_{ab}$ and $\tilde k_{ab}$, and the loop functions are given by: \begin{align} A(r_t, r_h) =& -\frac{1}{2} -2 \int [dX] \log \left[z+(1-z)r_t - x y r_h - i \epsilon \right] \nonumber \\ & + \int_0^1 dx \log \left[ x + (1-x) r_t - i \epsilon \right]\,, \nonumber \\ B(r_t, r_h) = & \int [dX] \frac{xy r_h - r_t }{z+(1-z)r_t - xy r_h} \,, \nonumber \\ G(m_1,m_2) \approx & \int [dX] \frac{ z}{- x z m_h^2 + x m_1^2 + (y+z) m_2^2 }\,, \nonumber \\ \tilde G(m_1,m_2) \approx & \int [dX] \frac{ y}{ - x z m_h^2 + x m_1^2 + (y+z) m_2^2 } \,. \end{align} The $\epsilon$ in $A(r_t, r_h)$ denotes an infinitesimal positive value. It can be seen that the terms associated with $k^\dagger_{2i} k_{i3}$, $\tilde{k}^\dagger_{2i} k_{i3}$, and $y^\dagger_{2i} y_{i3}$ in Eq.~(\ref{eq:CL}) are proportional to the lepton masses. The situation is similar to the $(c_R)_{ji}$ in the decays $\ell_i \to \ell_j \gamma$. Although $\mu_{hX}$ of TeV ($X=\Phi, \Delta)$ can enhance these effects, due to the effects of being related to $\mu_{hX} m_\ell /m^2_X$, their contributions are at least $10^{-2}$ smaller than those from $k^\dagger_{2i} \tilde{k}_{i3}$. Accordingly, the BR for $h \to \mu \tau$ is formulated as: \begin{align} & \text{BR}(h\to \mu \tau) \approx \frac{m_h}{16\pi \Gamma_h} (|C_L|^2 + | C_R|^2)\,, \end{align} where $\Gamma_h$ is the width of the Higgs. Due to $\text{BR}(h\to \mu \tau)$ being less than $1\%$, we use $\Gamma_h\approx \Gamma^{\rm SM}_h \approx 4.2$ MeV in our numerical estimations. Next, we discuss the decays for $b\to s \ell^+ \ell^-$. In order to include the effects of lepton non-unversality, we write the effective Hamiltonian as: \begin{equation} {\cal H} = \frac{G_F \alpha V_{tb} V^*_{ts}}{\sqrt{2}\pi} \left[ H_{1\mu} L^\mu + H_{2\mu} L^{5\mu} \right]\,, \label{eq:Hbsll} \end{equation} where the leptonic currents are denoted by $L^{(5)}_\mu= \ell \gamma_\mu (\gamma_5) \ell$; and the related hadronic currents are defined as: \begin{align} H_{1\mu} &= C^\ell_9 \bar s \gamma_\mu P_L - \frac{2m_b}{q^2} C_7 \bar s i \sigma_{\mu \nu} q^\nu P_R b\,, \nonumber \\ % H_{2\mu} & = C^{\ell}_{10} \bar s \gamma_\mu P_L b\,. \end{align} Here, the Wilson coefficients are read as: $C^\ell_{9(10)} = C^{\rm SM}_{9(10)} + C^{\rm NP,\ell}_{9(10)}$ and $C_7 = C^{\rm SM}_{7}$. The detailed angular distribution for $B\to (K\pi)_{K^*} \ell^+ \ell^-$ can be found in Refs.~\cite{DescotesGenon:2012zf,Chen:2002bq,Chen:2002zk,Altmannshofer:2008dz,Egede:2010zc}. Following the notations in Ref.~\cite{DescotesGenon:2012zf}, the angular observable $P'_5$ is defined by: \begin{align} P'_5 &= \frac{J_5}{ \sqrt{-J_{2c} J_{2s} }} \,, \quad J_5 = \sqrt{2} Re(A^L_0 A^{L*}_{\perp} )\,, \nonumber \\ J_{2c} &=- |A^L_0|^2 \,, \quad J_{2s} =\frac{1}{4} \left(|A^L_{\parallel}|^2 + |A^L_{\perp}|^2 \right) \,, \end{align} where $A^{L}_{0,\parallel,\perp}$ are related to the $B\to K^* $ transition form factors and the Wilson coefficients of $C^\ell_{9, 10}$ and $C_7$. Their explicit expressions can be found in Ref.~\cite{DescotesGenon:2012zf}. In this study, we do not directly investigate the angular analysis of $B\to K^* \ell^+ \ell^-$; instead, we refer to the results, which were done by using the global analysis to get the best-fit value of $C^{NP}_9 \approx -1.09$ for the new physics contributions~\cite{Descotes-Genon:2015uva}. Thus, we just derive the Wilson coefficients of $C^\ell_9$ and $C^\ell_{10}$ from the LQ contributions. With the Yukawa couplings in Eq.~(\ref{eq:int2}), the effective Hamiltonian for $b\to s \ell^+ \ell^-$ mediated by $\phi^{2/3}$ and $\delta^{4/3}$ can be respectively found as: \begin{align} & H_{\rm eff}^1 = \frac{k_{b \ell} k_{s \ell}}{2 m^2_{\Phi}} (\bar{s}\gamma^\mu P_L b)(\bar{\ell} \gamma_\mu P_{ R} \ell)\,, \nonumber \\ & H_{\rm eff}^2 = -\frac{y_{b \ell} y_{s \ell}}{2 m^2_{\Delta}} (\bar{s}\gamma^\mu P_L b)(\bar{\ell} \gamma_\mu P_{L} \ell)\,. \label{eq:Heff} \end{align} We can decompose the Eq.~(\ref{eq:Heff}) in terms of the effective operators $O_9$ and $O_{10}$, defined as $O_{9(10)} = \bar s \gamma_\mu P_L b \; \bar \ell \gamma^\mu (\gamma_5) \ell $. The associated Wilson coefficients of $O_{9,10}$ from the LQs then are found as: \begin{align} C^{LQ,\ell}_9 &= - \frac{1}{c_{\rm SM}} \left(\frac{k_{b \ell} k_{s \ell}}{4 m^2_{\Phi}} -\frac{y_{b \ell} y_{s \ell}}{4 m^2_{\Delta}} \right)\,, \nonumber \\ C^{LQ,\ell}_{10} &= \frac{1}{c_{\rm SM}} \left(\frac{k_{b \ell} k_{s \ell}}{4 m^2_{\Phi}} + \frac{y_{b \ell} y_{s \ell}}{4 m^2_{\Delta}} \right)\,, \label{eq:C9C10} \end{align} where $c_{\rm SM} = V_{tb} V^*_{ts} \alpha G_F/(\sqrt{2} \pi) $ is a scale factor from the SM effective Hamiltonian. It is worth mentioning that the interaction $C^{LQ,\mu}_{10} O_{10}$ can contribute to $B_s \to \mu^+ \mu^-$. Since the experimental data are consistent with the SM prediction, to consider the constraint from $B_s \to \mu^+ \mu^-$, we adopt the expression for the BR as~\cite{Hiller:2014yaa}: \begin{equation} \frac{\text{BR}(B_s \to \mu^+ \mu^-)}{\text{BR}(B_s \to \mu^+ \mu^-)^{\text SM}} = \left|1-0.24 C^{LQ,\mu}_{10} \right|^2. \end{equation} With $1\sigma$ errors, the allowed range for $C^{LQ,\mu}_{10}$ is obtained as $C^{LQ,\mu}_{10}=(0.21, 0.79)$. We use this result to constrain the free parameters. Since the $R_K$ is insensitive to the $B\to K$ transition form factors~\cite{Hiller:2003js}, in order to study the anomaly of $R_K$, we require that the allowed range of parameters has to satisfy~\cite{Hiller:2014yaa}: \begin{align} 0.7 \leq Re[X^e - X^\mu] \leq 1.5\,, \label{eq:X} \end{align} where $X^\ell = C^{LQ,\ell}_9 - C^{LQ,\ell}_{10}$, and the $R_K$ data with $1\sigma$ errors are used. Since the parameters in the decays $\ell_i \to \ell_j \gamma$, $h\to \mu \tau$, $\Delta a_\mu$, and $B\to K^{(*)} \ell^+ \ell^-$ are strongly correlated, in the following analysis, we take the current upper limits of $\text{BR}(\ell_i \to \ell_j \gamma)$ shown in Table~\ref{tab:Cif} as the inputs and attempt to find the allowed parameter space, such that the excesses in $\Delta a_\mu$ and $B\to K^{(*)} \ell^+ \ell^-$ can be satisfied, and the $\text{BR}(h\to \mu \tau)$ can be as large as possible. From $(c_R)_{ji}$ in Eq.~(\ref{eq:cRji}), the dominant effects on the radiative LFV processes are from the $\phi^{5/3}$ and the top-quark loop; thus, there is no possible cancellation in any of the decay amplitudes. With the upper bound of $\text{BR}(\mu\to e \gamma)$, we see that $k^\dagger_{13} \tilde k_{32}$ and $\tilde k^\dagger_{13} k_{32}$ have to be very small. In order to explain the excesses of muon $g-2$ and $B\to K^* \mu^+ \mu^-$, we set $k_{31}=\tilde k_{31}\approx 0$. As a result, $\text{BR}(h\to e \mu)$ is negligible in this model. The related parameters for $\tau \to (\mu ,e) \gamma$ decays are $k_{31,32}\tilde k_{33}$ and $\tilde k_{31,32} k_{33}$, respectively. These parameters simultaneously influence $h\to (\mu, e) \tau$, muon $g-2$, and $b\to s \mu^+ \mu^-$; therefore we have to analyze these processes together to get the allowed parameter space. Since Eqs.~(\ref{eq:cRji}), (\ref{eq:CL}), and (\ref{eq:C9C10}) involve many free parameters, in order to efficiently perform a numerical analysis, we set the ranges of relevant parameters as: \begin{align} & m_{LQ} \in [700\,, 1500\,] \; \text{GeV},\quad \mu_{LQ} \in [1\,, 100]\; \text{GeV}\,, \quad \{ k_{22}, \tilde k_{22}, y_{22} \} \in [-0.1\,, 0.1]\,,\nonumber \\ & \{ k_{33},\tilde k_{33}, y_{33} \} \in [-0.01,0.01]\,, \quad \{ k_{23},\tilde k_{23}, y_{23} \} \in [-0.1\,, 0.1]\,, \nonumber \\ & k_{32} \in \text{sign}(k_{22}) [0,0.5 ]\,, \quad \tilde k_{32} \in [-0.5\,, 0.5]\,, \quad y_{32} \in - \text{sign}(y_{22}) [0\,, 0.5 ] \,. \label{eq:range_scanning} \end{align} In order to avoid the constraints from $\tau\to \ell \gamma$ ($\ell = e, \mu$) and get $|C^{LQ,\mu}_{9}| \sim 1$, we set ($k_{33}/k_{32}, \tilde k_{33}/\tilde k_{32}) \sim 0.1$ in Eq.~(\ref{eq:range_scanning}). Additionally, the negative value of $C^{LQ,\mu}_{9}$ can be achieved when $k_{32}(y_{32})$ and $k_{22}(y_{22})$ are opposite in sign. As mentioned earlier, the Yukawa couplings in both decays $\tau \to \ell \gamma$ and $h\to \ell \tau$ are the same, we cannot remove the constraints from the radiative LFV processes in this model. The BRs for $h\to \ell \tau$ thus are of ${\cal }O(10^{-9})$ and much smaller than the current upper limits of ${\cal O}(10^{-4})$~\cite{Khachatryan:2015kon,Aad:2015gha}. One way to escape the constraint from $\tau \to \ell \gamma$ is to add a new LQ~\cite{Baek:2015mea}. Since we focus on the excesses of muon $g-2$ and $B\to K^* \mu^+ \mu^-$, we leave the more complicated model for further study. With the chosen ranges of parameters in Eq.~(\ref{eq:range_scanning}), we first show the values for $C^{LQ,\mu}_9$ and $C^{LQ,\mu}_{10}$ in Fig.~\ref{fig:scatter}(a), where the bounds from $\tau \to \ell \gamma$ have been considered; the horizontal band is from the measurement of $B_s\to \mu^+ \mu^-$; the vertical band is the range that can explain the excess of $B\to K^* \mu^+ \mu^-$, and we used $10^{5}$ parameter sets and obtained $824$ allowed points that satisfy the constraints. It can be seen that the $C^{LQ,\mu}_9$ and $C^{LQ,\mu}_{10}$ from the contributions of doublet $\Phi_{7/6}$ and triplet $\Delta_{1/3}$ LQs can simultaneously satisfy the constraint of $B_s\to \mu^+ \mu^-$ and explain the excess in $B\to K^* \mu^+ \mu^-$. From Eq.~(\ref{eq:mug2}), it is known that muon $g-2$ is associated with the Yukawa couplings $k_{32}\tilde k_{32}$. Although only $k_{32}$ is related to $C^{LQ,\mu}_9$ and $C^{LQ,\mu}_{10}$, since the Yukawa couplings $k_{q\ell}$, $\tilde k_{q\ell}$, and $y_{q\ell}$ are taken to be the same order of magnitude, we present the correlations of $\Delta a_\mu$ and $C^{LQ,\mu}_{9}$ in Fig.~\ref{fig:scatter}(b), where only the allowed range of $C^{LQ,\mu}_{9}$ is shown, and the region between two dashed lines denotes the $\Delta a_\mu$ data with $1\sigma$ errors. By plot (b), it can be seen clearly that the excesses in $\Delta a_\mu$ and $B\to K^* \mu^+ \mu^-$ can be simultaneously fitted in the model. \begin{figure}[hptb] \includegraphics[width=70mm]{C9-C10.eps} \includegraphics[width=70mm]{C9-damu.eps} \caption{ (a) The values of $C^{LQ,\mu}_9$ and $C^{LQ,\mu}_{10}$ using the ranges of parameters in Eq.~(\ref{eq:range_scanning}), where the bound from the $\text{BR}(B_s\to \mu^+ \mu^-)$ and the allowed range of $C^{LQ,\mu}_9$ from global analysis for $B\to K^* \mu^+ \mu^-$ are shown. (b) Correlation of $\Delta a_\mu$ and $C^{LQ,\mu}_9$, where we only show the values of $C^{LQ,\mu}_9$ that can fit the excess in $B\to K^* \mu^+ \mu^-$, and the band bounded by two dashed lines denotes the $\Delta a_\mu$ data with $1 \sigma$ errors~\cite{PDG}. } \label{fig:scatter} \end{figure} As discussed before, in order to avoid the constraint from $\mu\to e \gamma$, we set $k_{31}=\tilde k_{31}=0$ in our analysis; therefore, $C^{LQ,e}_{9(10)}$ for $B\to K e^+ e^-$ decay is only related to $y_{31} y_{21}$. Since $y_{31,21}$ are free parameters, for simplicity, we then take $|y_{31}|\sim |k_{31}|\sim 0$. As a result, $X^{e}=C^{LQ,e}_9 - C^{LQ,e}_{10} \approx 0$. In order to see whether the obtained $C^{LQ,\mu}_9$ and $C^{LQ,\mu}_{10}$ can fit the $R_K$ data, we show the correlation between $X^{\mu}$ and $C^{LQ,\mu}_{9}$ in Fig.~\ref{fig:scatter2}, where the band denotes the allowed range shown in Eq.~(\ref{eq:X}). It can be seen that the excesses of $R_K$ and $P'_5$ can be simultaneously explained when the measurement of $B_s\to \mu^+ \mu^-$ is satisfied. \begin{figure}[hptb] \includegraphics[width=75mm]{C9-Rk.eps} \caption{ Correlation between $X^\mu = C^{LQ,\mu}_9 - C^{LQ,\mu}_{10}$ and $C^{LQ,\mu}_9$, where the allowed range of $X^\mu$ is from the $R_K$ data with $1\sigma$ errors. \label{fig:scatter2}} \end{figure} \section{ Conclusion} In order to resolve the excesses of muon $g-2$ and $B\to K^{(*)} \ell^+ \ell^-$ decays, we investigate the extension of the SM by including leptoquarks, in which the particles are colored scalar bosons and can couple to quarks and leptons. In order to accommodate the measurement of $B_s\to \mu^+ \mu^-$ and the excesses of $B\to K^{(*)} \mu^+ \mu^-$, we study a model with one doublet and one triplet leptoquarks. After SSB, the couplings of the SM Higgs the LQs are described by $\mu_{hX}=\lambda_{HX} v$. If $\mu_{hX}$ is of ${\cal O}(100)$ GeV, the signal strength parameter $\mu^{\gamma\gamma}_{i}$ can significantly deviate from the SM prediction and is still consistent with the current Higgs measurements. In this study, lepton-flavor violating processes $\ell_i \to \ell_j \gamma$ give strict constraints on the Yukawa couplings $k_{31,33}$ and $\tilde k_{31,33}$. As a result, the branching ratios for the lepton-flavor violating Higgs decays $h\to \ell_i \ell_j$ are less than ${\cal O}(10^{-8}$). Nevertheless, the sizable couplings $k_{32,22}$, $\tilde k_{32,22}$, and $y_{32,22}$ can still explain the excess of muon $g-2$ and provide the necessary values for the Wilson coefficient $C^{\ell}_9$, such that the excesses in $B\to K^* \mu^+ \mu^-$ and $R_K$ can be resolved. \section*{Acknowledgments} This work was partially supported by the Ministry of Science and Technology of Taiwan R.O.C., under grant MOST-103-2112-M-006-004-MY3 (CHC). H. O. would like to thank the members of KIAS for their hospitality during his visit.
895d95b0ec5e0786bee06a5d7211e97139ca0885
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction}\label{int} The study of hypernuclei has been attracting great interest of nuclear physics community in providing the information from nucleon-nucleon (NN) interaction to hyperon-nucleon (YN) and hyperon-hyperon (YY) interactions. Due to the injection of hyperons a new dimension is added in normal nuclear system and hyperons serve as a potential probe for exploring many nuclear properties in domain of strangeness~\cite{bando1990,tan2004,lu2011}. However, hyperon-nucleon interaction is weaker than nucleon-nucleon but it is imperative as well as important to describe the nuclear many-body system with strangeness. Various theoretical approaches Skyrme Hartree Fock~\cite{3,4,5,6,7,8,9,10,11}, relativistic mean field~\cite{14,15,16,17,18,19}, cluster, variational, diffusion Monte Carlo~\cite{Hiyama2010,Hiyama2012,Nemura2005,nemura2005,Gal2009,Gal2011,Usmani2006,Usmani2008,Zhou2008,Vidana2004,Samanta2008}, and G-matrix~\cite{25,26,27} have been employed by scientific community to estimate the strength of hyperon-nucleon as well hyperon-hyperon interactions. Further, these models have established themselves as very effective in testing the existence of bound hypernuclei and the stability of nucleonic core against hyperon(s) addition or the occurrence of exotic strange matter which facilitate the path toward multi-strange systems. Magic numbers in nuclear physics are certain neutron and proton numbers in atomic nuclei, in which higher stability in the ground state is observed than in the neighbouring nuclides and are most abundant in nature. The various experimental signatures that show discontinuity at magic numbers are: the energy required for the separation of one and two nucleons, the energies of alpha and beta transitions, pairing energy and the excitation of low-lying vibrational~\cite{goeppert1958,solovev1992,nilsson1995}. The separation energy is sensitive to the collective or single particle inter play and provides a sufficient information about the nuclear structure effects. The discovery of magic numbers paved the way to great progress in understanding of nuclear structure and these numbers became the cornerstones for future theoretical developments in nuclear physics. It is worthy to mention that the several signatures are seen for the evolution of the magic gaps along the nuclear chart including superheavy region~\cite{sorlin2008,patra2012,zhang2005}. The quest for proton or neutron magic numbers in the elusive mass region of superheavy nuclei is of utmost importance as the mere existence of superheavy nuclei is the result of the interplay between the attractive nuclear force (shell effects) and the large disruptive coulomb repulsion between the protons that favours the fission~\cite{strutinsky1967,strutinsky1998}. It is well established that 2, 8, 20, 28, 50, 82 and 126 are the nucleonic magic number. In addition to this, Z = 120 and N = 172, 184 are predicted to be next magic number by various theoretical models in superheavy mass region~\cite{patra1999,patra1997,patra2004,rutz1997,meng2007}. These predictions have been made on the basis of separation energy, shell gaps, pairing energy and shell correction energy etc. It may therefore relevant to extend the line of thought to the hypernuclear chart. It is well known that the spin-orbit interaction in $\Lambda$ channel is weaker than nucleonic sector and thus the $\Lambda$ magic numbers are expected to be close to the harmonic oscillator ones: 2, 8, 20, 40 and 70. In this paper, our main motive is to make an extensive investigation to search the $\Lambda$ magic number within the RMF approach and to obtain the stability of triply magic system with doubly magic core. The magic numbers in nuclei are characterized by a large shell gap in single-particle energy levels. This means that the nucleon in the lower level has a comparatively large value of energy than that on higher level giving rise to more stability. The extra stability corresponding to certain number can be estimated from the sudden fall in the separation energy. The $\Lambda$ separation energy is considered to be one of the key quantity to reveal the nuclear response to the addition of lambda hyperon. Therefore, in present work, we obtain the the binding energy per particle and one-lambda as well as two-lambda separation energies for considered multi-hypernuclei. Moreover, two-lambda shell gaps is also calculated to make the clear presentation of the magic number which also support to the two-lambda separation energy. To mark the $\Lambda$ shell gaps, single-particle energy levels are analyzed that may correspond to $\Lambda$ magic number. In addition, to analyze the structural distribution as well as impact of $\Lambda$ hyperon on bubble structure for considered nuclei, total (nucleon plus $\Lambda$) density is reported. Nucleon and lambda mean field and spin-orbit interaction potentials are also observed. On the basis of binding energy per particle, the stability of triply magic hypernuclei is reported. RMF theory has been quite successful for studying the infinite nuclear systems and finite nuclei including the superheavy mass region~\cite{patra1999,patra1997,patra2004,rutz1997,meng2007,serot1992,gambhir1990,ring1996,serot1986,boguta1977,jha2007,ikram2013}. It is quite successful to study the equation of state for infinite nuclear matter as well as pure neutron matter, where the existence of strange baryons is expected~\cite{glendenning1998,schaffner2002}. In this context, addition of strangeness degree of freedom to RMF formalism is obvious for the suitable expansion of the model and such type of attempts have already been made~\cite{schaffner1994,rufa1990,glendenning1993,schaffner1993,mares1994,sugahara1994,vretenar1998,schaffner2002,lu2003,shen2006,win2008,ikram14}. RMF explains not only the structural properties of singly strange hypernuclei but also provides the details study of multi-strange systems containing several $\Lambda$'s, $\Sigma$'s or $\Xi$'s. In fact, RMF explains spin-orbit interaction very nicely in normal nuclei as well as hypernuclei. The contribution of spin-orbit interaction is very crucial in emerging the magic number in nucleonic sector and the same is expected in strange sector. The paper is organized as follows: A brief introduction on hypernuclei and magic number is given in Section~\ref{int}. Section~\ref{form} gives a brief description of relativistic mean field formalism for hypernuclei with inclusion of $\Lambda N$ and $\Lambda \Lambda$ interactions. The results are presented and discussed in Section~\ref{resu}. The paper is summarized in Section~\ref{sum}. \section{Formalism }\label{form} Relativistic mean field theory has been applied succesfully to study the structural properties of normal nuclei as well as hypernuclei~\cite{rufa1990,glendenning1993,mares1994,sugahara1994,vretenar1998,lu2003,win2008,ikram14}. The suitable expansion to hypernuclei has been made by including the lambda-baryon interaction Lagrangian with effective $\Lambda$N potential. The total Lagrangian density for single-$\Lambda$ hypernuclei has been given in many Refs.~\cite{rufa1990,glendenning1993,mares1994,sugahara1994,vretenar1998,lu2003,win2008,ikram14}. For dealing the multi-lambda hypernuclei in quantitative way, the additional strange scalar ($\sigma^*$) and vector ($\phi$) mesons have been included which simulate the $\Lambda \Lambda$ interaction ~\cite{schaffner1994,schaffner1993,schaffner2002,shen2006}. Now, the total Lagrangian density can be written as \begin{eqnarray} \mathcal{L}&=&\mathcal{L}_N+\mathcal{L}_\Lambda+\mathcal{L}_{\Lambda\Lambda} \;, \end{eqnarray} \begin{eqnarray} {\cal L}_N&=&\bar{\psi_{i}}\{i\gamma^{\mu} \partial_{\mu}-M\}\psi_{i} +{\frac12}(\partial^{\mu}\sigma\partial_{\mu}\sigma -m_{\sigma}^{2}\sigma^{2}) -{\frac13}g_{2}\sigma^{3} \nonumber \\ &&-{\frac14}g_{3}\sigma^{4} -g_{s}\bar{\psi_{i}}\psi_{i}\sigma -{\frac14}\Omega^{\mu\nu}\Omega_{\mu\nu} +{\frac12}m_{\omega}^{2}\omega^{\mu}\omega_{\mu} \nonumber \\ &&-g_{\omega }\bar\psi_{i}\gamma^{\mu}\psi_{i}\omega_{\mu} -{\frac14}B^{\mu\nu}B_{\mu\nu} +{\frac12}m_{\rho}^{2}{\vec{\rho}^{\mu}}{\vec{\rho}_{\mu}} -{\frac14}F^{\mu\nu}F_{\mu\nu} \nonumber \\ &&-g_{\rho}\bar\psi_{i}\gamma^{\mu}\vec{\tau}\psi_{i}\vec{\rho^{\mu}} -e\bar\psi_{i}\gamma^{\mu}\frac{\left(1-\tau_{3i}\right)}{2}\psi_{i}A_{\mu}\;, \nonumber \\ \mathcal{L}_{\Lambda}&=&\bar\psi_\Lambda\{i\gamma^\mu\partial_\mu -m_\Lambda\}\psi_\Lambda -g_{\sigma\Lambda}\bar\psi_\Lambda\psi_\Lambda\sigma -g_{\omega\Lambda}\bar\psi_\Lambda\gamma^{\mu}\psi_\Lambda \omega_\mu \;, \nonumber \\ \mathcal{L}_{\Lambda\Lambda}&=&{\frac12}(\partial^{\mu}\sigma^*\partial_{\mu}\sigma^* -m_{\sigma^*}^{2}\sigma^{*{2}}) -{\frac14}S^{\mu\nu}S_{\mu\nu} +{\frac12}m_{\phi}^{2}\phi^{\mu}\phi_{\mu} \nonumber \\ &&-g_{\sigma^*\Lambda}\bar\psi_\Lambda\psi_\Lambda\sigma^* -g_{\phi\Lambda}\bar\psi_\Lambda\gamma^\mu\psi_\Lambda\phi_\mu \;, \end{eqnarray} where $\psi$ and $\psi_\Lambda$ denote the Dirac spinors for nucleon and $\Lambda$-hyperon, whose masses are M and $m_\Lambda$, respectively. Because of zero isospin, the $\Lambda$-hyperon does not couple to ${\rho}$- mesons. The quantities $m_{\sigma}$, $m_{\omega}$, $m_{\rho}$, $m_{\sigma^*}$, $m_{\phi}$ are the masses of $\sigma$, $\omega$, $\rho$, $\sigma^*$, $\phi$ mesons and $g_s$, $g_{\omega}$, $g_{\rho}$, $g_{\sigma\Lambda}$, $g_{\omega\Lambda}$, $g_{\sigma^*\Lambda}$, $g_{\phi\Lambda}$ are their coupling constants, respectively. The nonlinear self-interaction coupling of ${\sigma}$ mesons is denoted by $g_2$ and $g_3$. The total energy of the system is given by $E_{total} = E_{part}(N,\Lambda)+E_{\sigma}+E_{\omega}+E_{\rho} +E_{\sigma^*}+E_{\phi}+E_{c}+E_{pair}+E_{c.m.},$ where $E_{part}(N,\Lambda)$ is the sum of the single-particle energies of the nucleons (N) and hyperon ($\Lambda$). The energies parts $E_{\sigma}$, $E_{\omega}$, $E_{\rho}$, $E_{\sigma^*}$, $E_{\phi}$, $E_{c}$, $E_{pair}$ and $E_{cm}$ are the contributions of meson fields, Coulomb field, pairing energy and the center-of-mass energy, respectively. In present work, for meson-baryon coupling constant, NL3* parameter set is used through out the calculations~\cite{lalazissis09}. To find the numerical values of used $\Lambda-$meson coupling constants, we adopt the nucleon coupling to hyperon couplings ratio defined as; $R_\sigma=g_{\sigma\Lambda}/g_s$, $R_\omega=g_{\omega\Lambda}/g_\omega$, $R_{\sigma^*}=g_{\sigma^*\Lambda}/g_s$ and $R_\phi=g_{\phi\Lambda}/g_\omega$. The relative coupling values are used as $R_\omega=2/3$, $R_\phi=-\sqrt{2}/3$, $R_\sigma=0.621$ and $R_{\sigma^*}=0.69$~\cite{chiapparini09,schaffner1994,dover1984}. In present calculations, we use the constant gap BCS approximation to include the pairing interaction and the centre of mass correction is included by $E_{cm}=-(3/4)41A^{-1/3}$. \section{RESULTS AND DISCUSSIONS}\label{resu} Before taking a detour on searching the $\Lambda$ magic behaviour in multi-$\Lambda$ hypernuclei, first we see the effects of introduced $\Lambda$ hyperon on normal nuclear core; how the binding energy and radii of normal nuclear system is affected by addition of $\Lambda$'s ? To analyze this, we consider a list of normal nuclei covering a range from light to superheavy mass region i.e. $^{16}O$ to $^{378}120$. Total binding energy (BE), binding energy per particle (BE/A), lambda binding energy ($B_\Lambda$) for $s-$ and $p-$ state and radii for considered core nuclei and corresponding hypernuclei are tabulated in Table~\ref{tab1}. The calculated $B_\Lambda$ is compared with available experimental data and we found a close agreement between them. This means the used parameter set is quite efficient to reproduce the experimental binding energy and ofcourse we can use it to make more calculations related to magicity in hypernuclei. Since, we are dealing with closed shell hypernuclei and hence our RMF calculations is restricted to spherical symmetric. The addition of $\Lambda$ hyperon to normal nuclei enhances the binding and shrinks the core of the system. This is because of glue like role of $\Lambda$ hyperon that residing on the $s-$state for most of the time. These observations are shown in Table~\ref{tab1}. Binding energy of hypernuclei are larger than their normal counter parts and a reduction in total radius ($r_{total}$) of hypernuclei is observed, that means the $\Lambda$ particle makes the core compact with increasing binding. For example, the total radius of $^{16}O$ and $^{209}Pb$ is 2.541 and 5.624 fm, which reduce to 2.536 and 5.616 fm by addition of single $\Lambda$ into the core of $O$ and $Pb$, respectively. Moreover, for the sake of comparison with experimental data, binding energy and radii of the hypernuclei produced by replacing the neutrons means having a constant baryon number are also framed in Table~\ref{tab1} and the shrinkage effect is also noticed. This results show that an important impact of $\Lambda$ hyperon on binding as well as size of the system. The increasing value of $B_\Lambda$ for $s-$state from light to superheavy hypernuclei confirming the potential depth of $\Lambda-$particle in nuclear matter which would be -28~MeV ~\cite{schaffner1994,millener1988}. \begin{table*} \caption{The calculated total binding energy and binding energy per particle for single-$\Lambda$ hypernuclei and their normal counter parts are listed here. The $\Lambda$ binding energy for $s-$ and $p-$state of considered hypernuclei are also mentioned and compared with available experimental values~\cite{hashimoto2006}, are given in parentheses. The radii are also displayed. Energies are given in unit of MeV and radii are in fm.} \renewcommand{\tabcolsep}{0.20cm} \renewcommand{\arraystretch}{1.3} \begin{ruledtabular} \begin{tabular}{cccccccccc} Nuclei / Hypernuclei&BE &BE/A &$B_\Lambda^s$ &$B_\Lambda^p$ &$r_{ch}$ &$r_{total}$ &$r_p$ &$r_n$ &$r_\Lambda$ \\ \hline $^{16}$O &126.27 &7.89& & &2.674& 2.541& 2.555& 2.527&\\ $^{16}_\Lambda$O &124.36 &7.77&-12.09 (12.5$\pm$0.35)&-2.66 (2.5$\pm$0.5)&2.673& 2.487& 2.550& 2.428&2.388\\ $^{17}_\Lambda$O &137.99 &8.12&-11.98&-3.07 &2.673& 2.536& 2.554& 2.526&2.468\\ $^{40}$Ca &341.43 &8.54& & &3.446& 3.331& 3.355& 3.307&\\ $^{40}_\Lambda$Ca &344.16 &8.60&-17.51 (18.7$\pm$1.1)&-9.32 (11.0$\pm$0.6)&3.439& 3.292& 3.346& 3.262&2.693\\ $^{41}_\Lambda$Ca &358.65 &8.75&-17.39&-9.46 &3.444& 3.315& 3.352& 3.304&2.737\\ $^{48}$Ca &414.17 &8.63& & &3.444& 3.496& 3.359& 3.591&\\ $^{48}_\Lambda$Ca &425.04 &8.86&-18.75&-10.95&3.439& 3.459& 3.353& 3.558&2.785\\ $^{49}_\Lambda$Ca &433.01 &8.84&-18.94&-11.16&3.440& 3.479& 3.355& 3.586&2.791\\ $^{56}$Ni &482.30 &8.61& & &3.696& 3.586& 3.610& 3.561&\\ $^{56}_\Lambda$Ni &487.47 &8.71&-20.48&-12.67&3.695& 3.560& 3.609& 3.533&2.816\\ $^{57}_\Lambda$Ni &502.96 &8.82&-20.72&-12.89&3.689& 3.567& 3.603& 3.555&2.817\\ $^{90}$Zr &783.17 &8.70& & &4.249& 4.245& 4.179& 4.297&\\ $^{90}_\Lambda$Zr &792.04 &8.80&-21.28&-15.25&4.246& 4.221& 4.175& 4.277&3.215\\ $^{91}_\Lambda$Zr &804.69 &8.84&-21.37&-15.36&4.247& 4.233& 4.177& 4.295&3.222\\ $^{124}$Sn &1048.19&8.45& & &4.642& 4.753& 4.580& 4.866&\\ $^{124}_\Lambda$Sn &1060.81&8.55&-22.24&-17.10&4.633& 4.729& 4.571& 4.849&3.493\\ $^{125}_\Lambda$Sn &1070.73&8.57&-22.28&-17.17&4.638& 4.742& 4.577& 4.864&3.503\\ $^{132}$Sn &1102.69&8.35& & &4.689& 4.854& 4.631& 4.985&\\ $^{132}_\Lambda$Sn &1118.13&8.47&-22.56&-17.65&4.680& 4.830& 4.620& 4.969&3.570\\ $^{133}_\Lambda$Sn &1125.49&8.46&-22.60&-17.71&4.686& 4.843& 4.627& 4.983&3.579\\ $^{208}$Pb &1639.32&7.88& & &5.499& 5.624& 5.448& 5.736&\\ $^{208}_\Lambda$Pb &1655.85&7.78&-23.56 (26.3$\pm$08)&-19.74 (21.3$\pm$0.7)&5.492& 5.604& 5.441& 5.719&4.067\\ $^{209}_\Lambda$Pb &1663.47&7.96&-23.54&-19.76&5.496& 5.616& 5.445& 5.734&4.076\\ $^{292}$120 &2063.09&7.06& & &6.271& 6.322& 6.225& 6.389&\\ $^{292}_\Lambda$120 &2103.33&7.12&-23.73&-20.94&6.262& 6.306& 6.216& 6.376&3.316\\ $^{293}_\Lambda$120 &2111.67&7.21&-23.59&-20.86&6.268& 6.316& 6.223& 6.387&3.320\\ $^{304}$120 &2140.81&7.04& & &6.302& 6.417& 6.258& 6.519&\\ $^{304}_\Lambda$120 &2184.09&7.19&-23.89&-21.03&6.298& 6.403& 6.253& 6.507&3.272\\ $^{305}_\Lambda$120 &2189.18&7.18&-23.79&-20.96&6.300& 6.411& 6.255& 6.518&3.301\\ $^{378}$120 &2385.44&6.31& & &6.714& 7.144& 6.678& 7.350&\\ $^{378}_\Lambda$120 &2432.02&6.11&-23.09&-20.81&6.896& 7.344& 6.863& 7.543&3.642\\ $^{379}_\Lambda$120 &2433.89&6.09&-23.43&-21.04&6.712& 7.138& 6.676& 7.350&3.549\\ \end{tabular} \end{ruledtabular} \label{tab1} \end{table*} \begin{figure} \includegraphics[width=1.0\columnwidth]{beperb.pdf} \caption{\label{be_perb}(color online) Binding energy per particle for triply magic multi-$\Lambda$ hypernuclei.} \end{figure} \subsection{Stability of hypernuclei} Binding energy provides the detailed information of various elements corresponding to their stability. Binding energy per particle increases upto the element iron whose atomic number is 26 and mass number 57. The information provided by the binding energy per particle curve is that iron and its neighbouring elements (Ni) are most stable i.e. they neither undergo fission or fusion. Thus the significance of the binding energy per particle curve lies in the fact that it is an indicator of nuclear stability and thus helps in classifying the elements which undergo fission, fusion and radioactive disintegration. We noticed a similar pattern of binding energy per particle in hypernuclear regime also. The triply magic system is produced by addition of $\Lambda$ magic number into the core of doubly shell closure such as $^{16}O$,$^{48}Ca$,$^{58}Ni$,$^{90}Zr$,$^{124}Sn$,$^{132}Sn$,$^{208}Pb$, $^{292}120$,$^{304}120$, $^{378}120$. Binding energy per particle of considered systems confirming that nickel with 8 $\Lambda$'s ($^{56+8\Lambda}Ni$) being the most tightly bound hypernucleus (BE/A = 9.5 MeV) as shown in Fig.~\ref{be_perb}. This results are in remarkable agreement with earlier predictions~\cite{schaffner1994, schaffner1993}. \begin{figure} \includegraphics[width=1.0\columnwidth]{lambda-be.pdf} \caption{\label{total_be}Binding energy per particle for light to supeheavy mass multi-$\Lambda$ hypernuclei.} \end{figure} \subsection{Binding energy and separation energy} To analyze the magic behaviour of lambda in multi-$\Lambda$ hypernuclei, we choose the nuclear core of doubly shell closure including predicted shell closure nuclei of superheavy mass region and than added the $\Lambda$ hyperons. Initially, we look for the binding energy per particle with respect to added hyperons for considered hypernuclei and plotted in Fig.~\ref{total_be}. The peak value of the graph corresponds to maximum stability for a particular hypernuclear system. It also explains that the injection of few $\Lambda$ hyperons enhance the binding of the light mass hypernuclei, conversely further addition reduces the binding. However for heavy mass region, the BE/A increases with addition of large number of hyperons and form a most bound system but further addition decreases the binding energy gradually. This means certain number of added $\Lambda$'s to a particular nuclear core form a most stable system. For example, the injection of 2 $\Lambda$'s provide a maximum stability to $^{16}O$. Along the similar line a maximum binding is observed for $^{48}Ca$ with $\Lambda=$ 8 and this number goes to 90 for superheavy core. In this way, we extract certain numbers of added $\Lambda$'s that is 2, 8, 18, 20, 40, 70, 90 which provide maximum stability for the considered systems (i.e. $^{16+2\Lambda}O$, $^{48+8\Lambda}Ca$, $^{58+8\Lambda}Ni$, $^{90+18\Lambda}Zr$, $^{124+20\Lambda}Sn$, $^{132+20\Lambda}Sn$, $^{208+40\Lambda}Pb$, $^{292+68\Lambda}120$, $^{304+70\Lambda}120$, $^{378+90\Lambda}120$) and these numbers may correspond to $\Lambda$ magic number in multi-$\Lambda$ hypernuclei. But, there exist several other strong signatures of marking the magic number such as; separation energy, shell gaps, pairing energy etc. Therefore, to analyze the actual behaviour of magicity, we make the analysis of such relevant parameters. In this regard, we estimate one- and two-lambda separation energy $S_\Lambda$, $S_{2\Lambda}$, which are known to be first insight of shell closure. In analogy of nucleonic sector, the magic number in multi-hypernuclei may characterized by the large lambda shell gaps in single-particle energy levels. The extra stability given by certain number of introduced $\Lambda$'s can also be detect from sudden fall in $\Lambda$ separation energy. Therefore, on quest of magicity in multi-$\Lambda$ hypernuclei, one- and two-lambda separation energy is estimated using the following expressions; \begin{equation*} S_\Lambda(N,Z,\Lambda)=BE(N,Z,\Lambda)-BE(N,Z,\Lambda-1) \end{equation*} and \begin{equation*} S_{2\Lambda}(N,Z,\Lambda)=BE(N,Z,\Lambda)-BE(N,Z,\Lambda-2). \end{equation*} These quantities are plotted in Fig.~\ref{s1-light},~\ref{s1-heavy},~\ref{s2-light} and~\ref{s2-heavy}. For a lambda chain, the $S_\Lambda$ and $S_{2\Lambda}$ becomes larger with increasing number of lambda $\Lambda$. For a fixed Z, N; $S_\Lambda$ and $S_{2\Lambda}$ decrease gradually with lambda number. A sudden decrease of $S_\Lambda$ and $S_{2\Lambda}$ just after the magic number in lambda chain like as neutron chain indicates the occurrence of $\Lambda$ shell closure. The sudden fall of $S_\Lambda$ at $\Lambda=$ 2, 8, 14, 18, 20, 28, 34, 40, 50, 58, 68, 70 and 82 can clearly be seen in considered hypernuclear candidates revealing a signature of magic character. Moreover, $\Lambda=$ 14 and 28 are observed only in light mass mulit-$\Lambda$ hypernuclei, even $\Lambda=$ 28 does not show pronounced energy separation. However, a good strength of sudden fall of $S_\Lambda$ at $\Lambda=$ 34 and 58 is clearly observed in heavy and superheavy mass region. Two-lambda separation energy provides more strong signature to quantify shell closure due to absence of odd-even effects. Figures~\ref{s2-light} and \ref{s2-heavy} reveal that sudden fall of $S_{2\Lambda}$ at $\Lambda=$ 2, 8, 14, 18, 20, 28, 34, 40, 50, 58, 68, 70 and 82 as observed in considered multi-$\Lambda$ hypernuclear candidates. These certain numbers corresponds to $\Lambda$ magic number in multi-$\Lambda$ hypernuclei and form a triply magic system with doubly magic core. This is the central theme of the paper. The significant falls of $S_\Lambda$ and $S_{2\Lambda}$ at $\Lambda=$14 is appeared in Ca and Ni hypernuclei. The lambda number 28 seems to be very much feeble magic number, contrary to nucleonic sector. The another new lambda number 68 suppose to be semi-magic arises due to subshells closure. For the sake of clear presentation of the results, we also make the analysis for two-lambda shell gaps ($\delta_{2\Lambda}$) and plotted as a function of added $\Lambda$'s. Summarizing the above results, we may say that based on one- and two-lambda separation energies $S_\Lambda$ and $S_{2\Lambda}$, the signatures of the magicity in RMF appears at 2, 8, 14, 18, 20, 28, 34, 40, 50, 58, 68, 70 and 82. The lambda number 28 and 68 are appeared in light and heavy hypernuclei, respectively and suppose to be feeble magic number. \begin{figure} \includegraphics[width=1.0\columnwidth]{s1.pdf} \caption{\label{s1-light}One lambda separation energy for medium mass multi-$\Lambda$ hypernuclei.} \end{figure} \begin{figure} \includegraphics[width=1.0\columnwidth]{sep1.pdf} \caption{\label{s1-heavy}Same as Fig.~\ref{s1-light} but for heavy to superheavy mass multi-$\Lambda$ hypernuclei.} \end{figure} \begin{figure} \includegraphics[width=1.0\columnwidth]{s2.pdf} \caption{\label{s2-light}Two lambda separation energy for medium mass multi-$\Lambda$ hypernuclei.} \end{figure} \begin{figure} \includegraphics[width=1.0\columnwidth]{sep2.pdf} \caption{\label{s2-heavy}Same as Fig.~\ref{s2-light} but for heavy to superheavy mass multi-$\Lambda$ hypernuclei. } \end{figure} \subsection{Two-lambda shell gap} The change of the two-lambda separation energies can also be quantified by the second difference of the binding energies, i.e., two-lambda shall gap which is expressed by: \begin{small} \begin{eqnarray*} \delta_{2\Lambda}(N,Z,\Lambda)&=&2BE(N,Z,\Lambda)-BE(N,Z,\Lambda+2)-BE(N,Z,\Lambda-2) \\ &=&S_{2\Lambda}(N,Z,\Lambda)-S_{2\Lambda}(N,Z,\Lambda+2). \end{eqnarray*} \end{small} A peak of two-lambda shell gaps indicates the drastic change of the two-lambda separation energies; which is used as one of the significant signature of magic number. The two-lambda shell gaps $\delta_{2\Lambda}$, for all considered hypernuclei as a function of added $\Lambda$ hyperons are shown in Fig~\ref{s-gaps}. A peak at certain $\Lambda$ number suggests the existence of lambda shell closure. However, the quality of magic number is represented by sharpness as well as magnitude of the peak. Figure.~\ref{s-gaps} reveals that the magnitude of the peak is found to be largest at $\Lambda=$ 2, 8, 20, 40 indicating the strong shell closures. Further, the peaks appeared at $\Lambda=$ 14, 18, 28, 34, 50, 58, 70 and 82 indicate the respective lambda magic number. Moreover, a peak with a very small magnitude is also appeared at $\Lambda=$ 68 due to closureness of subshell (2$d_{3/2}$) revealing $\Lambda$ semi-magic number. A peak with small magnitude is seen at $\Lambda$= 28 representing a feeble lambda magic number, contrary to nucleonic magic number. Pronounced peak is appeared at $\Lambda=$ 34 and 58 indicating a strong $\Lambda$ closed shells. \begin{figure} \includegraphics[width=1.0\columnwidth]{shell-gap.pdf} \caption{\label{s-gaps}(color online) Two lambda shell gap is shown for considered multi-$\Lambda$ hypernuclei. } \end{figure} \subsection{Density profile and bubble structure} A hypernucleus is a composed system of nucleons and hyperons and hence the gross structure of hypernucleus can be described by density distribution of nucleons as well as hyperons. It is well known and has mentioned earlier that the addition of a $\Lambda$ hyperon makes the nuclear core compact with increasing binding as well as density. Therefore, it is important to study the effects of large number of added $\Lambda$ hyperons on the nuclear density. Due to addition of hyperons, the magnitude of total density increases with increasing number of $\Lambda$'s as shown in Fig.~\ref{density}. On view the density profile, one can examine the most interesting feature of nuclei i.e. bubble structure, which measure the depression of central density and has already been observed in light to superheavy mass region~\cite{grasso2009,ekhan08,decharge03,ikram14}. It is to be noticed that several factors, including pairing correlations ~\cite{grasso2009}, tensor force~\cite{wang2013,nakada2013} and dynamic shape fluctuations~\cite{yao2012,yao2013,wu2014} turn out to have influence on depression of central density. The exotic structures like bubble and halo have been recently studied in $\Lambda-$hypernuclei~\cite{ikram14,ekhan2015}. Owing to weaker $\Lambda \Lambda$ attraction compared to the nucleon-nucleon one the lambda hyperons are more diffuse in a nucleus than nucleons and thus generating a hyperon density about $1/3$ smaller than the nucleonic density. Thus, it becomes quite important to look for the effect of large number of hyperons on neutron and proton density distributions. Since, there is no change of nucleon number and hence no anomalous effect of introduced $\Lambda$'s on neutron, proton densities is observed, individually. But the total density of the system is largely affected due to increasing number of $\Lambda$'s into the core. In considered multi-hypernuclei, the nucleonic core of some of them shows the depression of central density for example $^{16}O$, $^{90}Zr$, $^{292}120$, $^{304}120$ and $^{378}120$ as predicted earlier also~\cite{decharge03,ikram14}. It is found that the injected $\Lambda$'s reduce the depression of central density. For example, the depression of central part in $^{16}O$ is reduced by injection of 2 $\Lambda$'s and further more by 8 $\Lambda$'s. The $\Lambda$ particle attracts the nucleons towards the centre enhancing the central density and as a result remove the bubble structure partially or fully as reflected in Fig.~\ref{density}. Therefore, it is one of the important implication of $\Lambda$ particle to the nuclear system. Beyond the bubble structure, no anomalous behaviour of total density (core + $\Lambda$) in triply magic system is reported. \begin{figure} \includegraphics[width=1.0\columnwidth]{total-den.pdf} \caption{\label{density}(color online) Total (Nucleon plus lambda) density for considered triply magic hypernuclei with $\Lambda$=0, 2, 8, 14, 20, 34, 40, 58, 70 and 82.} \end{figure} \subsection{Spin-orbit interaction and mean field potentials} The spin-orbit interaction plays a significant role in reproducing the results quantitatively. It is the beauty of RMF in which the spin-orbit splitting is built-in naturally with exchange of scalar and vector mesons and thus describe a nuclear fine structure. It is not limited only to nuclei or superheavy nuclei but appears in hypernuclei also, however the strength of interaction is weaker than normal nuclei~\cite{boguta1981,noble1980,vretenar1998}. It is clearly seen from Figs.~\ref{potentials} that the spin-orbit potential for lambda hyperon is weaker than their normal counter parts and our results are consistent with theoretical predictions and experimental measurements~\cite{brockmann1977,keil2000,ajimura2001}. Here, nucleon ($V^N_{so}$) and lambda ($V^\Lambda_{so}$) spin-orbit interaction potentials are calculated for considered triply magic multi-hypernuclei. It is also conclude that the addition of $\Lambda$'s affects the nucleon as well as $\Lambda$ spin-orbit potential to a great extent. The nucleon ($V_N$) and lambda ($V_\Lambda$) mean field potentials are also investigated and plotted as a function of radial parameter shown in Fig.~\ref{meanpot}. The total depth of $\Lambda$ mean field potential is found to be around -30 MeV, which is in agreement with existing experimental data~\cite{keil2000}. It is to mention that the additions of $\Lambda$'s affects the depth of both nucleon as well as lambda mean potentials. The nucleonic potential depth in multi-lambda hypernuclei is approx to -80 MeV to -90 MeV. The shape of lambda potential looks like to be same as nucleonic potential and only the amount of depth is different. It is also to be noticed that the nucleonic potential looks like as V-shape type and shows the maximum depth around -90 MeV at r = 4 fm, while this amount of depth reaches to -70 MeV at r = 0 fm for $^{292}$120. It indicates a relatively low concentration of the particles at central region (r = 0) which is a direct consequence of depression of central density so-called bubble structure. \begin{figure} \includegraphics[width=1\columnwidth,height=8cm]{spinorbit_pot.pdf} \caption{\label{potentials}(color online) Spin-orbit interaction potentials of nucleon and lambda for considered triply magic hypernuclei with $\Lambda$=0, 2, 8, 14, 20, 34, 40, 58, 70 and 82. The upper part in each panel represents the nucleonic spin-orbit and the lower one representing the lambda spin-orbit interaction.} \end{figure} \begin{figure} \includegraphics[width=1.0\columnwidth]{mean_pot.pdf} \caption{\label{meanpot}(color online) Total mean field potentials for considered triply magic hypernuclei with $\Lambda$=0, 2, 8, 14, 20, 34, 40, 58, 70 and 82. Dashed lines represents the lambda mean potential and the nucleon potential is shown by solid line.} \end{figure} \subsection{Single-particle energies} Any kinds of change in a system can be observed from their single-particle energy levels. To analyze the impact of $\Lambda$ hyperon on nucleon single-particle energy levels, the filled neutron and proton levels for Ca hypernuclei are plotted as a function of added hyperons as shown in Fig.~\ref{levels-ca}. Figure~\ref{levels-ca} reveals that the neutron and proton energy levels goes dipper with addition of $\Lambda$'s as a result increase the stability of the system. The added hyperons increase the nucleon separation energy and as a result form a more bound system with increasing binding then their normal counter parts, which also leads to an extension of drip-line~\cite{Samanta2008}. For example, neutron $s_{1/2}$(n) level has an energy of about -54.037 MeV for the core of Ca hypernucleus, while this amount reaches to -62.499 MeV for $^{48+18\Lambda}Ca$ system with 18 $\Lambda$'s. Also, a same trend is observed for proton levels where, $s_{1/2}$(p) has an energy 51.787 MeV for the core of Ca hypernuclei and this value reaches to -60.0477 MeV with addition of 18 $\Lambda$'s. This results show that the $\Lambda$ hyperons draw the nuclear system towards more stability with increasing strangeness. Moreover, the same trend of neutron and proton energy levels is observed for other multi hypernuclei where both the levels would go dipper with increasing number of $\Lambda$ hyperon to nucleonic core but we do not make a plot for the same. A inversion of proton levels is seen, where $d_{3/2}$ fill faster than $s_{1/2}$ and this type of filling is also observed in lambda levels. Further, we analyze the lambda single-particle energy levels for $^{48+n\Lambda}Ca$, $^{208+n\Lambda}Pb$ and $^{304+n\Lambda}120$ hypernuclei to extract the lambda shell gaps for confirming the $\Lambda$ magic number. The lambda energy levels as function of added $\Lambda$'s are given in Fig.~\ref{levels-ca},~\ref{levels-pb},~\ref{levels-120}. The filling of $\Lambda$'s is same as the nucleons following the shell model scheme with lambda spin-orbit interaction potential. It is observed that the single-particle gap of spin-orbit splitting in lambda levels is smaller than the nucleons due to weakening strength of lambda spin-orbit interaction. By analyzing the lambda single-particle energy levels of Ca hypernuclei it is found that large energy gap exist in $1d_{5/2}$ to $1d_{3/2}$ or $2s_{1/2}$ and that's why lambda magic number 14 is emerged. Further, $2s_{1/2}$ and $1d_{3/2}$ are very much close to each other due to weaker strength of $\Lambda$ spin-orbit interaction. However, $\Lambda=$ 20 is clearly seen due to large energy gap by $f_{7/2}$ to lower orbital. In case of Pb, the large shell gaps for $\Lambda =$ 2, 8, 18, 20, 28, 34, 40, 50, 58, 70 and 82 is appeared. However, the single-particle gap for lambda number 28 is not so strong as compared to others suggesting the feeble magic number. The inversion of normal level scheme is noticed and the higher levels fill faster than lower one and hence this types of filling is responsible to emerge the new more magic number. For example, the filling of $1d_{3/2}$ before $2s_{1/2}$ shows a shell gap at $\Lambda=$ 18. Along the similar line, due to inversion between \{$1f_{5/2}$, $2p_{3/2}$\} and \{$1g_{7/2}$, $2d_{5/2}$\} the $\Lambda$ closed shells 34 and 58 is observed, respectively. In case of superheavy multihypernuclei, large single-particle shell gaps are appeared for lambda number 2, 8, 18, 20, 34, 40, 50, 58, 70 and 82. It is quite worth to notice that pronounced energy gaps is noticed in $^{208}Pb$ and $^{304}120$ at $\Lambda=$ 34, 58 are being suggested to be strong $\Lambda$ shell closure. The sharp peaks observed in $\delta_{2\Lambda}$ at $\Lambda=$ 2, 8, 20, 34, 40 and 58 is clearly reflected from lambda single-particle energies, where a large energy gap is exist for the filling of these number of lambda hyperons. \begin{figure} \includegraphics[width=1.0\columnwidth]{canpl.pdf} \caption{\label{levels-ca}(color online) Single-particle energy levels for triply magic Ca multi-hypernuclei for $\Lambda=$ 2, 8, 14, 18, 20 and 28.} \end{figure} \begin{figure} \includegraphics[width=1.0\columnwidth]{pbnpl.pdf} \caption{\label{levels-pb}(color online) Single-particle energy levels for triply magic Pb multi-hypernuclei for $\Lambda$ = 2, 8, 18, 20, 28, 34, 40, 50, 58, 70, 82 and 90.} \end{figure} \begin{figure} \includegraphics[width=1.0\columnwidth]{304npl.pdf} \caption{\label{levels-120}(color online) Single-particle energy levels for triply magic $^{304}$120 multi-hypernuclei for $\Lambda$ = 2, 8, 18, 20, 28, 34, 40, 50, 58, 70, 82 and 90.} \end{figure} \begin{table} \caption{Lambda magic number produced in various considered multi-hypernuclei are tabulated here.} \begin{ruledtabular} \begin{tabular}{cccccccccccccc} \multicolumn{1}{c}{Hypernuclei}&\multicolumn{13}{c}{Lambda magic number}\\ \cline{1-1}\cline{2-14} \hline $^{16+n\Lambda}$O &2&8&-&-&-&-&-&-&-&-&-&-&-\\ $^{48+n\Lambda}$Ca &2&8&14&-&20&-&-&-&-&-&-&-&-\\ $^{58+n\Lambda}$Ni &2&8&14&-&20&28&34&-&-&-&-&-&-\\ $^{90+n\Lambda}$Zr &2&8&14&-&20&28&-&40&50&-&-&-&-\\ $^{124+n\Lambda}$Sn &2&8&14&-&20&28&34&40&50&-&-&-&-\\ $^{132+n\Lambda}$Sn &2&8&14&-&20&28&34&40&50&-&-&-&-\\ $^{208+n\Lambda}$Pb &2&8&-&18&20&28&34&40&50&58&-&70&82\\ $^{292+n\Lambda}$120&2&8&-&18&-&28&34&40&50&58&68&-&82\\ $^{304+n\Lambda}$120&2&8&-&18&20&28&34&40&50&58&68&70&82\\ $^{378+n\Lambda}$120&2&8&14&18&20&-&34&40&50&58&68&-&82\\ \end{tabular} \end{ruledtabular} \label{tab2} \end{table} \subsection{Magicity} Various signatures of the evolution of magic shell gaps have been discovered across the nuclear landscape during the past few decades~\cite{zhang2005,patra2012} such as (i) A large binding energy then neighbouring nuclides, (ii) Sudden fall at separation energy, (iii) A large shell gap, etc. It becomes therefore quite relevant to extend the prediction of magic numbers to the hypernuclear chart. Looking for the magic behaviour firstly, we emphasize on binding of some selected nuclei whose nucleonic core is doubly magic such as $O$, $Ca$, $Ni$, $Zr$, $Sn$, $Pb$ and $120$. Some certain number of $\Lambda$ hyperon binds the nuclear core with maximum stability that may correspond to $\Lambda$ magic number in hypernuclei and might form a triple magic system with doubly magic nuclear core as initially discussed in Ref.~\cite{zofka1989} and recently in Ref.~\cite{ekhan2015}. For example, $\Lambda=$ 2, produce a maximum binding for $^{16}O$ and $^{48}Ca$ reveal the maximum binding with $\Lambda=$ 8. Also, $^{378}120$ shows a peak binding on addition of 90 $\Lambda$'s. In this way, we extract some lambda number which are 2, 8, 18, 20, 40, 70, 90 produce a maximum stability for their particular system and suppose to be $\Lambda$ magic number as much as close to harmonic oscillator number. But several other strong signatures are exist to identify the magicity and we make the analysis in this direction to look out the correct $\Lambda$ magic number. After analyzing the $S_\Lambda$ and $S_{2\Lambda}$, for considered light to superheavy mass multi-hypernuclei it is noticed that a sudden fall is observed at $\Lambda=$ 2, 8, 14, 20, 28, 34, 40, 50, 58, 68, 70 and 82. And hence, the analysis suggest that these numbers are supposed to be $\Lambda$ magic number in multi-lambda hypernuclei and form a triply magic system having doubly magic nucleonic core. In order to identify the $\Lambda$ magic number strongly, two-lambda shell gaps is examined which provide more strong signature of magicity and favoured $S_{2\Lambda}$. Pronounced peak in two-lambda shell gap is observed at $\Lambda=$ 2, 8, 20 and 40 indicating a strong shell closure. The peaks observed with a significant magnitude at $\Lambda=$ 14, 18, 34, 50, 58, 70 and 82 indicating a shell closure also. Further to testify, we look for the lambda shell gaps by examining the single-particle energy levels. We noticed that a large single-particle gap is appeared in $^{208}Pb$ hypernucleus at $\Lambda=$ 2, 8, 18, 20, 28, 34, 40, 50, 58 and 70 confirming the lambda magic number. The analysis of single-particle energy levels for $^{378}120$ multi hypernuclei clear the results by showing a large energy gap in 2, 8, 18, 20, 34, 40, 50, 58, 70 and 82. It is to mention that a significant shell gap is observed for 34 and 58 suggesting a strong $\Lambda$ shell closure. The inversion of normal level scheme is responsible to emerge the $\Lambda$ magic number 34 and 58. The experimentally confirmation of nucleonic shell closure of 34 supports our predictions~\cite{stepp2013,maierbeck2009}. The nucleonic number 14, 16, 18 and 32 have also been in discussion and expected to be shell closure~\cite{jia2016,ozawa2000,kanungo2002,rodriguez2007}. In addition, nucleon number 16 and 32 have also been experimentally confirmed in exotic nuclei as a new neutron magic number~\cite{kanungo2009,wienholtz2013}. Therefore, it is concluded that the $\Lambda$ magicity quite resembles with the nuclear magicity and it is expected that our predictions might be used as significant input to make the things clear regarding new sub shell closure. The predicted $\Lambda$ magic number in multi-hypernuclei are framed in Table~\ref{tab2}. The present lambda magic numbers are quite agreeable with the prediction of Ref.~\cite{zofka1989} and Ref.~\cite{ekhan2015} where Bruckner-Hartee Fock calculations using the lambda density functional have been made. It is clear from the plot of $\delta_{2\Lambda}$, that 34, 58 and 70 have peaks of great magnitude, while 68 in a feeble magnitude and suppose to be subshell closure. Moreover, strong nucleonic magic number 28 is observed very feeble in lambda magicity. It is to mention that the lambda number 14 is appeared in medium mass hypernuclei, on contrary to this 18 is observed in superheavy mass multi hypernuclei. \section{Summary and conclusion}\label{sum} In summary, we have suggested the possible $\Lambda$ magic number i.e. 2, 8, 14, 18, 20, 28, 34, 40, 50, 58, 68, 70, 82 in multi-$\Lambda$ hypernuclei within the relativistic mean field theory with effective $\Lambda$N as well as $\Lambda$$\Lambda$ interactions. The survey of $\Lambda$ magic number is made on the basis of binding energy, one- and two-lambda separation energies $S_\Lambda$, $S_{2\Lambda}$, and two-lambda shell gaps $\delta_{2\Lambda}$. It is noticed that pronounced single particle energy gap is observed for lambda number 34 and 58 in Pb and superheavy multi hypernuclei representing the strong $\Lambda$ magic number. Our predictions are strongly supported by nuclear magicity, where nucleon number 34 is experimentally confirmed as a neutron shell closure~\cite{stepp2013,maierbeck2009}. It is expected that the weakening strength of lambda spin-orbit interaction potential is responsible for emerging the new lambda magic number. The predicted $\Lambda$ magic numbers are in remarkable agreement with earlier predictions~\cite{ekhan2015,zofka1989} and hypernucler magicity quite resembles with nuclear magicity. In the analogy of nuclear stability, we noticed a similar pattern of binding energy per particle in hypernuclear regime and Ni hypernucleus with 8 $\Lambda$'s is found to be most tightly bound triply magic system in hypernuclear landscape. The addition of $\Lambda$ hyperons have significant impact on nucleon distribution and remove the bubble structure partially or fully. The spin-orbit interaction potentials and mean field potentials is also studied for predicted triply magic hypernuclear systems and the added $\Lambda$'s affect both the potential to a large extent. The present results may be used as a significant input to produce the triply magic hypernuclei in laboratory in future. It is also concluded that the addition of $\Lambda$ hyperons draw the nuclear system towards more stability with increasing strangeness. We noticed that the core of superheavy nuclei has more affinity to absorb large number of hyperons. This means such systems are able to simulate the strange hadronic matter containing large number of heavy hyperons such as $\Sigma$'s and $\Xi$'s including several $\Lambda$'s and the formation of such systems has large implication in nuclear-astrophysics. \section{Acknowledgments} One of the authors (MI) acknowledge the hospitality provided by Institute of Physics, Bhubaneswar during the work.
f2c93076cabe2117c5b3eeb29231e2822423ac8a
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In the last decade, many studies have permitted a deeper understanding of the relationship between gravity and gauge theories from the point of view of scattering amplitudes (see \cite{review} for a comprehensive review). One of the most interesting results is color-kinematics duality \cite{color-kinematics}, which allows the construction of gravity amplitudes by replacing color factors by a second copy of the kinematic numerators. This double copy structure has a historic antecedent in the Kawai, Lewellen, and Tye (KLT) relations \cite{KLT}, showing how, at tree level, closed string amplitudes admit a decomposition in terms of products of open string amplitudes. Similar structures have been found in various other setups \cite{stieberger_taylor,ours}. It seems clear that, at the level of scattering amplitudes, there is a sense in which gravity can be considered the ``square'' of a gauge theory. Given this double copy structure, a natural question to ask is how certain properties of gauge theory amplitudes translate into the gravitational side. One of these is the existence of radiation zeros \cite{rad_zeros_rev}. This is a peculiar feature of certain scattering processes where one or more massless gauge bosons are radiated, consisting in the vanishing of the amplitude for certain phase space configurations. The phenomenon was first identified in processes involving gauge bosons trilinear couplings, in particular $u\bar{d}\rightarrow W^{+}\gamma$ \cite{rad-zeros}. It has been experimentally observed both at the Tevatron \cite{tevatron_zeros} and LHC \cite{LHC_zeros}. Their existence has been also studied in graviton photoproduction \cite{passarino}. These so-called Type-I zeros appear for momentum configurations satisfying the constraint $Q_{i}=\kappa \, p_{i}\cdot k$, where $k$ is the momentum of the gauge boson, $Q_{i}$ and $p_{i}$ are the charge and momenta of the other particles, and $\kappa$ is a numerical constant. In Ref. \cite{heyssler_stirling} it was realized that zeros in the amplitude may also occur when the spatial momenta of the particles involved in the process lie on the same plane. These Type-II or planar zeros have been identified in the processes $e^{+}e^{-}\rightarrow W^{+}W^{-}\gamma$ \cite{soft1} and $e^{+}e^{-}\rightarrow \tau\bar{\tau}\gamma$ \cite{soft2}, in both cases in the soft photon limit. So far, the only study of planar zeros beyond the soft limit has been carried out in the interesting work \cite{harland-lang}, where the five parton amplitude in QCD was analyzed. Using the maximally helicity violating (MHV) formalism, planar zeros were found both for the $gg\rightarrow ggg$ and $q\bar{q}\rightarrow ggg$ processes. In the case of the five-gluon amplitude for general color factors the planar zero condition depends on the color quantum numbers of incoming and outgoing gluons. The present article has a double aim. One is to study the conditions for the emergence of planar zeros in the five-gluon amplitude. We show that planar zeros are a ``projective" property of the amplitude, in the sense that they are preserved by a simultaneous rescaling of the stereographic coordinates labeling the flight directions of the three outgoing gluons. In terms of stereographic coordinates, we find that the existence of planar zeros is determined by a cubic algebraic curve whose integer coefficients are given in terms of the color factors. In the case of SU($N$) gauge groups, we find that the casuistic of curves obtained for different color configurations gets broader as the rank $N$ increases, starting with the case of SU(2) where no physical zeros are found for external particles with well-defined color quantum numbers. Our second target consists in exploiting color-kinematics duality to study planar zeros in the gravitational case, where we find that the five-graviton amplitude vanishes whenever the process is planar. This can be understood applying the BCJ prescription to the equation determining the zeros in the gauge case. By replacing color factors with kinematic numerators satisfying color-kinematics duality, the condition for the planar zero is seen to be identically satisfied without further kinematic constraints. The plan of the paper is as follows. In Section \ref{sec:5-gluon} we review the calculation of the five gluon amplitude using the MHV formalism. Section \ref{sec:gluon_planar} is devoted to the conditions for planar zeros in the gauge case, while in Section \ref{sec:permutations} we study the transformation of the loci of planar zeros under permutations of the color labels of the external gluons. In Section \ref{sec:graviton} the graviton amplitude is obtained using color-kinematics duality and the condition for the existence of amplitude zeros is obtained. Finally, in Section \ref{sec:conclusions} we summarize our conclusions. \section{The five-gluon amplitude} \label{sec:5-gluon} In this section we revisit the construction of the five-gluon amplitude \begin{align} g(p_{1},a_{1})+g(p_{2},a_{2})\longrightarrow g(p_{3},a_{3})+g(p_{4},a_{4})+g(p_{5},a_{5}), \end{align} where we take all momenta incoming. The tree level amplitude is computed in terms of 15 nonequivalent trivalent diagrams, leading to the expression \begin{align} \mathcal{A}_{5}&=g^{3}\left({c_{1}n_{1}\over s_{12}s_{45}}+{c_{2}n_{2}\over s_{23}s_{15}}+{c_{3}n_{3}\over s_{34}s_{12}} +{c_{4}n_{4}\over s_{45}s_{23}}+{c_{5}n_{5}\over s_{15}s_{34}}+{c_{6}n_{6}\over s_{14}s_{25}} +{c_{7}n_{7}\over s_{13}s_{25}}+{c_{8}n_{8}\over s_{24}s_{13}} \right. \nonumber \\[0.2cm] &+ \left.{c_{9}n_{9}\over s_{35}s_{24}}+{c_{10}n_{10}\over s_{14}s_{35}}+{c_{11}n_{11}\over s_{15}s_{24}} +{c_{12}n_{12}\over s_{12}s_{35}}+{c_{13}n_{13}\over s_{23}s_{14}}+{c_{14}n_{14}\over s_{25}s_{34}} +{c_{15}n_{15}\over s_{13}s_{45}}\right), \label{eq:amplitude_gluon_general} \end{align} where we have introduced the kinematic invariants \begin{align} s_{ij}=(p_{i}+p_{j})^{2}=2 \, p_{i}\cdot p_{j}, \hspace*{1cm} i<j. \end{align} The color factors in Eq. \eqref{eq:amplitude_gluon_general} are given by\footnote{Our conventions for the color factors differ from those in Refs. \cite{review,color-kinematics}.} \begin{align} c_{1} &= f^{a_{1}a_{2}b} f^{ba_{3}c} f^{ca_{4}a_{5}}, \hspace*{1cm} c_{2}=f^{a_{1}a_{5}b} f^{ba_{4}c} f^{ca_{3}a_{2}}, \nonumber \\[0.2cm] c_{3} &= f^{a_{3}a_{4}b} f^{ba_{5}c} f^{ca_{1}a_{2}}, \hspace*{1cm} c_{4}=f^{a_{4}a_{5}b} f^{ba_{1}c} f^{ca_{2}a_{3}}, \nonumber \\[0.2cm] c_{5} &= f^{a_{5}a_{1}b} f^{ba_{2}c} f^{ca_{3}a_{4}}, \hspace*{1cm} c_{6}=f^{a_{1}a_{4}b} f^{ba_{3}c} f^{ca_{5}a_{2}}, \nonumber \\[0.2cm] c_{7} &= f^{a_{1}a_{3}b} f^{ba_{4}c} f^{ca_{5}a_{2}}, \hspace*{1cm} c_{8}=f^{a_{1}a_{3}b} f^{ba_{5}c} f^{ca_{4}a_{2}}, \label{eq:color_factors} \\[0.2cm] c_{9} &= f^{a_{3}a_{5}b} f^{ba_{1}c} f^{ca_{2}a_{4}}, \hspace*{1cm} c_{10}=f^{a_{4}a_{1}b} f^{ba_{2}c} f^{ca_{3}a_{5}}, \nonumber \\[0.2cm] c_{11} &= f^{a_{1}a_{5}b} f^{ba_{3}c} f^{ca_{4}a_{2}}, \hspace*{1cm} c_{12}=f^{a_3a_5b} f^{ba_4c} f^{ca_1a_2}, \nonumber \\[0.2cm] c_{13} &= f^{a_{1}a_{4}b} f^{ba_{5}c} f^{ca_{3}a_{2}}, \hspace*{1cm} c_{14}=f^{a_{5}a_{2}b} f^{ba_1c} f^{ca_{3}a_{4}}, \nonumber \\[0.2cm] c_{15} &= f^{a_{1}a_{3}b} f^{ba_{2}c} f^{ca_{4}a_{5}}, \nonumber \end{align} and satisfy nine independent Jacobi identities \begin{align} c_{3}-c_{5}+c_{14}&=0, \hspace*{1.15cm} c_{3}-c_{1}-c_{12}=0, \nonumber \\[0.2cm] c_{4}-c_{1}+c_{15}&=0, \hspace*{1.15cm} c_{4}+c_{2}-c_{13}=0, \nonumber \\[0.2cm] c_{5}+c_{2}-c_{11}&=0, \hspace*{1cm} c_{13}-c_{6}+c_{10}=0, \label{eq:jacobi_id}\\[0.2cm] c_{14}-c_{7}+c_{6}&=0, \hspace*{1.15cm} c_{7}-c_{8}+c_{15}=0, \nonumber \\[0.2cm] c_{8}-c_{9}-c_{11}&=0, \hspace*{0.85cm} (c_9-c_{10}+c_{12}=0). \nonumber \end{align} On general grounds, the amplitude can be written in terms of color-ordered amplitudes as \begin{align} \mathcal{A}_{5}=g^{3}\sum_{\sigma\in S_{4}}c[1,\sigma(2,3,4,5)]A_{5}[1,\sigma(2,3,4,5)], \end{align} where the sum is over noncyclic permutations of the external legs. However, the set of color-ordered amplitudes is over complete, a fact expressed by the Kleiss-Kuijf relations \cite{kleiss_kuijf}. In the case of the five-point amplitude, there are $5\times 4$ different ways of choosing a basis in the space of independent color structures $\mathsf{TCS}_{5}$ \cite{kol_shir}. We select one of these basis by fixing the incoming gluons (see Fig. \ref{fig:KK-rel}), so the five-gluon amplitude in Eq. \eqref{eq:amplitude_gluon_general} can be re-expressed in terms of $3!$ color ordered amplitudes according to \begin{align} \mathcal{A}_{5}=g^{3}\sum_{\sigma\in S_{3}}c[1,2,\sigma(3,4,5)]A_{5}[1,2,\sigma(3,4,5)], \label{eq:kleiss-kuijf_exp} \end{align} \begin{figure}[t] \centerline{\includegraphics[scale=0.37]{perm_topologies.pdf}} \caption[]{Graphic representation of the choice of subamplitudes basis in the implementation of the Kleiss-Kuijf relations.} \label{fig:KK-rel} \end{figure} where the subamplitudes are explictly given in terms of the numerators $n_{i}$ by \begin{align} A_{5}[1,2,3,4,5]&={n_{1}\over s_{12}s_{45}}-{n_{2}\over s_{23}s_{15}}+{n_{3}\over s_{34}s_{12}}+{n_{4}\over s_{45}s_{23}} +{n_{5}\over s_{15}s_{34}}, \nonumber \\[0.2cm] A_{5}[1,2,3,5,4]&=-{n_{1}\over s_{12}s_{45}}-{n_{13}\over s_{23}s_{14}}+{n_{12}\over s_{35}s_{12}}-{n_{4}\over s_{45}s_{23}} +{n_{10}\over s_{14}s_{35}}, \nonumber \\[0.2cm] A_{5}[1,2,4,3,5]&=-{n_{12}\over s_{12}s_{35}}-{n_{11}\over s_{24}s_{15}}-{n_{3}\over s_{34}s_{12}}+{n_{9}\over s_{35}s_{24}} -{n_{5}\over s_{15}s_{34}}, \\[0.2cm] A_{5}[1,2,4,5,3]&={n_{12}\over s_{12}s_{35}}-{n_{8}\over s_{24}s_{13}}-{n_{1}\over s_{45}s_{12}}-{n_{9}\over s_{35}s_{24}} -{n_{15}\over s_{13}s_{45}}, \nonumber \\[0.2cm] A_{5}[1,2,5,3,4]&=-{n_{3}\over s_{12}s_{34}}-{n_{6}\over s_{25}s_{14}}-{n_{12}\over s_{35}s_{12}}+{n_{14}\over s_{34}s_{25}} -{n_{10}\over s_{14}s_{35}}, \nonumber \\[0.2cm] A_{5}[1,2,5,4,3]&={n_{3}\over s_{12}s_{34}}-{n_{7}\over s_{25}s_{13}}+{n_{1}\over s_{12}s_{45}}-{n_{14}\over s_{34}s_{25}} +{n_{15}\over s_{13}s_{45}}. \nonumber \end{align} Going back to the expression for the color factors in Eq. \eqref{eq:color_factors}, these partial amplitudes are respectively associated with the six color factors $c_{7}$, $c_{8}$, $c_{6}$, $c_{13}$, $c_{11}$, and $c_{2}$. At this point we can exploit the generalized gauge freedom in the definition of the numerators to implement color-kinematics duality, so the numerators $n_{i}$ mimic the Jacobi identities \eqref{eq:jacobi_id}. Solving the corresponding equations we can eliminate $n_{7}$ to $n_{15}$ finding the following solution for the numerators in terms of the basis of color-ordered amplitudes \begin{align} n_{1}&=-n_{12}=n_{15}=s_{12}s_{45}A_{5}[1,2,3,4,5], \nonumber \\[0.2cm] n_{2}&=n_{3}=n_{4}=n_{5}=n_{11}=n_{13}=n_{14}=0, \nonumber \\[0.2cm] n_{6}&=n_{7}=n_{10}=s_{14}s_{35}A_{5}[1,2,3,5,4]+s_{14}(s_{35}+s_{45})A_{5}[1,2,3,4,5], \\[0.2cm] n_{8}&=n_{9}=s_{14}s_{35}A_{5}[1,2,3,5,4]+(s_{14}s_{35}+s_{14}s_{45}+s_{12}s_{45})A_{5}[1,2,3,4,5]. \nonumber \end{align} Color-kinematics duality is independent of the polarization of the gluons. Here we are going to use the MHV formalism and assign negative helicity to the incoming gluons. Using the Parke-Taylor formula \cite{parke_taylor} we have \begin{align} A_{5}[1^{-},2^{-},\sigma(3^{+},4^{+},5^{+})]=i{\langle 12\rangle^{4}\over \langle 12\rangle\langle 2\sigma(3)\rangle \langle \sigma(3)\sigma(4)\rangle \langle \sigma(4)\sigma(5)\rangle\langle \sigma(5)1\rangle}, \end{align} for any permutation $\sigma\in S_{3}$ of the last three indices. Expressing in addition the kinematic invariants in terms of spinors, $s_{ij}=\langle ij\rangle[ji]$, we arrive at the following expressions for the numerators \begin{align} n_{1}&=-n_{12}=n_{15}=i{\langle 12\rangle^{4}[21][54]\over \langle 23\rangle\langle 34\rangle \langle 51\rangle}, \nonumber \\[0.2cm] n_{6}&=n_{7}=n_{10}=i{\langle 12\rangle^{4}[14][52]\over \langle 23\rangle\langle 34\rangle \langle 51\rangle}, \nonumber \\[0.2cm] n_{8}&=n_{9}=i{\langle 12\rangle^{4}[24][51]\over \langle 23\rangle\langle 34\rangle \langle 51\rangle}, \hspace*{1cm} \label{eq:numerators} \\[0.2cm] n_{2}&=n_{3}=n_{4}=n_{5}=n_{11}=n_{13}=n_{14}=0. \nonumber \end{align} With this result, the five-gluon amplitude can be written as \begin{align} \mathcal{A}_{5}&=-ig^{3}\langle 12\rangle^{3}\left({c_{2}\over \langle 23\rangle\langle 34\rangle\langle 45\rangle\langle 51\rangle} +{c_{6}\over \langle 25\rangle\langle 53\rangle\langle 34\rangle\langle 41\rangle} +{c_{7}\over \langle 25\rangle\langle 54\rangle\langle 43\rangle\langle 31\rangle}\right. \nonumber \\[0.2cm] &\left.+{c_{8}\over \langle 24\rangle\langle 45\rangle\langle 53\rangle\langle 31\rangle} +{c_{11}\over \langle 24\rangle\langle 43\rangle\langle 35\rangle\langle 51\rangle} +{c_{13}\over \langle 23\rangle\langle 35\rangle\langle 54\rangle\langle 41\rangle}\right). \label{eq:5g_amplitude_spinors} \end{align} Alternatively, this expression can be obtained from Eq. \eqref{eq:kleiss-kuijf_exp} by a direct application of the Parke-Taylor formula. The spinor products appearing in the five-gluon amplitude can now be recast in terms of momenta. Working in the center-of-mass frame, the incoming momenta take the form \begin{align} p_{1}={\sqrt{s}\over 2}(1,0,0,1), \hspace*{1cm} p_{2}={\sqrt{s}\over 2}(1,0,0,-1). \end{align} On the other hand, for the three outgoing gluons their spatial momenta are parametrized using stereographic coordinates $\zeta_{a}\in \mathbb{C}$ (with $a=3,4,5$) according to \begin{align} p_{a}=-\omega_{a}\left(1,{\zeta_{a}+\overline{\zeta}_{a}\over 1+\zeta_{a}\overline{\zeta}_{a}},i{\overline{\zeta}_{a}-\zeta_{a}\over 1+\zeta_{a}\overline{\zeta}_{a}},{\zeta_{a}\overline{\zeta}_{a}-1\over 1+\zeta_{a}\overline{\zeta}_{a}}\right), \end{align} where the global minus sign reflects that all momenta are taken entering the diagram. The stereographic coordinates $\zeta_{a}$ are related to the rapidity $y_{a}$ and the azimuthal angle $\phi_{a}$ by \begin{align} \zeta_{a}=e^{y_{a}+i\phi_{a}}. \end{align} \section{Gauge planar zeros} \label{sec:gluon_planar} We focus now on planar five-gluon scattering with general color quantum numbers. Since the incoming particles travel along the $z$ axis, without loss of generality we can take all momenta lying on the $xz$-plane. This means that $p_{a}^{y}=0$ and therefore $\zeta_{a}$ has to be real and the outgoing momenta read \begin{align} p_{a}=-{\omega_{a}\over 1+\zeta_{a}^{2}}(1+\zeta_{a}^{2},2\zeta_{a},0,\zeta_{a}^{2}-1). \end{align} Alternatively, the planarity condition implies that all emitted particles have azimuthal angles with either $\phi_{a}=0$ or $\phi_{a}=\pi$. Implementing momentum conservation $p_{1}+\ldots+p_{5}=0$ gives three independent equations that determine the gluon energies $\omega_{a}$ in terms of the center-of-mass energy $\sqrt{s}$ and the flight directions of the outgoing gluons labelled by $\zeta_{a}$, \begin{align} \omega_{3}&={\sqrt{s}\over 2}{(1+\zeta_{3}^{2})(1+\zeta_{4}\zeta_{5})\over (\zeta_{3}-\zeta_{4})(\zeta_{3}-\zeta_{5})}, \nonumber \\[0.2cm] \omega_{4}&={\sqrt{s}\over 2}{(1+\zeta_{4}^{2})(1+\zeta_{3}\zeta_{5})\over (\zeta_{4}-\zeta_{3})(\zeta_{4}-\zeta_{5})}, \label{eq:energies_dehomog}\\[0.2cm] \omega_{5}&={\sqrt{s}\over 2}{(1+\zeta_{5}^{2})(1+\zeta_{3}\zeta_{4})\over (\zeta_{5}-\zeta_{3})(\zeta_{5}-\zeta_{4})}. \nonumber \end{align} Furthermore, the finite positive energy condition $0\leq \omega_{a}<\infty$ imposes constraints on the possible values of $\zeta_{a}$. In particular, let us remark that finite energy implies that $\zeta_{a}\neq \zeta_{b}$ for $3\leq a<b\leq 5$. Using this parametrization, the amplitude \eqref{eq:5g_amplitude_spinors} takes the form \begin{align} \mathcal{A}_{5}&={2ig^{3}\over\sqrt{s}}{(\zeta_{3}-\zeta_{4})(\zeta_{3}-\zeta_{5})(\zeta_{4}-\zeta_{5})\over (1+\zeta_{3}\zeta_{4})(1+\zeta_{3}\zeta_{5})(1+\zeta_{4}\zeta_{5})}\left[-c_{2}{\zeta_{5}-\zeta_{3}\over \zeta_{3}} -c_{6}{\zeta_{4}-\zeta_{5}\over \zeta_{5}}\right. \nonumber \\[0.2cm] &\left.+c_{7}{\zeta_{3}-\zeta_{5}\over \zeta_{5}}-c_{8}{\zeta_{3}-\zeta_{4}\over \zeta_{4}} +c_{11}{\zeta_{5}-\zeta_{4}\over\zeta_{4}}+c_{13}{\zeta_{4}-\zeta_{3}\over \zeta_{3}} \right]. \label{eq:amplitude_zetas} \end{align} In order to find the zeros of the amplitude, we notice that the finite energy condition implies that the prefactor can never vanish. As a consequence, we find the following equation depending on the color factors \begin{align} c_{2}{\zeta_{5}-\zeta_{3}\over \zeta_{3}} +c_{6}{\zeta_{4}-\zeta_{5}\over \zeta_{5}}-c_{7}{\zeta_{3}-\zeta_{5}\over \zeta_{5}} \hspace*{3cm}\nonumber \\[0.2cm] +c_{8}{\zeta_{3}-\zeta_{4}\over \zeta_{4}} -c_{11}{\zeta_{5}-\zeta_{4}\over\zeta_{4}}-c_{13}{\zeta_{4}-\zeta_{3}\over \zeta_{3}}=0. \label{eq:planar_cond1} \end{align} The planar zero condition just derived is a homogeneous equation of vanishing degree. Since the amplitude \eqref{eq:amplitude_zetas} diverges whenever any of the $\zeta_{a}$ vanishes, we can multiply the previous equation by $\zeta_{3}\zeta_{4}\zeta_{5}$ without generating spurious solutions in the physical region. Taking projective coordinates \begin{align} (\zeta_{3},\zeta_{4},\zeta_{5})=\lambda(1,U,V), \hspace*{1cm} \lambda,U,V\neq 0 \label{eq:projective_coord} \end{align} the planar zeros of the five-gluon amplitude are determined by the loci defined by the following equation \begin{align} c_{7}U-c_{8}V-c_{6}U^{2}+c_{11}V^{2}+(c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13})UV+c_{13}U^{2}V-c_{2}UV^{2}=0. \label{eq:cubic_eq} \end{align} Moreover, this equation is homogeneous in the color factors and therefore independent of the normalization of the gauge group generators. Since there exists a normalization of the generators that makes all structure constants integer numbers \cite{sattinger-weaver}, the planar zeros are determined by a cubic curve with integer coefficients. In terms of the projective coordinates \eqref{eq:projective_coord}, the energies of the outgoing particles take the form \begin{align} \omega_{3}&={\sqrt{s}\over 2}{(1+\lambda^{2})(1+\lambda^{2}UV)\over \lambda^{2}(1-U)(1-V)}, \nonumber \\[0.2cm] \omega_{4}&={\sqrt{s}\over 2}{(1+\lambda^{2}U^{2})(1+\lambda^{2}V)\over \lambda^{2}(U-1)(U-V)}, \label{eq:omegas_UV}\\[0.2cm] \omega_{5}&={\sqrt{s}\over 2}{(1+\lambda^{2}V^{2})(1+\lambda^{2}U)\over \lambda^{2}(V-1)(V-U)}.\nonumber \end{align} We have seen already that in order to keep the amplitude finite we have to exclude $U=0$ and $V=0$ from the physical region. Now, energy finiteness further demands that $U\neq 1$, $V\neq 1$, and $U\neq V$. By requiring $\omega_{a}\geq 0$ we find that, for example, the region $U>0$, $V>0$ has to be considered unphysical as well. Indeed, if this is the case all three numerators in \eqref{eq:omegas_UV} are positive whereas the three denominators cannot have the same sign simultaneously. As a consequence, at least one of the energies has to be negative and the configuration is excluded. Studying the values of $U$ and $V$ in which the three energies are simultaneously positive for a given $\lambda$, we arrive at the physical regions shown in Fig. \ref{fig:physical_regions}. Notice that the plot is symmetric under the exchange $U\leftrightarrow V$. The conformal structure of the equation defining the planar zeros indicates that each solution of Eq. \eqref{eq:cubic_eq} can be realized in infinitely many physical setups, depending on the value of the parameter $\lambda$. Notice that the boundaries of the allowed regions depend on $\lambda$ as well, so they move as this parameter varies, while the position of the zeros, being a projective invariant, remain fixed. \begin{figure}[t] \centerline{\includegraphics[scale=0.5]{physical_regions.pdf}} \caption[]{Physical regions in the $UV$ plane for a given value of $\lambda$. The shadowed regions are unphysical points where the energy is negative for at least one of the outgoing gluons. Dashed lines correspond to the soft limits in which one or several energies tend to zero. Solid lines, including the axes, represent also unphysical points.} \label{fig:physical_regions} \end{figure} A particularly interesting regime is the soft limit, in which one or various of the emitted gluon energies tend to zero. From Eq. \eqref{eq:omegas_UV} we see that the points in the $(U,V)$ plane for which $\omega_{a}$ vanish are given by \begin{align} UV&=-{1\over \lambda^{2}} \hspace*{1cm} (\omega_{3}=0), \nonumber \\[0.2cm] V&=-{1\over \lambda^{2}} \hspace*{1cm} (\omega_{4}=0), \\[0.2cm] U&=-{1\over \lambda^{2}} \hspace*{1cm} (\omega_{5}=0), \nonumber \end{align} which are indicated by the dashed lines in Fig. \ref{fig:physical_regions}. On general grounds, a planar zero corresponding to a point of the cubic \eqref{eq:cubic_eq} can be physically captured in the soft limit provided there is a value of $\lambda$ for which this point collides against any of the ``soft'' lines defining the boundaries of the physical region. The first example to analyze is the case of incoming gluons in a singlet state for arbitrary gauge group, already studied in \cite{harland-lang}. Using the fact that $f^{da_{3}b}f^{ba_{4}c}f^{ca_{5}d}\sim f^{a_{3}a_{4}a_{5}}$, we find \begin{align} c_{2}=c_{6}=-c_{7}=c_{8}=-c_{11}=-c_{13}=-f^{a_{3}a_{4}a_{5}}. \end{align} The cubic equation then reads \begin{eqnarray} U+V+U^{2}+V^{2}-6 UV +U^{2}V+UV^{2}=0. \label{eq:cubid_eq_singlet} \end{eqnarray} The associated algebraic curve is represented in Fig. \ref{fig:singlet_case}. Comparing with Fig. \ref{fig:physical_regions} we see that for small enough $\lambda$ there is indeed a large part of the curve lying on physically allowed regions. In particular, for $\lambda<1$ there are solutions with arbitrarily large $|U|$ and $|V|$. \begin{figure}[t] \centerline{\includegraphics[scale=0.55]{singlet_case_curve2.pdf}\hspace*{2cm}\includegraphics[scale=0.55]{singlet_case_curve.pdf}} \caption[]{Algebraic curve associated with the planar zeros for incoming gluons in a singlet state, Eq. \eqref{eq:cubid_eq_singlet}. The right panel shows a blowup of the region around $(U,V)=(0,0)$.} \label{fig:singlet_case} \end{figure} We study next the loci defined by Eq. \eqref{eq:cubic_eq} for SU($N$) gauge groups with different ranks and various color configurations: \paragraph{SU($\mathbf{2}$).} In the case of SU(2) it is easy to write a closed expression for the color factors \begin{align} c_{2}&=\delta^{a_{3}a_{4}}\epsilon^{a_{2}a_{5}a_{1}}-\delta^{a_{2}a_{4}}\epsilon^{a_{3}a_{5}a_{1}}, \nonumber \\[0.2cm] c_{6}&=\delta^{a_{5}a_{3}}\epsilon^{a_{2}a_{4}a_{1}}-\delta^{a_{2}a_{3}}\epsilon^{a_{5}a_{4}a_{1}}, \nonumber \\[0.2cm] c_{7}&=\delta^{a_{1}a_{4}}\epsilon^{a_{3}a_{5}a_{2}}-\delta^{a_{3}a_{4}}\epsilon^{a_{1}a_{5}a_{2}}, \nonumber \\[0.2cm] c_{8}&=\delta^{a_{2}a_{5}}\epsilon^{a_{4}a_{1}a_{3}}-\delta^{a_{4}a_{5}}\epsilon^{a_{2}a_{1}a_{3}}, \\[0.2cm] c_{11}&=\delta^{a_{4}a_{3}}\epsilon^{a_{2}a_{5}a_{1}}-\delta^{a_{2}a_{3}}\epsilon^{a_{4}a_{5}a_{1}}, \nonumber \\[0.2cm] c_{13}&=\delta^{a_{4}a_{5}}\epsilon^{a_{1}a_{2}a_{3}}-\delta^{a_{1}a_{5}}\epsilon^{a_{4}a_{2}a_{3}}, \nonumber \end{align} where a convenient normalization of the gauge group generators has been chosen. In principle, the color factors can only take the values $0,\pm 1$, and $\pm 2$, since each term on the right-hand side of these equations is either $0$ or $\pm 1$. However, the case $\pm 2$ is excluded. The reason is that due to the structure of indices of the Levi-Civita tensor, sharing the last two entries, they cannot have oposite signs. As a consequence, they cannot add up and we conclude that for SU(2) the color factors satisfy $c_{i}=0,\pm 1$. An exploration of the possible external color numbers show that there are no solutions containing physical points. We illustrate this with a few examples. Our first case has color structure $(a_{1},a_{2},a_{3},a_{4},a_{5})=(2,3,1,1,1)$, giving the same value for all color factors \begin{align} c_{2}=c_{6}=c_{7}=c_{8}=c_{11}=c_{13}=1. \end{align} The resulting cubic equation completely factorizes as \begin{align} (U-1)(V-1)(U-V)=0. \label{eq:curve1_SU(2)} \end{align} We see that the three solutions lie outside the physical region and as a consequence there are no planar zeros for this gauge configuration. Next we try $(a_{1},a_{2},a_{3},a_{4},a_{5})=(2,2,2,1,3)$, which corresponds to color factors \begin{align} c_{2}=c_{7}=c_{8}=c_{13}=0, \hspace*{1cm} c_{6}=-c_{11}=1. \end{align} In this case the equation for the zeros becomes quadratic and factorizes as \begin{align} (U-V)^{2}=0. \end{align} The geometric loci of zeros coincide again with the unphysical region corresponding to two particles in the final state with infinite energy. As a last example, we take $(a_{1},a_{2},a_{3},a_{4},a_{5})=(1,2,2,2,3)$, which gives \begin{align} c_{2}=c_{8}=c_{11}=c_{13}=0, \hspace*{1cm} c_{6}=c_{7}=1. \end{align} In this case the cubic again degenerates into a quadratic equation \begin{align} U(U-1)=0, \end{align} which has no physical solutions. To summarize, a scan of possible values of the external color numbers show that the only curves obtained in this case coincide with unphysical regions in the plot of Fig. \ref{fig:physical_regions}, $U=0,1$, $V=0,1$ or $U=V$. The only possibility for planar zeros in this case is to consider external states without well-defined color numbers, such as the singlet case studied above. \paragraph{SU($\mathbf{3}$).} We work out a first example where we take color quantum numbers $(a_{1},a_{2},a_{3},a_{4},a_{5})=(7,7,6,1,5)$ and color factors \begin{align} c_{2}=-c_{7}=c_{8}=-c_{13}=2, \hspace*{1cm} c_{6}=-c_{11}=-1. \end{align} The planar zeros are given by the factorized cubic \begin{align} (U+V-2)(U+V-2UV)=0. \label{eq:curve_SU(3)_ex1} \end{align} This is a hyperbola together with its tangent at $(U,V)=(1,1)$ (see the left panel of Fig. \ref{fig:SU(3)_case1}). The loci has nonvanishing intersection with the physically allowed region in the $UV$ plane for appropriate values of $\lambda$. \begin{figure}[t] \centerline{\includegraphics[scale=0.55]{SU3_plot1.pdf} \hspace*{2cm}\includegraphics[scale=0.55]{SU3_plot2.pdf} } \caption[]{Left panel: curve \eqref{eq:curve_SU(3)_ex1} giving the planar zeros for the SU(3) five-gluon amplitude with $(a_{1},a_{2},a_{3},a_{4},a_{5})=(7,7,6,1,5)$. Right panel: the same for $(a_{1},a_{2},a_{3},a_{4},a_{5})=(1,4,1,2,6)$, corresponding to Eq. \eqref{eq:second_hyp_SU(2)}.} \label{fig:SU(3)_case1} \end{figure} A different hyperbola is obtained for $(a_{1},a_{2},a_{3},a_{4},a_{5})=(1,4,1,2,6)$ with \begin{align} c_{2}=-c_{11}=-1,\hspace*{1cm} c_{6}=-4, \hspace*{1cm} c_{7}=c_{8}=0, \hspace*{1cm} c_{13}=-2. \end{align} The equation determining the zeros also factorizes in this case, giving \begin{align} (2U-V)(-2U+V+UV)=0. \label{eq:second_hyp_SU(2)} \end{align} Again, we have a hyperbola and one of its tangents, this time at the origin. The curves are shown in the RHS panel of Fig. \ref{fig:SU(3)_case1}. As in the SU(2) cases all examples explored for the gauge group SU(3) show factorization of the cubic equation. In this latter case, however, not only the type of curves is enlarged to include hyperbolas which were absent for SU(2), but the curves contain physical points. In addition, considering quantum numbers in a SU(2) subgroup of SU(3) generates the curves obtained for the former group. \paragraph{SU(5).} Enlarging the gauge group to SU(5) brings more general types of cubic algebraic curves. This is for example the case taking $(a_{1},a_{2},a_{3},a_{4},a_{5})=(17,19,19,18,23)$. The resulting color factors are \begin{eqnarray} c_{2}=c_{13}=0, \hspace*{1cm} c_{6}=c_{8}=2, \hspace*{1cm} c_{7}=c_{11}=1. \end{eqnarray} Since $c_{2}$ and $c_{13}$ vanish, it results in the following quadratic equation determining the planar zeros \begin{eqnarray} U - 2 U^{2} - 2 V + 2 U V + V^{2}=0. \label{eq:alg_curve1_SU(5)} \end{eqnarray} Unlike the examples encountered for SU(2) and SU(3), this curve does not factorize and corresponds to the hyperbola shown in the LHS panel of Fig. \ref{fig:plotsSU(5)1}. \begin{figure}[t] \centerline{\includegraphics[scale=0.55]{SU_5_example1.pdf}\hspace*{2cm}\includegraphics[scale=0.55]{SU_5_example2.pdf}} \caption[]{Two examples of planar zero curves for the group SU(5): the hyperbola in the LHS panel corresponds to $(a_{1},a_{2},a_{3},a_{4},a_{5})=(17,19,19,18,23)$, resulting in Eq. \eqref{eq:alg_curve1_SU(5)}. Equation \eqref{eq:SU(5)curve_ex2} is represented on the RHS panel, corresponding to color indices $(a_{1},a_{2},a_{3},a_{4},a_{5})=(19,18,23,17,19)$.} \label{fig:plotsSU(5)1} \end{figure} A second interesting example is provided by $(a_{1},a_{2},a_{3},a_{4},a_{5})=(19,18,23,17,19)$. The corresponding color factors are \begin{eqnarray} c_{2}=c_{11}=0, \hspace*{1cm} c_{6}=c_{8}=2, \hspace*{1cm} c_{7}=c_{13}=1. \end{eqnarray} The resulting equation for the zero \begin{eqnarray} U - 2 U^{2} - 2 V + 2 U V + U^{2} V=0 \label{eq:SU(5)curve_ex2} \end{eqnarray} is the cubic curve shown in the RHS panel of Fig. \ref{fig:plotsSU(5)1}. As a last example we take $(a_{1},a_{2},a_{3},a_{4},a_{5})=(19,19,18,23,17)$, with color factors \begin{eqnarray} c_{2}=-c_{7}=-2, \hspace*{1cm} c_{6}=c_{8}=-c_{11}=-c_{13}=1. \end{eqnarray} We get the cubic curve \begin{eqnarray} 2U-U^{2}-V-U^{2}V-V^{2}+2UV^{2}=0, \label{eq:curveSU(3)Ex3} \end{eqnarray} which, as shown in Fig. \ref{fig:plotSU(5)Ex3}, contains a singular point at $(U,V)=(1,1)$. \begin{figure}[t] \centerline{\includegraphics[scale=0.55]{SU_5_example3.pdf}\hspace*{2cm} \includegraphics[scale=0.55]{SU_5_example3_2.pdf}} \caption[]{Singular curve in Eq. \eqref{eq:curveSU(3)Ex3} corresponding to $(a_{1},a_{2},a_{3},a_{4},a_{5})=(19,19,18,23,17)$. The RHS panel blows up the region around the origin.} \label{fig:plotSU(5)Ex3} \end{figure} We see how SU(5) provides more general types of curves than the ones found for unitary groups of lower rank. We also have to take into account that SU(5) contains SU(3) and SU(2) subgroups. Using the standard generators (see, for example, \cite{mohapatra}) these subgroups are respectively generated by $\{T^{1},\ldots,T^{8}\}$ and $\{T^{21}, T^{22},T^{23}\}$. Thus, setting the external indices in the subsets $(1,\ldots,8)$ or $(21,22,23)$ we recover previous examples. For instance, $(a_{1},a_{2},a_{3},a_{4},a_{5})=(7,7,6,1,5)$ gives the curve shown in the LHS panel of Fig. \ref{fig:SU(3)_case1}, whereas $(a_{1},a_{2},a_{3},a_{4},a_{5})=(22,23,21,21,21)$ reproduces Eq. \eqref{eq:curve1_SU(2)}. \section{Planar zeros and color permutations} \label{sec:permutations} It is interesting to see how the zeros here investigated transform under permutations of the color quantum numbers of the external particles. We begin considering those permutations preserving the choice of amplitudes basis implied by Eq. \eqref{eq:kleiss-kuijf_exp}. These are elements of $S_{3}$ permuting the color indices of the three outgoing gluons (see Fig. \ref{fig:KK-rel}). In order to find the action of these permutations on the geometric loci of planar zeros, we can see how the color factors \eqref{eq:color_factors} transform under permutations of the $(a_{3},a_{4},a_{5})$ color indices. Here instead we use a more geometric approach and work with the homogenization of the cubic equation \eqref{eq:cubic_eq} \begin{align} c_{7}Z^{2}U-c_{8}Z^{2}V-c_{6}ZU^{2}+c_{11}ZV^{2}+(c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13})ZUV \nonumber \\[0.2cm] +c_{13}U^{2}V-c_{2}UV^{2}=0. \label{eq:cubic_eq_hom} \end{align} The group $S_{3}$ acts passively by permutation of the coordinates $(Z,U,V)$. Let us discuss the geometrical meaning of these transformations. Equation \eqref{eq:cubic_eq_hom} is defined in the whole projective plane, which is covered by the three affine patches centered at $(1,0,0)$, $(0,1,0)$, and $(0,0,1)$. The group $S_{3}$ is generated by $S_{2}$ transformations interchanging the two coordinates within each patch, together with cyclic permutations of the three patches. This defines a coset decomposition of $S_{3}$ with respect to its cyclic subgroup generated by $(123)$. The physically allowed regions shown in Fig. \ref{fig:physical_regions} correspond to the patch centered at $(1,0,0)$. It has been already pointed out that it is invariant under the interchange of the two coordinates $U\leftrightarrow V$. Moreover, the corresponding plots in the other two affine coordinate patches are identical to this one. This is easy to see from Eq. \eqref{eq:energies_dehomog}, where it is glaring that cyclic permutations of the three patches only interchange the energies of the three outgoing gluons. Thus, the positivity conditions remain algebraically the same in any of the three affine patches. The final conclusion is that $S_{3}$ only acts on the axes labels of the plot in Fig. \ref{fig:physical_regions}. This is a passive version of the fact that the energies of the outgoing gluons are determined by momentum conservation alone and that the color structures play no role in it. Applying a permutations of $S_{3}$ to Eq. \eqref{eq:cubic_eq_hom}, we find that the color factors transform under the (six-dimensional) regular representation of the group: in particular, if we write $\boldsymbol{C}=(c_{2},c_{6},c_{7},c_{8},c_{11},c_{13})^{T}$ the group $S_{3}$ acts through the matrices \begin{align} (1)(2)(3)&=\left( \begin{array}{cccccc} 1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \end{array} \right), \hspace*{1.3cm} (123)= \left( \begin{array}{cccccc} 0 & 0 & 0 & 1 & 0 & 0 \\ 1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \\ 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 \end{array} \right), \nonumber \\[0.2cm] (132)&=\left( \begin{array}{cccccc} 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 \\ 1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \\ 0 & 0 & 1 & 0 & 0 & 0 \end{array} \right),\hspace*{1cm} (12)(3)= \left( \begin{array}{cccccc} 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \\ 1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \end{array} \right), \hspace*{1cm} \label{eq:regular_rep_S3}\\[0.2cm] (13)(2)&=\left( \begin{array}{cccccc} 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \\ 1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 & 0 \end{array} \right),\hspace*{1cm} (1)(23)= \left( \begin{array}{cccccc} 0 & 0 & 0 & 0 & 0 & 1 \\ 0 & 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 & 0 \\ 1 & 0 & 0 & 0 & 0 & 0 \end{array} \right). \nonumber \end{align} Notice that the combination of color factors in the coefficient of the $ZUV$ term itself transforms with the one-dimensional parity representation, \begin{align} \sigma(c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13})=(-1)^{\pi(\sigma)}(c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13}), \end{align} where $\pi(\sigma)=0,1$ for even and odd permutations respectively. This is a consequence of the fact that $ZUV$ is an invariant under the permutation group. Applying the transformations given by the matrices \eqref{eq:regular_rep_S3} to the curve \eqref{eq:cubid_eq_singlet} obtained in the case of the scattering of two gluons in a singlet state, $\boldsymbol{C}=(1,1,-1,1,-1,-1)^{T}$, we find that it is invariant, since $\boldsymbol{C}$ transforms with the parity of the permutation, $\sigma(\boldsymbol{C})=(-1)^{\pi(\sigma)}\boldsymbol{C}$. This means the curve shown in the plot in Fig. \ref{fig:singlet_case} describes the planar zeros in all three coordinate patches. Incidentally, notice that the curve is invariant as well under the interchange of the two coordinates in any of the three corresponding plots. The curves presented in Section \ref{sec:gluon_planar} are expressed in the patch $(1,0,0)$. Under permutation of the two coordinates, some solutions remain invariant, such as Eq. \eqref{eq:curve_SU(3)_ex1}, or get mapped into a different solution. We can also consider the transformation of the curves with respect to general color permutations. They form the group $\mathsf{TCS}_{5}$ which is identified with the cyclic Lie operad $\mathsf{Lie}(\!(5)\!)$. Its structure has been studied in Ref. \cite{kol_shir}, where it was found that its action on the six independent color structures is given by the following six-dimensional representation of $S_{5}$ \begin{align*} \yng(3,1,1)\,\,\,. \end{align*} Unlike the transformations of $S_{3}$ studied above, those in $\mathsf{TCS}_{5}\backslash S_{3}$ do not act on the geometric loci of planar zeros by permutation of the projective coodinates $(Z,U,V)$. The most obvious example is provided by the interchange of the color indices of the incoming gluons $a_{1}$, $a_{2}$. From Eq. \eqref{eq:color_factors} we find that this transformation acts linearly on $\boldsymbol{C}$ through the matrix \begin{align} (12)(3)(4)(5)=\left( \begin{array}{rrrrrr} 0 & 0 & -1 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & -1 & 0 \\ -1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & -1 \\ 0 & -1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & -1 & 0 & 0 \end{array} \right), \end{align} where we have used the cycle notation for the elements of $S_{5}$. It is interesting to point out that this transformation leaves invariant the coefficient of the $ZUV$ coefficient in \eqref{eq:cubic_eq_hom}. This last property is not shared by other transformations in $\mathsf{TCS}_{5}\backslash S_{3}$. For example, \begin{align} (134)(25)=\left( \begin{array}{rrrrrr} -1 & 0 & 0 & 0 & 0 & 1 \\ 0 & 0 & 1 & 0 & 0 & 0 \\ 0 & -1 & 1 & 0 & 0 & 0 \\ -1 & 0 & 0 & 0 & 1 & 0 \\ -1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & -1 & 0 & 0 \end{array} \right), \hspace*{0.5cm} (1245)(3)=\left( \begin{array}{rrrrrr} -1 & 1 & -1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 \\ -1 & 0 & 0 & 0 & 0 & 1 \\ -1 & 0 & -1 & 1 & 0 & 1 \\ 0 & 1 & -1 & 0 & 0 & 0 \end{array} \right), \end{align} act on $c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13}$ respectively as \begin{align} (134)(25)(c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13})&=c_{2}-c_{6}+c_{7}-c_{8}-c_{11}-c_{13}, \nonumber \\[0.2cm] (1245)(3)(c_{2}+c_{6}-c_{7}+c_{8}-c_{11}-c_{13})&=c_{2}-c_{6}-c_{7}+c_{8}-c_{11}+c_{13}. \end{align} Through its transformations of the color factors, $\mathsf{TCS}_{5}$ acts on the curves determining the planar zeros. In fact, this action can be used to generate the whole orbit of projective curves associated with the permutations of the color quantum numbers of the interacting gluons. \section{Graviton planar zeros from color-kinematics duality} \label{sec:graviton} We turn now to the problem of planar zeros in the five-graviton tree level amplitude. The gravitational amplitude can be constructed from its gluon counterpart \eqref{eq:amplitude_gluon_general} using the BCJ prescription \cite{color-kinematics}, \begin{align} -i\mathcal{M}_{5}&=\left({\kappa\over 2}\right)^{3}\left({n_{1}^{2}\over s_{12}s_{45}}+{n_{2}^{2}\over s_{23}s_{15}}+{n_{3}^{2}\over s_{34}s_{12}} +{n^{2}_{4}\over s_{45}s_{23}}+{n_{5}^{2}\over s_{15}s_{34}}+{n_{6}^{2}\over s_{14}s_{25}} +{n_{7}^{2}\over s_{13}s_{25}}+{n_{8}^{2}\over s_{24}s_{13}} \right. \nonumber \\[0.2cm] &+ \left.{n_{9}^{2}\over s_{35}s_{24}}+{n_{10}^{2}\over s_{14}s_{35}}+{n_{11}^{2}\over s_{15}s_{24}} +{n_{12}^{2}\over s_{12}s_{35}}+{n_{13}^{2}\over s_{23}s_{14}}+{n_{14}^{2}\over s_{25}s_{34}} +{n_{15}^{2}\over s_{13}s_{45}}\right), \end{align} provided the numerators $n_{i}$ satisfy color-kinematics duality, with $\kappa$ the gravitational coupling. Taking the graviton polarizations $(1^{-},2^{-},3^{+},4^{+},5^{+})$, we use our expression for the MHV gauge amplitude given in Eq. \eqref{eq:5g_amplitude_spinors} to write \begin{align} -i\mathcal{M}_{5} &=-i\left({\kappa\over 2}\right)^{3}\langle 12\rangle^{3}\left({n_{2}\over \langle 23\rangle\langle 34\rangle\langle 45\rangle\langle 51\rangle} +{n_{6}\over \langle 25\rangle\langle 53\rangle\langle 34\rangle\langle 41\rangle} +{n_{7}\over \langle 25\rangle\langle 53\rangle\langle 43\rangle\langle 31\rangle}\right. \nonumber \\[0.2cm] &\left.+{n_{8}\over \langle 24\rangle\langle 45\rangle\langle 53\rangle\langle 31\rangle} +{n_{11}\over \langle 24\rangle\langle 43\rangle\langle 35\rangle\langle 51\rangle} +{n_{13}\over \langle 23\rangle\langle 35\rangle\langle 54\rangle\langle 41\rangle}\right). \end{align} Using the form of the gauge theory numerators in Eq. \eqref{eq:numerators}, after a bit of algebra we arrive at the simpler expression \cite{bgk} \begin{align} -i\mathcal{M}_{5}&=-\left({\kappa\over 2}\right)^{3}{\langle 12\rangle^{7}[41][52]\over \langle 12\rangle\langle 14\rangle \langle 23\rangle \langle 25\rangle \langle 34\rangle \langle 35\rangle \langle 45\rangle} \left(1-{\langle 14\rangle\langle 25\rangle [42][51]\over \langle 15\rangle \langle 24\rangle[41][52]}\right). \end{align} The term inside the parenthesis can be further simplified taking into account the relation $s_{ij}=\langle ij\rangle[ji]$, \begin{align} 1-{\langle 14\rangle\langle 25\rangle [42][51]\over \langle 15\rangle \langle 24\rangle[41][52]} =1-\left({\langle 14\rangle\langle 25\rangle\over \langle 15\rangle\langle 24\rangle}\right)^{2} \left({s_{15}s_{24}\over s_{14}s_{25}}\right), \end{align} which in turn can be written as a function of $\Delta \phi_{45}$, the difference of azimuthal angles of particles $4$ and $5$ (see the Appendix of Ref. \cite{harland-lang}), \begin{align} \left({\langle 14\rangle\langle 25\rangle\over \langle 15\rangle\langle 24\rangle}\right)^{2} \left({s_{15}s_{24}\over s_{14}s_{25}}\right)=e^{2i\Delta\phi_{45}}. \end{align} With this, the five-graviton tree level amplitude reads \begin{align} -i\mathcal{M}_{5}&=-\left({\kappa\over 2}\right)^{3}{\langle 12\rangle^{7}[41][52]\over \langle 12\rangle\langle 14\rangle \langle 23\rangle \langle 25\rangle \langle 34\rangle \langle 35\rangle \langle 45\rangle} \left(1-e^{2i\Delta\phi_{45}}\right). \label{eq:grav_ampl} \end{align} Given our choice of reference frame, planarity implies that for any two outgoing momenta their azimuthal angles must satisfy $\Delta\phi_{ij}=0,\pi$. In both cases we find from \eqref{eq:grav_ampl} that the gravitational amplitude vanishes \begin{align} -i\mathcal{M}_{5}\Big|_{\rm planar}=0. \end{align} Unlike the gauge case where the cancellation condition depends on the color factors of the incoming particles, the graviton amplitude automatically vanishes in the limit of planar scattering. We now show that this is a consequence of color kinematic duality. In gauge theories we have seen that the vanishing of the amplitude in the planar case leads to a nontrivial condition on the momenta of the outgoing particles given in Eq. \eqref{eq:planar_cond1}. In fact, the numerators in Eq. \eqref{eq:numerators} have been chosen to satisfy color-kinematics duality, so we can obtain the planar zero condition for gravity by replacing the color factors $\{c_{2},c_{6},c_{7},c_{8},c_{11},c_{13}\}$ in \eqref{eq:planar_cond1} with the corresponding numerators $\{n_{2},n_{6},n_{7},n_{8},n_{11},n_{13}\}$. In terms of the stereographic coordinates, the latter are given by \begin{align} n_{6}&=n_{7}=is^{3\over 2}{(\zeta_{3}-\zeta_{5})\zeta_{5}\over \zeta_{3}(1+\zeta_{4}\zeta_{5})}, \nonumber \\[0.2cm] n_{8}&=is^{3\over 2}{(\zeta_{3}-\zeta_{5})\zeta_{4}\over \zeta_{3}(1+\zeta_{4}\zeta_{5})}, \\[0.2cm] n_{2}&=n_{11}=n_{13}=0. \nonumber \end{align} After this substitution, the condition for the existence of a zero in the amplitude is identically satisfied \begin{align} -n_{2}{\zeta_{5}-\zeta_{3}\over \zeta_{3}} -n_{6}{\zeta_{4}-\zeta_{5}\over \zeta_{5}}+n_{7}{\zeta_{3}-\zeta_{5}\over \zeta_{5}} -n_{8}{\zeta_{3}-\zeta_{4}\over \zeta_{4}} +n_{11}{\zeta_{5}-\zeta_{4}\over\zeta_{4}}+n_{13}{\zeta_{4}-\zeta_{3}\over \zeta_{3}} \nonumber \\[0.2cm] =is^{3\over 2}{(\zeta_{3}-\zeta_{5})\over \zeta_{3}(1+\zeta_{4}\zeta_{5})}\Big(-\zeta_{4}+\zeta_{5}+\zeta_{3}-\zeta_{5}- \zeta_{3}+\zeta_{4}\Big)=0. \label{eq:planar_cond1_grav} \end{align} This implies that, in the gravitational case, planarity is enough to make the amplitude vanish, without additional kinematic conditions to be satisfied by the stereographic coordinates of the outgoing gravitons. \section{Closing remarks} \label{sec:conclusions} \begin{figure}[t] \centerline{\includegraphics[scale=1.0]{vector_plot1.pdf}\hspace*{0.5cm}\includegraphics[scale=1.0]{vector_plot2.pdf}} \caption[]{Graphic representation of the kinematic configuration on the interaction plane for two planar zeros. In both cases the gauge group is SU(3) with color configuration $(7,7,6,1,5)$, so they correspond to two points on the curve shown on the left panel of \ref{fig:SU(3)_case1}. The left panel represents the solution $(\zeta_{3},\zeta_{4},\zeta_{5})=(-1.95,0.4,-4.3)$, whereas in the right panel $(\zeta_{3},\zeta_{4},\zeta_{5})=(0.85,2.5,-0.8)$. In the projective coordinates \eqref{eq:projective_coord}, they correspond to $\lambda=-1.95$, $U=-0.21$, $V=2.21$ and $\lambda=0.85$, $U=2.94$, $V=-0.94$ respectively.} \label{fig:vector_plots} \end{figure} We have studied the presence of planar zeros in both Yang-Mills theories and gravity. For the case of gauge theories, we have represented in Fig. \ref{fig:vector_plots} the kinematics on the interaction plane for two typical planar zeros within the same color configuration. By varying the value of $\lambda\equiv \zeta_{3}$ while keeping $U\equiv\zeta_{4}/\zeta_{3}$ and $V\equiv \zeta_{5}/\zeta_{3}$ constant, these processes can be deformed into a different one with the emission, for example, of one or more soft gluons while the total amplitude remains equal to zero. This happens because planar zeros live in the projective $U$-$V$ plane and are therefore invariant under a simultaneous rescaling of the three outgoing stereographic coordinates. Without loss of generality we considered the situation in which the scattering takes place in the $y=0$ plane. Planar zeros on a different interaction plane can be obtained by applying rotations to the solutions studied here. In particular, the Lorentz group acts on the stereographic coordinates parametrizing the direction of the momenta through $\mbox{SL}(2,\mathbb{C})$ transformations \cite{strominger_et_al} \begin{align} \zeta_{k}'={a\zeta_{k}+b\over c\zeta_{k}+d}, \hspace*{2cm} ad-bc=1, \end{align} where for the incoming particles we have $\zeta_{1}=\infty$ and $\zeta_{2}=0$. Rotations can be spotted by looking for transformations leaving invariant the energies \eqref{eq:energies_dehomog}, together with those of the incoming particles. They are given by \begin{align} \left( \begin{array}{cc} a & b \\[0.2cm] c & d \end{array} \right)=\left( \begin{array}{cc} \xi & -\sqrt{1-\xi^{2}} \\[0.2cm] \sqrt{1-\xi^{2}} & \xi \end{array} \right). \end{align} For real $|\xi|\leq 1$, we parametrize $\xi=\cos\phi$. This corresponds to a rotation of the interaction plane of angle $2\phi$ with respect to the $x$-axis. Alternatively, for $|\xi|>1$, setting $|\xi|=\cosh{\chi}$ the transformation implements a rotation of angle $\sin\phi'=\tanh{2\chi}$ around the $y$-axis. With the results here presented we have shed some light on the origin of the planar zeroes present in Yang-Mills scattering amplitudes. Our results can be generalized to an arbitrary number of external legs at Born level. It will be worth further investigating the effect of quantum corrections. We have also connected, via the BCJ duality, these zeroes to the corresponding ones in gravity. We are currently studying how this picture is modified when the scattering of open and closed strings is considered. \section*{Acknowledgments} We would like to thank Lucian Harland-Lang and Lev Lipatov for discussions. A.S.V. and D. M. J. acknowledge support from the Spanish Government grant FPA2015-65480-P and Spanish MINECO Centro de Excelencia Severo Ochoa Programme (SEV-2012-0249). The work of M.A.V.-M. has been partially supported by Spanish Government grant FPA2015-64041-C2-2-P.
5c91ff88c8161c3ea4b03d30db4026e6160f01ce
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} After the discovery of the extraordinary properties of graphene, the next challenge is to develop mechanisms that allow the enhancement and modulation of such properties. Along these lines, strain engineering is currently one of the trending topics in graphene science because of the possibility to induce new physical phenomena by means of mechanical strain. Examples are modifications on the Fermi velocity,\cite{pereira2009} the modulation of Landau levels spectra,\cite{betancur2015} the generation of pseudo-magnetic fields,\cite{guinea2010} the modulation of the electrical,\cite{kim2009} and thermal conductivities,\cite{ning2009} just to name a few. Furthermore, with the recent advances in experimental techniques to apply strain, there are different reports of uniaxial,\cite{uniax-controlado,isot-comp} biaxial,\cite{isot-comp,biax} and shear\cite{shear} strain in graphene. Interestingly, it has been shown that uniaxial strain can be applied in a controlled, reversible and non destructive way,\cite{uniax-controlado} making it of particular interest. Two of the most studied properties of uniaxially strained graphene are its electronic and vibrational structure. Currently, it is well know that in uniaxially strained graphene the crossing point of the valence and conduction bands at the Fermi level, the so called Dirac point, shifts away from the corner of the Brillouin zone (the ${\bf K}$ point), with no band gap opening.\cite{pereira2009,commentFarjam,Mohr2009,Choi2010} Among the vibrational structure of graphene, the $E_{2g}$ phonon mode at the center of the Brillouin zone (the $\Gamma$ point), is particularly interesting because it is responsible for the G-band in the Raman spectroscopy.\cite{ferrari2007} Under uniaxial strain the $E_{2g}$ phonon mode is split in two modes, one parallel and the other perpendicular to the axis of the applied strain.\cite{Mohiuddin2009,cheng2011} That effect is useful to characterize the direction and strength of the uniaxial strain \cite{Mohiuddin2009,Huang2009,frank2013} via Raman spectroscopy. Even more, the full phonon dispersion,\cite{Mohr2009} Gr\"{u}neisen parameters,\cite{cheng2011} and the origin of the phonon instability at the ideal strength\cite{Liu2007,Hwang2014} have been studied in uniaxially strained graphene using first principles calculations. Regardless the level of understanding of the electronic and vibrational structure of uniaxially strained graphene, some fundamental and important microscopic properties like the electron-phonon (e-ph) coupling needs to be studied in detail. In pristine graphene the e-ph coupling induces strong anomalies in the phonon dispersion,\cite{piscanec2004} contributes to the intrinsic electronic resistivity,\cite{sohier2014} is responsible for most of the linewidth in the Raman G-band,\cite{ferrari2007} and determines the scattering rules for the double resonance Raman 2D-band\cite{ferrari2007}. Even more interesting could be the possibility to induce electron-phonon superconductivity by means of atomic-decorating,\cite{profeta2012,Ludbrook2015} heavy doping,\cite{margine2014} and a combination of doping and biaxial tensile strain.\cite{Chen-si2} Therefore, in order to have a deep understanding of the effects of uniaxial strain on the vibrational, thermal and transport properties, a detailed study of the e-ph coupling in uniaxially strained graphene is mandatory. A key feature of the e-ph coupling are Kohn anomalies:\cite{KA1} anomalous behavior in the phonon dispersion due to an electronic screening of the ionic vibration, which are fully determined by the geometry of the Fermi surface. In graphene the Fermi surface is the Dirac point, thus the Kohn anomalies take place only at the ${\Gamma}$ and ${\bf K}$ points, which are shown as a discontinuity in the frequency derivative of the phonon dispersion of the highest optical (HO) branches.\cite{piscanec2004} Therefore, the e-ph coupling is localized on the transversal (TO) and longitudinal (LO) optical branches at ${\Gamma}$ in the $E_{2g}$ phonon mode. Meanwhile, at ${\bf K}$ the e-ph coupling is almost entirely localized on the TO $A'_1$ phonon mode, with a very small contribution from the double degenerated $E'$ phonon mode on the LO and the longitudinal acoustic (LA) phonon branches,\cite{piscanec2004} which is usually neglected in the study of properties related to the e-ph coupling. In uniaxially strained graphene, the loss of hexagonal symmetry and the shift of the Dirac point is expected to have an effect on the e-ph coupling but, to the best of the authors knowledge, this is not yet reported. In this work we have employed first principles density functional theory (DFT) calculations to systematically study the effects of uniaxial strain along the armchair (AC) and zigzag (ZZ) directions on the Kohn anomalies and e-ph coupling in graphene. In particular, we determine the displacement of the Kohn anomaly from ${\bf K}$, its frequency softening, vibrational modes and e-ph coupling. \footnote{We use the definition of e-ph coupling in graphene following the work of Piscanec et al.\cite{piscanec2004} The details are explained in Appendix \ref{epc-definition}.} We show that uniaxial strain induce a substantial enhancement of the e-ph coupling in the LA branch around ${\bf K}$ with respect to pristine graphene, generating a non-negligible Kohn anomaly. This paper is organized as follow: In Sec. \ref{methodology} we describe the computational details of our first principles calculations. In Sec. \ref{res-struct} we present the changes in the bond length and average force constants that will be useful for forthcoming discussions. The Kohn anomalies in the phonon dispersion, its vibrational phonon modes and frequency shifts are shown in Sec. \ref{results-Kohn anomaly}. An analysis of the uniaxial strain effects in the e-ph coupling is discussed in Sec. \ref{e-ph coupling-res}. In Sec. \ref{conclusions-sec} we summarize our main findings. Finally, we include four Appendix sections with several useful relations used along this article, concerning the structural properties of uniaxially strained graphene, the Kohn anomaly shift, the classical atomic displacement of the discussed phonon modes, and the calculated e-ph coupling quantities. \section{Computational details} \label{methodology} The present calculations were performed within DFT, in the framework of the Mixed Basis Pseudopotential approach (MBPP).\cite{MBPP} Core electrons were replaced by norm conserving pseudopotentials\cite{vanderbiltPP} with non-linear core-corrections included. Valence states were represented by a combination of $s$ and $p$ type localized functions at each atomic sites, complemented with plane waves up to a kinetic energy of 25 Ry. The exchange-correlation functional was treated with the PBE\cite{PBE} parameterization of the generalized gradient approximation. During the structural optimization, the carbons positions were relaxed until the interatomic forces were 0.0001 Ry/Bohr or less. For phonon and e-ph coupling calculations we employed the density functional perturbation theory as implemented in the MBPP code.\cite{Pert} Special attention was paid to the integration in the irreducible Brillouin zone with a 72$\times$72$\times$1 Monkhorst-Pack $k$-point mesh and a small Gaussian broadening of 0.10 eV. This was needed in order to avoid electronic smearing effects on the Kohn anomalies and at the same time obtain converged phonon frequencies. Dynamical matrices were calculated using 12$\times$12$\times$1 and 9$\times$9$\times$1 $q$-points grids for pristine and uniaxial strained graphene, respectively. Full phonon dispersion and force constants were obtained via standard Fourier interpolation. In order to resolve the Kohn anomalies on the phonon dispersion we also computed several low-symmetry $q$-points corresponding to the full $q$-grid of 72$\times$72$\times$1. For the evaluation of the e-ph coupling properties we used a denser $k$-grid of 144$\times$144$\times$1, within a Gaussian broadening varying from 0.05 to 0.30 eV which, however, does not affect our final results. To simulate a single atomic layer, we used the supercell approach and we left at least 12 ${\rm \AA}$ of vacuum space between successive layers to avoid spurious supercell effects on the electronic states and phonon frequencies. \section{Results and Discussion} \subsection{Structural properties} \label{res-struct} For pristine graphene we have obtained a lattice parameter of 2.465 \AA , which corresponds to a bond length of 1.423 \AA. Taking the derivative of the acoustic phonon branches in the limit of ${\bf q}\rightarrow 0$, we estimate a Young modulus of 369 N/m and a Poisson's ratio of 0.182. The calculated elastic constants values are in agreement with the previously experimental and computational reported values. For instance, Politano \textit{et al.}\cite{Politano20124903} perform phonon dispersion measurements from macroscopic graphene samples, and estimate a Young modulus of 342 N/m and a Poisson's ratio of 0.19 from the sound velocities of the TA and LA phonon branches. In the context of previous DFT-based reports, the Young Modulus value varies from 344 to 356 N/m,\cite{Cadelano2010,Liu2007,Bera2010} and the Poisson's ratio from 0.162 to 0.186,\cite{Cadelano2010,Gui2008,Liu2007,Bera2010} depending on the exchange-correlation functional and other numerical approximations. \begin{figure} \includegraphics*[scale=1.0]{fig-esquema.pdf} \caption{ (a) Schematic representation of the lattice vectors, C-C distances ($\alpha$ and $\beta$), and strain directions (AC and ZZ) employed in this work. (b) First Brillouin zone in the reciprocal space with the high symmetry points for uniaxially strained graphene.} \label{esquema} \end{figure} As we described in the Appendix \ref{uniax-strain}, the structural properties of graphene under ZZ and AC strain are defined by the relation between the parallel or applied strain $\varepsilon_{\parallel}$, the perpendicular contraction $\varepsilon_{\bot}$, and the C-C distances $\alpha$ and $\beta$ (see Fig. \ref{esquema} for the definition of strain directions, and the real and reciprocal lattice). In Fig. \ref{struct-prop}(a) we present the computed values for $\varepsilon_{\bot}$ as a function of $\varepsilon_{\parallel}$, and for reference we have included the linear dependence for a constant Poisson's ratio. From that, it clearly shows a non-linear behavior for $\varepsilon_{\parallel} > 2 \%$, which indicates a non-constant Poisson's ratio, in agreement with previous works.\cite{Liu2007,wei2009,cheng2011} Hereinafter, for simplicity $\varepsilon_{\parallel}$ will be referred only as strain. \begin{figure} \includegraphics*[scale=1.0]{fig-struct-prop.pdf} \caption{ Structural properties for uniaxially strained graphene. (a) Evolution of the perpendicular strain as a function of the parallel strain. (b) Changes of bond distances as a function of uniaxial strain. } \label{struct-prop} \end{figure} The changes in the interatomic C-C distances $\alpha$ and $\beta$ are show in Fig. \ref{struct-prop}(b). Although we consider only tensile strain, the C-C distances do not increase in all cases. For ZZ strain there is a small contraction in $\alpha$, corresponding to the bond perpendicular to the direction of the applied strain. In a classical picture, the contraction of $\alpha$ should increase the force constant related to this bond, contrary of what is expected when a material is under tension. To corroborate this picture, we calculate the average force constant related to atom-atom bonds, defined by \begin{equation} \label{forcecte} I(b) = \sqrt{\frac{1}{3}\sum_{i j}\Phi_{i j}^2 (b)}, \end{equation} where $\Phi_{i j}(b)$ represents the force constant matrix assigned to a bond $b$. The respective $I(\alpha)$ and $I(\beta)$ are shown in Fig. \ref{force-ctes} for both ZZ and AC uniaxial strain. In all cases we found that the dominant change in $I(b)$ comes from the longitudinal component of the force constant. Just as expected from the change in the length of the C-C bonds, all the average force constants decrease, except for a small hardening in $I(\alpha)$ under ZZ strain. Such behavior is a key feature in the forthcoming discussion of the phonon frequency shift for the Kohn anomaly. \begin{figure} \includegraphics*[scale=1.0]{fig-force-ctes.pdf} \caption{ Average force constants normalized with respect to value for the bond in pristine graphene as a function of uniaxial strain.} \label{force-ctes} \end{figure} \subsection{Kohn anomalies} \label{results-Kohn anomaly} In order to determine the position of the Kohn anomalies under uniaxial strain, we need to determine the distance $\Delta$ between the Dirac point and the ${\bf K}$ point in the electronic structure, as we described in Appendix \ref{Dirac-cone-shift}. We estimate the evolution of $\Delta$ as a function of the applied strain by an interpolation of the electronic bands at the Fermi level (see Fig. \ref{dp-shift}(a)). We found that $\Delta$ is bigger for strain in the AC direction than in ZZ, although for strains lower than 3 \% it is almost independent of the strain direction. Then, the position of the Kohn anomaly should be at the phonon nesting vector ${\bf q}_{ZZ}$ or ${\bf q}_{AC}$, presented for the unit cell of the reciprocal space in Figs. \ref{dp-shift}(b) and \ref{dp-shift}(c) for the ZZ and AC strain, respectively. \begin{figure} \includegraphics*[scale=1.0]{fig-dp-shift.pdf} \caption{ Dirac point shift from ${\bf K}$. (a) $\Delta$ as a function of ZZ and AC strain. Note that $a$ is the lattice parameter of the strained system and is given by Eq. \ref{latt-cte}. (b) Representation of ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$. The cross marks represent the Dirac points shifted from ${\bf K}$ and ${\bf K'}$.} \label{dp-shift} \end{figure} The phonon dispersion around the Kohn anomaly in uniaxially strained graphene for $\varepsilon = 5 \%$, is show in Fig. \ref{ph-disp}. For an easy reference and comparison, each branch and its respective phonon mode will be identified by its polarization in pristine graphene: LO, TO, and LA. As general trends, at ${\Gamma}$ we can observe the splitting of the $E_{2g}$ phonon mode (see Fig. \ref{ph-disp}(a)), and that the derivative discontinuity of the HO branches depends on the chosen direction along the Brillouin zone. Around ${\bf K}$, the Kohn anomaly in the HO branch shows the expected shift according to our estimation for $\Delta$ (dotted line in Fig. \ref{ph-disp}(b)). More interesting is the new derivative discontinuity on the LA branch at approximately 125 and 129 meV for the AC and ZZ strain, respectively. The fact that such discontinuities occur at the nesting vector that connect two Dirac points (${\bf q}_{ZZ}$ and ${\bf q}_{AC}$) is a direct indication of a Kohn anomaly in the LA branch. This is confirmed in Sec. \ref{e-ph coupling-res} with the analysis of e-ph coupling in the LA branch. \begin{figure} \includegraphics*[scale=0.6, trim=6.5cm 0cm 5cm 0cm clip]{fig-ph-disp.pdf} \caption{(a) Kohn anomalies under uniaxial 5 \% of AC (blue) and ZZ (red) strain at ${\Gamma}$, and (b) ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$. Dotted lines represent the position of ${\bf q}_{ZZ/AC}$. Symbols correspond to computed frequencies: circles for LO, inverse triangles for TO, and squares for LA branches.} \label{ph-disp} \end{figure} For a further discussion of the phonon modes at the Kohn anomaly, we first focus in the ${\Gamma}$ point. As has been reported previously,\cite{Mohiuddin2009,cheng2011} the splitting of the $E_{2g}$ phonon mode results in two modes with eigenvectors which are perpendicular (with smaller softening) and parallel to the strain direction (see Fig. \ref{gamma-modes}(a)). This effect is measured in Raman spectroscopy via the G-band, and because its relevance in graphene characterization, we adopt the same nomenclature that identify as ${\rm G}^+$ (${\rm G}^-$) the band with smaller (higher) softening. \begin{figure} \includegraphics*[scale=1]{fig-gamma-modes.pdf} \caption{(a) Representation of the phonon modes for ${\rm G}^+$ and ${\rm G}^-$ at ${\Gamma}$ and (b) its respective phonon frequency shift. The arrows indicate the instantaneous displacement of the carbon atoms at a particular time.} \label{gamma-modes} \end{figure} The phonon frequency shift for the ${\rm G}^+$ and ${\rm G}^-$ bands are show in Fig. \ref{gamma-modes}(b). The present results for the shift and splitting of the G-band are in good agreement with previous theoretical\cite{Mohiuddin2009,Mohr2009,cheng2011} and experimental\cite{Mohiuddin2009} reports. However, it is important to mention that there are a wide range of reported values because the different set ups and conditions to induce strain on graphene, as well as other effects like substrate interaction, temperature, and the number of graphene layers. Beside that, in the studied range of strain we obtain an almost linear softening in ${\rm G}^+$ and ${\rm G}^-$, which becomes independent of the strain direction for deformations lower than 2 \%. For the anomalies at ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ in the TO and LA branches, we found a polarization of the phonon eigenvectors $\eta_{\kappa s}^{{\bf q}\nu}$ as a function of the strain, such that the atoms moves on ellipses with the mayor axis parallel (LA) and perpendicular (TO) to the strain direction, and whose eccentricity approaches to one as the strain increases, until the ellipses become almost straight lines (see Appendix \ref{app-atom-disp} for a proper description of the classical atomic displacement in graphene). During this evolution the phase difference $\varPhi$ between the atomic displacements along the $x$ and $y$ direction is $\pm \pi/2$, the major and minor axes of the ellipses are defined by the magnitude of the phonon eigenvectors, and the relations $\left|\eta_{x}^{TO}\right| = \left|\eta_{y}^{LA}\right|$ and $\left|\eta_{y}^{TO}\right| = \left|\eta_{x}^{LA}\right|$ are always fulfilled. This means that the Kohn anomaly shift from the high symmetry point ${\bf K}$ induces a mixing of the phonon eigenvectors $\eta_{\kappa s}^{TO}$ and $\eta_{\kappa s}^{LA}$, which belong to the same irreducible representation of the point group of ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$, as in pristine graphene for ${\bf q}$ points outside the high symmetry points $\Gamma$, ${\bf K}$, and ${\bf M}$. Thus the classical atomic displacement on each anomaly are in mutually perpendicular ellipses, but with the same magnitudes for the major and minor axes. The magnitude of the phonon eigenvectors and the eccentricity of the resulting ellipses for the Kohn anomaly in the TO branch are show in Fig. \ref{eigenvectK} with the norm $\sqrt{\left|\eta_{x}\right|^2+\left|\eta_{y}\right|^2}=1$ assumed for simplicity. Within the ZZ (AC) strain along the $x(y)$ Cartesian axis (see Fig. \ref{esquema}), it is clear that the phonon eigenvectors tend to align in the strain direction, especially for ZZ strain where the eccentricity approaches one faster than for the AC strain, resulting in a straight line displacement. \begin{figure} \includegraphics*[scale=1]{fig-eigenvectK.pdf} \caption{ Polarization of the phonon eigenvectors at ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ as a function of strain. (a) Modulus of $\eta_{x,y}$ for the TO branch, and (b) the respective eccentricity for the elliptical displacement of atoms. } \label{eigenvectK} \end{figure} Using the same nomenclature as that for the splitting of the G-band (${\rm G}^{-}$ and ${\rm G}^{+}$) for the anomalies at the TO and LA branches, we will employ the $+$($-$) superindex to indicate that the phonon mode has eigenvectors perpendicular (parallel) to the strain direction and the smaller (higher) softening. A schematic representation of the ${\rm TO}^{+}$ and ${\rm LA}^{-}$ modes and the behavior of the phonon frequency shift are show in Fig. \ref{q-modes}. Unlike the phonon frequency shift in ${\rm G}^{-}$, ${\rm G}^{+}$, and ${\rm LA}^{-}$, in the case of ${\rm TO}^{+}$ the phonon softening is non-linear and becomes nearly constant starting from 2 \% of ZZ strain. In ${\rm TO}_{ZZ}^{+}$ the atoms move along the AC direction, inducing a large distortion of the $\alpha$ bond. Thus, the constant frequency softening is a consequence of the very small increment of the force constant for the $\alpha$ bond, whose length remains almost constant under ZZ strain (see Fig. \ref{struct-prop} and Fig. \ref{force-ctes}). In ${\rm TO}_{AC}^{+}$ the atoms move along the ZZ direction, the atomic distortion is not along the $\alpha$ bond, and therefore the frequency softening is not yet constant as in ${\rm TO}_{ZZ}^{+}$. \begin{figure} \includegraphics*[scale=1]{fig-q-modes.pdf} \caption{(a) Representation of the phonon modes for the TO and LA branches at ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ and (b) the respective phonon frequency shift. The arrows indicate the instantaneous displacement of the carbon atoms at a particular time.} \label{q-modes} \end{figure} \subsection{Electron-phonon coupling} \label{e-ph coupling-res} The computed values for the average e-ph coupling matrix-element square over the Fermi surface in pristine graphene are $\langle g_{{\Gamma},G}^2\rangle = 0.0400 {\rm \ eV}^2$ and $\langle g_{{\bf K},TO}^2\rangle = 0.0989 {\rm \ eV}^2$, which are in excellent agreement with previously reported values.\cite{piscanec2004,Lazzeri2006, Yan2009} We also obtain a value of $0.0037 {\rm \ eV}^2$ for the double degenerate LO and LA branches at ${\bf K}$, which is very small in comparison with the TO branch. The effect of uniaxial strain on the e-ph coupling matrix element square over the Fermi surface is shown in Fig. \ref{epc-fig}. We report the evolution of $\langle g^2\rangle$ for ${\rm G}^+$ and ${\rm G}^-$ at the ${\Gamma}$ point (Fig. \ref{epc-fig}a), meanwhile for the ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ points we analyze the TO, LO and LA branches (Fig. \ref{epc-fig}b). In the case of the LO branch at ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ we found that $\langle g^2\rangle$ remains practically constant for both ZZ and AC strain, and for clarity it has not been included in Fig. \ref{epc-fig}. \begin{figure} \centering \includegraphics*[scale=1]{fig-epc.pdf} \caption{Electron-phonon coupling $\langle g^2\rangle $ for strained graphene at (a) $\Gamma$ point, and (b) ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$.} \label{epc-fig} \end{figure} At the ${\Gamma}$ point, we found that after the splitting of the $E_{2g}$ phonon mode under uniaxial strain, the e-ph coupling in ${\rm G}^+$ (${\rm G}^-$) slightly increases (decreases) with almost no dependence on the strain direction. The overall change at ${\Gamma}$ for 5\% of uniaxial strain, considering the sum of both ${\rm G}^+$ and ${\rm G}^-$, shows a small reduction in $\langle g_{\Gamma}^2\rangle$ of 0.0017 eV$^2$, which corresponds to -2\% with respect to the value for the pristine case. A more complex behavior takes place in ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ as a function of uniaxial strain: the $\langle g^2\rangle$ in the TO branch starts to decrease with a non-linear dependence, while in the LA branch the e-ph coupling increases with almost the same rate as the corresponding one of the TO. In the same way as the phonon eigenvectors polarization behave under uniaxial strain, these changes occur faster in ZZ than AC, with the LA branch overcoming the TO after 3\% of ZZ strain. However, if we take into account the sum of all the branches in ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$, the total $\langle g_{{\bf q}}^2\rangle$ for ZZ and AC strains are very similar, and increases only by 2\% with respect to the pristine case. Therefore, considering both the ${\Gamma}$ and ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ contributions to the e-ph coupling, we have that the total $\langle g^2\rangle$ in uniaxially strained graphene remains practically constant. To understand the trend shown in Fig. \ref{epc-fig}, it is important to note that $\langle g^2\rangle \sim (\delta V)^2/\omega$ (see eqs. \ref{gelements} and \ref{g2}). Thus, considering only the contribution of $\omega$, due to the phonon softening of the Kohn anomalies, one would expected an increment in $\langle g^2\rangle$. However, from Fig. \ref{epc-fig} we can see that $\langle g^2\rangle$ decreases for some phonon modes as a function of strain. On the other hand, it should be noted that for ZZ and AC strain the pattern of the atomic vibrations in the $\Gamma$ point remains the same as in pristine graphene, even though the atomic distances $\alpha$ and $\beta$ are not equal. Meanwhile for ${\bf q}_{\rm ZZ}$ and ${\bf q}_{\rm AC}$, as a result of the Kohn anomaly shift from the high symmetry point ${\bf K}$, the mixing of the phonon eigenvectors for the TO and LA branches induces an important modification of the atomic vibration with respect to pristine graphene. Therefore, the behavior of $\langle g^2\rangle$ as function of uniaxial strain is mainly due to the change in the polarization of the phonon eigenvectors, which contributes to the enhancement(reduction) of the e-ph coupling in the LA(TO) branch. Regarding the anomaly in the phonon dispersion for the LA branch at ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ discussed in Sec. \ref{results-Kohn anomaly}, based on the substantial increment of $\langle g^2\rangle$, it could be assigned to an emergent Kohn anomaly in uniaxially strained graphene. This feature is a major difference in the e-ph coupling between pristine and uniaxially strained graphene, due to the presence of a new intervalley phonon-scattering channel for electronic states close to the Dirac point, in addition to the TO branch. On the other hand, it was previously reported that in comparison to many-body theories which includes electronic correlation effects, standard DFT underestimates the e-ph coupling of the $E_{2g}$ and $A_1'$ phonon modes of graphene.\cite{lazzeri2008} Therefore, the inclusion of such many-body effects could give rise to a stronger Kohn anomaly and e-ph coupling than our results. However, the use of linear response theory to compute the phonon dispersion and e-ph coupling are, at present, not implemented in many-body methodologies such as GW. Here we have shown that uniaxial strain induces a non-negligible Kohn anomaly even for small strain rates, which opens the possibility to be experimentally observed. It would be even more interesting to evaluate the contribution of this anomaly to those graphene properties which depend on the e-ph coupling. For example, it could be important to determine if this new Kohn anomaly contributes to the splitting of the double resonance Raman scattering 2D-band, \cite{mohr2010,Yoo2D,Venezuela2D,Marzari2D,Popov2D} or to the intrinsic electronic resistivity,\cite{sohier2014} where until now only the optical $A'_1$ intervalley phonon mode was considered. \section{Conclusions} \label{conclusions-sec} We have performed a first principles study of the structural properties, Kohn anomalies, and e-ph coupling for uniaxially strained graphene in the ZZ and AC directions. For ZZ strain we found a small contraction of the bond perpendicular to the strain direction, that increases the corresponding force constant. Evaluating the shift of the Dirac point from ${\bf K}$, the phonon nesting vectors ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ were calculated. Analyzing the phonon dispersion we found that a Kohn anomaly in ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$ emerges as a function of the uniaxial strain, in the LA branch. For both, the original Kohn anomaly in the TO branch and the new anomaly in the LA branch, there is a polarization of the phonon eigenvectors as induced by the strain, in directions parallel and perpendicular to the applied strain, in the same way as is known to occur for ${\rm G}^+$ and ${\rm G}^-$ at ${\Gamma}$. The softening of frequency in the Kohn anomaly shows a linear behavior as a function of the strain, except for the TO branch which shows a non-linear softening, and becomes almost constant for ZZ strains higher than 3\%. From the analysis of the average e-ph coupling matrix element square over the Fermi surface as a function of the uniaxial strain, for the ${\Gamma}$ point we found that the strain has a small effect on the ${\rm G}^+$ and ${\rm G}^-$ phonon modes. For the TO branch there is a reduction of $\langle g^2\rangle$ at ${\bf q}_{ZZ}$ and ${\bf q}_{AC}$, while for the LA branch there is a large enhancement of the e-ph coupling as a function of strain. Such behavior is mainly a consequence of the change in the polarization of the phonon eigenvectors because of mixing of the LA and TO modes induced by the uniaxial strain. Finally, it is important to emphasize that uniaxial strain in graphene induces a Kohn anomaly and enhancement of the e-ph coupling in the LA phonon branch, in contrast with the view commonly assumed that the e-ph coupling around the ${\bf K}$ point is present only in the TO phonon branch. \begin{acknowledgments} This research was partially supported by Consejo Nacional de Ciencia y Tecnolog\'ia (CONACYT, M\'exico) under grant No. 83604. One of the authors (M.E.C.-Q.) gratefully acknowledges a student grant from CONACYT-M\'exico, and the hospitality of the Karlsruher Institut f\"ur Technologie and the Benem\'erita Universidad Aut\'onoma de Puebla. \end{acknowledgments}
30c7d0ee8fc77feb0c458622502508f321cdcd6d
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} We consider a diffusion process in ${\mathbb R}^n$ solution of \[ dX_{t}=\sum_{j=1}^{d}\sigma _{j}(t,X_{t})\circ dW_{t}^{j}+b(t,X_{t})dt,\quad \quad X_0=x_0. \] where $W=(W^{1},...,W^{d})$ is a standard Brownian motion and $\circ dW_{t}^{j} $ denotes the Stratonovich integral. We assume suitable regularity properties for $\sigma_{j},b:{\mathbb R}^{+}\times {\mathbb R}^{n}\rightarrow {\mathbb R}^{n}$ (see \eqref{Not2} for details). We also assume that the coefficients $\sigma_{j},b$ verify the strong H\"{o}rmander condition of order one (that is, involving the $\sigma _{j}$'s and their first order Lie brackets $[\sigma _{i},\sigma_{j}]$'s) locally around a skeleton path \[ dx_{t}(\phi )=\sum_{j=1}^{d}\sigma _{j}(t,x_{t}(\phi ))\phi _{t}^{j}dt+b(t,x_{t}(\phi ))dt,\quad \quad x_0(\phi)=x_0 \] (this is formally written in property $(H_2)$ of \eqref{Not6}). In such a framework, in this paper we find exponential lower and upper bounds for the probability that the diffusion $X$ remains in a small tube around the skeleton path $x(\phi)$. Several works have considered this subject, starting from Stroock and Varadhan in \cite{StroockVaradhan:72}, where such result is used to prove the support theorem for diffusion processes. In their work, the tube is written in terms of the Euclidean norm, but later on different norms have been used to take into account the regularity of the trajectories (\cite{BenArousGradinaruLedoux:1994, FrizLyonsStrook:06}) and their geometric structure (\cite{pigato:14}). This kind of problems is also related to the Onsager-Machlup functional and large or moderate deviation theory, see e.g. \cite{capitaine:00,IkedaWatanabe:89,guillin2003}. In this work, we construct the tube using a distance coming from a norm which reflects the non isotropic structure of the problem, i.e. the fact that the diffusion process $X_{t}$ propagates with speed $\sqrt{t}$ in the direction of the diffusion vector fields $\sigma _{j}$ and with speed $t=\sqrt{t}\times \sqrt{t}$ in the direction of $[\sigma _{i},\sigma _{j}]$. We also prove that this distance is locally equivalent with the standard control (Carath\'eodory) metric. A key step in proving our tube estimates is given by the use of the density estimates provided in \cite{BCP1}. Generally speaking, there is a strong connection between tube and density estimates. In this work we use a concatenation of short time density estimates to prove a tube estimate, but one may proceed in reverse order: tubes estimates, for instance, can be used to provide lower bounds for the density. In \cite{[BFM]}, tube estimates for locally elliptic diffusions are proved, and applied to find lower bounds for the probability to be in a ball at fixed time and bounds for the distribution function. In \cite{BDM}, this is applied to lognormal-like stochastic volatility models, finding estimates for the tails of the distribution, and estimates on the implied volatility. The paper is organized as follows. In Section \ref{notationsresults}, we state our main result, given in Theorem \ref{mttubesS}, and we propose some examples of application. The proof of Theorem \ref{mttubesS} is developed in Section \ref{diffusion}. In Section \ref{sectioncontrol} we study the local equivalence between the control metric and the distance we use to define the tube when the diffusion coefficients depend on the space variable only. As a straightforward consequence, we can state our tube estimate result in terms of the Carath\'eodory metric (see Theorem \ref{tubecontrold}). \section{Notation and main results}\label{notationsresults} We recall the notation from \cite{BCP1} and introduce some new ones. We consider vector fields $\sigma_{j},b:{\mathbb R}^{+}\times {\mathbb R}^{n}\rightarrow {\mathbb R}^{n}$ which are four time differentiable in $x\in {\mathbb R}^{n}$ and one time differentiable in time $t\in {\mathbb R}^{+}$, and suppose that the derivatives with respect to the space $x\in {\mathbb R}^{n}$ are one time differentiable with respect to $t$. Hereafter, for $k\geq 1$, $\alpha =(\alpha _{1},\ldots,\alpha _{k})\in \{1,...,n\}^{k}$ represents a multi-index with length $|\alpha| =k$ and $% \partial _{x}^{\alpha }=\partial _{x_{\alpha _{1}}}\cdots\partial _{x_{\alpha _{k}}}$. We allow the case $k=0$ by setting $\alpha=\emptyset$ (the void multiindex), $|\alpha|=0$ and $\partial^\alpha_x=\mathrm{Id}$. For $(t,x)\in {\mathbb R}^{+}\times {\mathbb R}^{n}$ we denote by $n(t,x)$ a constant such that \be{Not2} \begin{array}{c} \forall s\in[ (t-1)\vee 0,t+1],\ \forall y\in B(x,1)\mbox{ one has }\smallskip\\ \displaystyle \sum_{|\alpha|=0}^4\Big(\left\vert \partial _{x}^{\alpha }b(s,y)\right\vert +\left\vert \partial _{t}\partial _{x}^{\alpha }b(s,y)\right\vert +\sum_{j=1}^{d}\left\vert \partial _{x}^{\alpha }\sigma _{j}(s,y)\right\vert+\left\vert \partial _{t}\partial _{x}^{\alpha }\sigma _{j}(s,y)\right\vert \Big) \leq n(t,x). \end{array} \end{equation} For $f,g:{\mathbb R}^{+}\times {\mathbb R}^{n}\rightarrow {\mathbb R}^{n}$ we define the directional derivative (w.r.t. the space variable $x$) $\partial_{g}f(t,x)=\sum_{i=1}^{n}g^{i}(t,x)\partial _{x_{i}}f(t,x)$, and we recall that the Lie bracket (again w.r.t. the space variable) is defined as $[g,f](t,x)=\partial _{g}f(t,x)-\partial _{f}g(t,x)$. Let $M\in \mathcal{M}_{n\times m}$ be a matrix with full row rank. We write $M^T$ for the transposed matrix, and $MM^T$ is invertible. We denote by $\lambda_*(M)$ (respectively $\lambda^*(M))$ the smallest (respectively the largest) singular value of $M$. We recall that singular values are the square roots of the eigenvalues of $MM^T$, and that, when $M$ is symmetric, singular values coincide with the absolute values of the eigenvalues of $M$. In particular, when $M$ is a covariance matrix, $\lambda_*(M)$ and $\lambda^*(M)$ coincide with the smallest and the largest eigenvalues of $M$. We consider the following norm on ${\mathbb R}^{n}$: \be{Not4} \left\vert y\right\vert _{M}=\sqrt{\left\langle (MM^T)^{-1}y,y\right\rangle}. \end{equation} We introduce the $n\times d^2$ matrix $A(t,x)$ defined as follows. We set $m=d^2$ and define the function \be{lip} l(i,p)=(p-1)d+i\in\{1,\dots,m\},\quad p,i\in\{1,\dots,d\}. \end{equation} Notice that $l(i,p)$ is invertible. For $l=1,\ldots,m$, we set the (column) vector field $A_{l}(t,x)$ in ${\mathbb R}^n$ as follows: \be{Al} \begin{split} A_{l}(t,x) &=[\sigma_{i},\sigma_{p}](t,x) \quad \mbox{if}\quad l=l(i,p) \quad \mbox{with}\quad i\neq p, \\ &= \sigma_{i}(t,x)\quad \mbox{if}\quad l=l(i,p) \quad \mbox{with}\quad i= p \end{split} \end{equation} and we set the $n\times m$ matrix $A(t,x)$ to be the one having $A_1(t,x),\ldots,A_m(t,x)$ as its columns, that is \be{Alucia} A(t,x)=[A_1(t,x),\ldots,A_m(t,x)]. \end{equation} We denote by $\lambda (t,x)$ the smallest singular value of $A(t,x)$, so \be{Not9'} \lambda (t,x)^2= \lambda_*(A(t,x))^2= \inf_{\left\vert \xi \right\vert =1}\sum_{i=1}^{m}\left\langle A_{i}(t,x),\xi \right\rangle ^{2}. \end{equation} For fixed $R>0$ we define the $m\times m$ diagonal scaling matrix $D_R$ as \begin{equation}\label{DR} \begin{split} (D_R)_{l,l}&=R \quad \mbox{if}\quad l=l(i,p) \quad \mbox{with}\quad i\neq p, \\ &= \sqrt{R} \quad \mbox{if}\quad l=l(i,p) \quad \mbox{with}\quad i= p \end{split} \end{equation} and the scaled directional matrix \be{Adlucia} A_R(t,x)=A(t,x)D_R. \end{equation} Notice that the $l$th column of the matrix $A_R(t,x)$ is given by $\sqrt{R}\sigma_{i}(t,x)$ if $l=l(i,p)$ with $ i= p$, and if $i\neq p$ then the $l$th column of $A_R(t,x)$ is $R[\sigma_{i},\sigma_{p}](t,x)=[\sqrt{R}\sigma_{i},\sqrt{R}\sigma_{p}](t,x)$. For a control $\phi \in L^2([0,T],{\mathbb R}^n)$ we consider the skeleton $x(\phi)$ associated to \eqref{equation}, that is, \be{skeleton} dx_{t}(\phi )=\sum_{j=1}^{d}\sigma _{j}(t,x_{t}(\phi ))\phi _{t}^{j}dt+b(t,x_{t}(\phi ))dt, \quad x_0(\phi)=x_0. \end{equation} In the following, we also need a function $R\,:\, [0,T]\rightarrow (0,1]$ that will play the role of a radius function (for the tube around $x(\phi)$). We consider now a ``regularity property'' already introduced in \cite{BallyKohatsu:10}, which is needed to control the growth of certain quantities along the skeleton path. For $\mu \geq 1$ and $0<h\leq 1$ we denote by $L(\mu ,h)$ the following class of functions: \begin{equation}\label{Lhmu} L(\mu,h)=\big\{f:{\mathbb R}^{+}\rightarrow {\mathbb R}^{+}\mbox{ such that }f(t)\leq \mu f(s)\quad \mbox{for}\quad \left\vert t-s\right\vert \leq h\big\}. \end{equation} From now on, we make use of the following hypotheses: there exist some functions $% n:[0,T]\rightarrow \lbrack 1,\infty )$ and $\lambda :[0,T]\rightarrow (0,1]$ such that for some $\mu \geq 1$ and $0<h\leq 1$ we have \begin{equation} \begin{array}{lcl} (H_{1})\quad & n(t,x_{t}(\phi )) \leq n_{t},\quad\forall t\in \lbrack 0,T],\smallskip \\ (H_{2})\quad & \lambda (t,x_{t}(\phi )) \geq \lambda _{t},\quad\forall t\in \lbrack 0,T],\smallskip \\ (H_{3})\quad & R_.,\,\left\vert \phi _{.}\right\vert ^{2},\,n_{.},\,\lambda _{.} \in L(\mu ,h). \end{array} \label{Not6} \end{equation} Recall that $\phi \in L^2([0,T],{\mathbb R}^n)$ is the control giving the skeleton path and $R\,:[0,T]\to (0,1]$ stands for the radius function. \br{dghd} Hypothesis $(H_{2})$ implies that for each $t\in (0,T),$ the space ${\mathbb R}^n$ is spanned by the vectors $(\sigma _{i}(t,x_{t}),[\sigma _{j},\sigma _{p}](t,x_{t}))_{i,j,p=1,...,d,j<p}$, meaning that a strong H\"{o}rmander condition locally holds along the curve $x_{t}(\phi ).$ \end{remark} Let $X$ denote a process in ${\mathbb R}^n$ solving \be{equation} dX_{t}=\sum_{j=1}^{d}\sigma _{j}(t,X_{t})\circ dW_{t}^{j}+b(t,X_{t})dt,\quad \quad X_0=x_0, \end{equation} $W$ being a standard Brownian motion in ${\mathbb R}^d$. Remark that $(H_1)$ is only a local assumption: we do not assume global Lipschitz continuity or sublinear growth properties for the coefficients, so the above SDE might not have a unique solution. We only assume to work with a continuous adapted process $X$ solving \eqref{equation} on the time interval $[0,T]$. For $K,q,K_*,q_*>0$, $\mu \geq 1$, $h\in(0, 1]$, $n\,:\,[0,T]\to [1,+\infty)$, $\lambda\,:\,[0,T]\to (0,1]$ and $\phi\in L^2([0,T],{\mathbb R}^n)$, we set the functions \begin{equation}\label{H-R} \begin{array}{l} H_t=K \left(\frac{\mu n_t}{\lambda_t}\right)^q,\smallskip\\ R_t^*(\phi)=\exp\left(-K_* \left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \mu^{2q_*}\right) \left( h\wedge \inf_{0\leq \delta\leq h} \left\{ \delta \big/ \int_t^{t+\delta} |\phi_s|^2 ds \right\} \right). \end{array} \end{equation} The main result of this paper is the following: \begin{theorem}\label{mttubesS} Let $\mu\geq 1$, $h\in(0,1]$, $n\,:\,[0,T]\to [1,+\infty)$, $\lambda\,:\,[0,T]\to (0,1]$, $R\,:\,[0,T]\to (0,1]$ and $\phi\in L^2([0,T],{\mathbb R}^n)$ be such that $(H_1)$--$(H_3)$ in \eqref{Not6} hold. Then there exist $K,q,K_*,q_*>0$ such that, for $H$ and $R^*(\phi)$ as in \eqref{H-R}, if $R_t\leq R_t^*(\phi)$ one has \begin{equation}\label{tuberesultstrong} \begin{array}{rl} \displaystyle \exp\left(- \int_0^T H_t \left(\frac{1}{R_t}+|\phi_t|^2\right)dt \right) \leq &\displaystyle {\mathbb P}\left(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \right)\\ &\displaystyle \leq \exp\left(- \int_0^T e^{-H_t} \left(\frac{1}{R_t}+|\phi_t|^2\right)dt \right). \end{array} \end{equation} \end{theorem} The proof of Theorem \ref{mttubesS} is developed in Section \ref{diffusion}. We discuss here some comments and examples. \br{1} The estimate \eqref{tuberesultstrong} allows for a regime shift, meaning that the dimension of the space generated by the $\sigma_i$'s and the $[\sigma_i,\sigma_j]$'s may change along the tube, and this is accounted by the variation of $A_R$ along $x_t(\phi)$. \end{remark} \br{4} The fact that $R\in L(\mu,h)$ implies that $\inf_{t\in [0,T]} R_t>0$. So, the radius of the tube is small, but cannot go to $0$ at any time. \end{remark} \begin{remark} The lower bound holds even if the inequality $R_t\leq R_t^*(\phi)$ is not satisfied, in the form \[ \exp\left(- \int_0^T H_t \left(\frac{1}{h}+\frac{1}{R_t}+|\phi_t|^2\right)dt \right) \leq{\mathbb P}\left(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \right). \] Details are given in next Theorem \ref{th-lower}. \end{remark} \br{jfgsjgt} Suppose $X_{t}=W_{t}$ and $x(\phi)=0$, so that $n_{t}=1$, $\lambda_{t}=1$, $\mu =1$ and $\phi _{t}=0$. Take $R_t=R$ constant. Then $% \left\vert X_{t}-x_{t}(\phi)\right\vert _{A_{R}(t,x_{t}(\phi ))}=R^{-1/2}W_{t}$ and we obtain $\exp (-C_1 T/R)\leq {\mathbb P}(\sup_{t\leq T}\left\vert W_{t}\right\vert \leq \sqrt{R})\leq \exp (-C_2 T/R)$ which is consistent with the standard estimate (see \cite{IkedaWatanabe:89}). \end{remark} A global two-sided bound for the density of $X_t$ is proved in \cite{KusuokaStroock:87}, under the \emph{strong} H\"{o}rmander non-degeneracy condition. It is also assumed that the coefficients do not depend time, i.e. $b(t,x)=b(x),\, \sigma(t,x)=\sigma(x)$, and that $b(x)=\sum_{j=1}^d \alpha_i \sigma_i(x)$, with $\alpha_i\in C_b^\infty({\mathbb R}^n)$ (i.e. the drift is generated by the vector fields of the diffusive part, which is a quite restrictive hypothesis). This bound is Gaussian in the control metric that we now define. For $x,y\in {\mathbb R}^{n}$ we denote by $C(x,y)$ the set of controls $\psi \in L^{2}([0,1];{\mathbb R}^d)$ such that the corresponding solution of \[ du_{t}(\psi )=\sum_{j=1}^{d}\sigma _{j}(u_{t}(\psi))\psi _{t}^{j}dt, \quad\quad u_{0}(\psi )=x \] satisfies $u_{1}(\psi )=y$. The control (Carath\'eodory) distance is defined as \begin{equation*} d_{c}(x,y)=\inf \Big\{\Big(\int_{0}^{1}\left\vert \psi _{s}\right\vert ^{2}ds% \Big)^{1/2}:\psi \in C(x,y)\Big\}. \end{equation*} The result in \cite{KusuokaStroock:87} is the following. Let $p_\delta(x,\cdot)$ denote the density of $X_\delta$ with starting condition $X_0=x$. Then there exists a constant $M\geq 1$ such that \[ \begin{split} &\frac{1}{M |B_{d_c}(x,\sqrt \delta)|}\exp\left( -\frac{M d_c(x,y)^2}{\delta}\right)\\ &\quad \quad \quad \leq p_\delta(x,y) \leq \frac{M}{|B_{d_c}(x,\sqrt\delta)|}\exp\left( -\frac{d_c(x,y)^2}{M \delta}\right) \end{split} \] where $\delta\in(0,T]$, $x,y\in{\mathbb R}^n$, $B_d(x,r)=\{y\in {\mathbb R}^n: d(x,y)<r\}$ and $|B_{d_c}(x,r)|$ denotes its Lebesgue measure. Remark that now, as in \cite{KusuokaStroock:87}, $\sigma(t,x)=\sigma(x)$. We define the semi distance $d$ via: $d(x,y) < \sqrt{R}$ if $|x-y|_{A_R(x)}<1$, and prove in Section \ref{sectioncontrol} the local equivalence of $d$ and $d_c$. This allows us to state Theorem \ref{mttubesS} in the control metric: \bt{tubecontrold} Suppose that the diffusion coefficients $\sigma_j$, $j=1,\ldots,d$, in \eqref{equation} depend on the space variable $x$ only and that the hypotheses of Theorem \ref{mttubesS} hold. Then, \begin{multline} \exp\left(- \int_0^T H_t \left(\frac{1}{R_t}+|\phi_t|^2\right)dt \right)\\ \leq {\mathbb P}\left(\sup_{0\leq t\leq T} d_c(X_t,x_t(\phi))\leq \sqrt{R_t} \right)\leq \exp\left(- \int_0^T e^{-H_t} \left(\frac{1}{R_t}+|\phi_t|^2\right)dt \right). \end{multline} \end{theorem} We prove the tube estimates in Section \ref{diffusion}, whereas the equivalence between the matrix norm and the Carath\'eodory distance is given in Section \ref{sectioncontrol}. \medskip We present now two examples of application. \medskip \textbf{Example 1}. {\small \textsc{[Grushin diffusion]}} Consider a positive, fixed $R$ and the two dimensional diffusion process \begin{equation*} X_{t}^{1}=x_{1}+W_{t}^{1},\quad X_{t}^{2}=x_{2}+\int_{0}^{t}X_{s}^{1}dW_{s}^{2}. \end{equation*}% Here $$ A_R A_R^T (x)=\left(\begin{array}{cc} R & 0 \\ 0 & R(x_1^2+2R) \end{array} \right), $$ so the associated norm is $|\xi|_{A_R(x)}^2= \frac{\xi_1^2}{R}+\frac{\xi_2^2}{R(x_1^2+2R)}$. On $\{x_1= 0\}$, $|\xi|_{A_R(x)}^2= \frac{\xi_1^2}{R}+\frac{\xi_2^2}{2R^2}$ and consequently $\{\xi :\left\vert \xi \right\vert _{A_{R }(x)}\leq 1\}$ is an ellipsoid. If we take a path $x(t)$ with $x_1(t)$ which keeps far from zero then we have ellipticity along the path and we may use estimates for elliptic SDEs (see \cite{[BFM]}). If $x_{1}(t)=0$ for some $t\in \lbrack 0,T]$ we need our estimate. Let us compare the norm in the two cases: if $x_{1}>0$ the diffusion matrix is non-degenerate and we can consider the norm $\left\vert \xi \right\vert _{B_{R}(x)}$ with $B_R(x)=R \sigma(x).$ We have% \begin{equation*} \left\vert \xi \right\vert _{B_{R }(x)}^{2}=\frac{1}{R }\xi _{1}^{2}+\frac{1}{R x_{1}^{2}}\xi _{2}^{2}\geq \frac{1}{R }\xi _{1}^{2}+\frac{1}{R (x_1^2+2R)}\xi _{2}^{2}=\left\vert \xi \right\vert _{A_{R }(x)}^{2}, \end{equation*}% and the two norms are equivalent for $R$ small. Let us now take $x_{t}(\phi )=(0,0).$ We have $n_{s}=1$ and $\lambda _{s}=1$ and $X_{t}-x_{t}(\phi)=(W_{t}^{1},% \int_{0}^{t}W_{s}^{1}dW_{s}^{2})$, so we obtain% \begin{equation*} \begin{split} e^{-C_1 T/R }&\leq {\mathbb P}\left(\sup_{t\leq T}\left\{ \frac{1}{R}\left\vert W_{t}^{1}\right\vert ^{2}+% \frac{1}{2R^{2}}\left\vert \int_{0}^{t}W_{s}^{1}dW_{s}^{2}\right\vert ^{2}\right\} \leq 1\right)\\ &={\mathbb P}\left(\sup_{t\leq T}(\left\vert X_{t}-x_{t}\right\vert _{A_R(x_t)}^{2}\leq 1 \right)\leq e^{-C_2 T/R }. \end{split} \end{equation*} \noindent \textbf{Example 2.} {\small \textsc{[Principal invariant diffusion on the Heisenberg group]}} Consider on ${\mathbb R}^3$ the vector fields $\partial_{x_1}-\frac{x_2}{2} \partial_{x_3}$ and $\partial_{x_2}-\frac{x_1}{2} \partial_{x_3}$. The associated Markov process is the triple given by a Brownian motion on ${\mathbb R}^2$ and its L\'evy area, that is \[ X_{t}^{1}=x_{1}+W_{t}^{1},\quad X_{t}^{2}=x_{2}+W_{t}^{2},\quad X_{t}^{3}=x_{3}+\frac{1}{2}\int_{0}^{t}X_{s}^{1}dW_{s}^{2}- \frac{1}{2} \int_{0}^{t}X_{s}^{2}dW_{s}^{1}. \] We refer e.g. to \cite{Driver2005340,Bakry20081905,Li2007497}, where gradient bounds for the heat kernel are obtained, and \cite{bau}. Since the diffusion is in dimension $n=3$ and the driving Brownian in dimension $d=2$, ellipticity cannot hold. Direct computations give \[ \sigma_1(x)= \left(\begin{array}{c} 1 \\ 0 \\ - \frac{x_2}2 \end{array}\right), \quad \sigma_2(x)= \left(\begin{array}{c} 0 \\ 1 \\ \frac{x_1}2 \end{array}\right), \quad [\sigma_1,\sigma_2](x)= \partial_{\sigma_1}\sigma_2- \partial_{\sigma_2}\sigma_1 = \left(\begin{array}{c} 0 \\ 0 \\ 1 \end{array}\right). \] Therefore $\sigma_1(x),\sigma_2(x),[\sigma_1,\sigma_2](x)$ span ${\mathbb R}^3$ and hypoellipticity holds. In $x=0$ we have $|\xi|^2_{A_R(0)}= \frac{\xi_1^2+\xi_2^2}{R} + \frac{\xi_3^2}{2 R^2}$, so taking the control $\phi\equiv 0$ and denoting $A_{t}(W)=\frac{1}{2}\int_{0}^{t}X_{s}^{1}dW_{s}^{2}- \frac{1}{2} \int_{0}^{t}X_{s}^{2}dW_{s}^{1}$ (the L\'evy area), we obtain \[ \begin{split} {\mathbb P}\left( \sup_{t\leq T/R} |W^1_t|^2+|W^2_t|^2 + \frac{|A_t(W)|^2}{2} \leq 1 \right) &= {\mathbb P}\left( \sup_{t\leq T} \frac{|W^1_t|^2+|W^2_t|^2}{R} + \frac{|A_t(W)|^2}{2 R^2} \leq 1 \right)\\ &= {\mathbb P}\left(\sup_{t\leq T} |X_t|^2_{A_R(x_t(\phi))}\leq 1 \right). \end{split} \] Appling our estimate we have \[ e^{-C_1 T/R}\leq {\mathbb P}\left( \sup_{t\leq T/R} |W^1_t|^2+|W^2_t|^2 + \frac{|A_t(W)|^2}{2} \leq 1 \right) \leq e^{-C_2 T/R}. \] \section{Tube estimates}\label{diffusion} The proof of Theorem \ref{mttubesS} is inspired by the approach in \cite{[BFM]}. A similar procedure is also used in \cite{pigato:14} in a weak H\"ormander framework. Such a proof strongly uses the estimates for the density developed in \cite{BCP1} and it is crucial that these estimates hold in a time interval of a fixed small length. This is because the proof consists in a ``concatenation'' of such estimates in order to recover the whole time interval $[0,T]$. And since the ``concatenation'' works around the skeleton path $x(\phi)$, it suffices that the properties for all objects hold only locally around $x(\phi)$, as required in \eqref{Not6}. In order to set-up this program, we need the precise behavior of the norm $|\cdot |_{A_R}$. So, we first present the desired properties for $|\cdot |_{A_R}$ (Section \ref{norms}) and then we proceed with the proof of Theorem \ref{mttubesS} (Section \ref{diffusion}). \subsection{Matrix norms}\label{norms} Recall the definitions \eqref{Alucia} and \eqref{Adlucia} for $A(t,x)$ and $A_R(t,x)$ respectively. We work with the norm $\left\vert y\right\vert _{A_R(t,x)}^{2}=\left\langle (A_RA^T_{R}(t,x))^{-1}y,y\right\rangle$, $y\in {\mathbb R}^n$. \label{page-norm1} \begin{lemma} \label{NORM1} Let $x\in{\mathbb R}^n$, $t\geq 0$, $R>0$ and recall that $\lambda^*(A(t,x))$ and $\lambda_*(A(t,x))$ denote the largest and lowest singular value of $A(t,x)$. \begin{itemize} \item[$i)$] For every $y\in {\mathbb R}^n$ and $0<R\leq R^{\prime }\leq 1$ \begin{align} \sqrt{\frac{R}{R^{\prime }}}\left\vert y\right\vert _{A_R(t,x)}&\geq \left\vert y\right\vert _{A_{R^{\prime }}(t,x)}\geq \frac{R}{R^{\prime }}\left\vert y\right\vert _{A_R(t,x)}\label{Norm2} \\ \frac{1}{\sqrt{R} \lambda ^*(A(t,x))}\left\vert y\right\vert &\leq \left\vert y\right\vert _{A_R(t,x)}\leq \frac{1}{R\lambda _*(A(t,x))}% \left\vert y\right\vert . \label{Norm3} \end{align} \item[$ii)$] For every $z\in {\mathbb R}^{m}$ and $R>0$ \begin{equation} \left\vert A_R(t,x)z\right\vert _{A_R(t,x)}\leq \left\vert z\right\vert . \label{Norm4} \end{equation} \item[$iii)$] For every $\varphi\in L^2([0,T];{\mathbb R}^m)$, \begin{equation} \Big|\int_0^r\varphi_s\,ds\Big|^2_{A_R(t,x)} \leq r\int_0^r|\varphi_s|^2_{A_R(t,x)}\,ds,\quad r\in [0,T]. \label{Norm4bis} \end{equation} \end{itemize} \end{lemma} \begin{proof} For fixed $x\in{\mathbb R}^n$ and $t\geq 0$, during the proof we omit in $A(t,x)$ and $A_R(t,x)$ the dependence on $(t,x)$, so we simply write $A$ and $A_R$ \smallskip $i)$ For $0<R\leq R^{\prime }\leq 1$, it is easy to check that \begin{equation*} \frac{R^{\prime }}{R}A_RA_R^T\leq A_{R^{\prime }}A_{R^{\prime }}^T\leq \left( \frac{R^{\prime }}{R}\right) ^{2}A_RA_R^T \end{equation*}% which is equivalent to (\ref{Norm2}). This also implies (taking $ R^{\prime }=1$ so $A_{R^{\prime }}=A)$ that% \begin{align*} &\frac{1}{R}\lambda _*(A_R)^2\leq \lambda _*(A)^2\leq \frac{1}{R^{2}% }\lambda _*(A_R)^2\\ &\frac{1}{R}\lambda ^*(A_R)^2 \leq \lambda ^*(A)^2\leq \frac{1}{R^{2}}\lambda ^*(A_R)^2 \end{align*}% which immediately gives (\ref{Norm3}). $ii)$ For $z\in {\mathbb R}^m$, we write $z=A_R^T y+w$ with $y\in {\mathbb R}^{n}$ and $% w\in (\mathrm{Im}A_R^T)^{\bot }= \mathrm{Ker}\, A_R.$ Then $A_Rz=A_RA_R^T y$ so that% \begin{eqnarray*} \left\vert A_Rz\right\vert _{A_R}^{2} &=&\left\vert A_RA_R^T y\right\vert _{A_R}^{2}=\left\langle (A_RA_R^T)^{-1}A_RA_R^T y,A_RA_R^T y\right\rangle \\ &=&\left\langle y,A_RA_R^T y\right\rangle=\left\langle A_R^T y,A_R^T y\right\rangle =\left\vert A_R^T y\right\vert ^{2}\leq \left\vert z\right\vert ^{2} \end{eqnarray*} and (\ref{Norm4}) holds. \smallskip $iii)$ For $\varphi\in L^2([0,T];{\mathbb R}^m)$ and $r\in [0,T]$, \begin{align*} &\Big|\int_0^r\varphi_sds\Big|^2_{A_R} =\big\langle(A_RA_R^T)^{-1}\int_0^r\varphi_sds,% \int_0^r\varphi_sds\big\rangle =\int_0^r\int_0^r\big\langle(A_RA_R^T)^{-1}\varphi_s,\varphi_u% \big\rangle dsdu \\ &=\frac 12\int_0^r\int_0^r\big\langle(A_RA_R^T)^{-1}(\varphi_s-\varphi_u),\varphi_s-\varphi_u% \big\rangle dsdu\\ &\quad-\int_0^r\int_0^r\Big(\big\langle(A_RA_R^T)^{-1}\varphi_s,\varphi_s\big\rangle -\big\langle(A_RA_R^T)% ^{-1}\varphi_u,\varphi_u\big\rangle \Big)dsdu \\ &=\frac 12\int_0^r\int_0^r\Big(|\varphi_s-\varphi_u|^2_{{A_R}}-2|\varphi_s|^2_{{A_R}}% \Big)dsdu \\ &\leq \int_0^r\int_0^r|\varphi_u|^2_{A_R}dsdu = t\int_0^t|\varphi_u|^2_{A_R}du. \end{align*} \end{proof} Next Lemma \ref{NORM3} is strictly connected to Remark \ref{1}, where we stressed that our result allows for a regime switch along the tube. In fact, here we fix $R>0$, two points $(t,x)$ and $(s,y)$ and we get an equivalence between the norms $|\cdot|_{A_{R }(t,x)}$ and $|\cdot|_{A_{R }(s,y)}$ without supposing that in these two points the H\"ormander condition holds ``under the same regime''. To compensate this lack of uniformity, we suppose that the distance between $(t,x)$ and $(s,y)$ is bounded by $\sqrt{R}$, and we will need to take this fact into account. In the concatenation procedure of next Section \ref{sect-proof}, the size of the intervals, to which we apply our density estimates, will have to depend on the radius of the tube. We set $$ O=\{(t,x)\in [0,T]\times {\mathbb R}\,:\,\lambda(t,x)>0\} $$ which is open, and under \eqref{Not2}, we define \begin{equation}\label{DD} \mathcal{D}=\Big\{C\,:\,O\to {\mathbb R}_+ \mbox{ such that } C(t,x)=K\Big(\frac{n(t,x)}{\lambda(t,x)}\Big)^q,\ K,q>0\Big\}. \end{equation} We also define $$ 1/\mathcal{D}=\Big\{c\,:\,O\to {\mathbb R}_+ \mbox{ such that } 1/c\in\mathcal{D}\}. $$ \bl{NORM3} Assume \eqref{Not2} and let $\mathcal{D}$ as in \eqref{DD}. There exists $C^*\in \mathcal{D}$ such that for every $(t,x),(s,y)\in O$ and $R\in(0,1]$ satisfying \be{Norm5} \left\vert x-y\right\vert +|t-s| \leq \sqrt{R}/C^*(t,x), \end{equation} then for every $z\in {\mathbb R}^{n}$ one has \be{Norm6} \frac{1}{4}\left\vert z\right\vert^2 _{A_{R }(t,x)}\leq \left\vert z\right\vert^2 _{A_{R }(s,y)}\leq 4\left\vert z\right\vert^2 _{A_{R }(t,x)}. \end{equation} \end{lemma} \begin{proof} \eqref{Norm6} is equivalent to \[ 4(A_{R }A_{R }^T)(t,x)\geq (A_{R }A_{R }^T)(s,y)\geq \frac{1}{4}(A_{R }A_{R }^T)(t,x), \] so we prove the above inequalities. Let $A_{R,k}$, $k=1,\ldots,m$, denote the columns of $A_R$. We use $(a+b)^{2}\geq \frac{1}{2}a^{2}-b^{2}$: \begin{eqnarray*} \langle A_{R }A_{R }^T(s,y)z,z\rangle &=&\sum_{k=1}^{m}\left\langle A_{R ,k}(s,y),z\right\rangle ^{2} \\ &=&\sum_{k=1}^{m}(\left\langle A_{R ,k}(t,x),z\right\rangle +\left\langle A_{R ,k}(s,y)-A_{R ,k}(t,x),z\right\rangle )^{2} \\ &\geq &\frac{1}{2}\sum_{k=1}^{m}\left\langle A_{R ,k}(t,x),z\right\rangle ^{2}-\sum_{k=1}^{m}\left\langle A_{R ,k}(s,y)-A_{R ,k}(t,x),z\right\rangle^{2}. \end{eqnarray*}% We use \eqref{Not2}: for every $(s,y)$ such that $|t-s|\leq 1$ and $|x-y|\leq 1$, we have \begin{eqnarray*} \langle A_{R }A_{R }^T(s,y)z,z\rangle &\geq &\frac{1}{2}\sum_{k=1}^{m}\left\langle A_{R ,k}(t,x),z\right\rangle ^{2}-C_1 n(t,x)^\alpha R (\left\vert x-y\right\vert ^{2}+\left\vert t-s\right\vert ^{2}) \left\vert z\right\vert ^{2}, \end{eqnarray*}% in which $C_1>0$ and $\alpha\geq 1$ denote universal constants. Notice that \[ \sum_{k=1}^{m}\left\langle A_{R,k}(t,x),z\right\rangle ^{2} = \langle A_{R }A_{R }^T(t,x)z,z\rangle \geq\lambda_*^2(A_{R}(t,x)) |z|^2 \geq R^2 \lambda_*^2(A(t,x))|z|^2. \] We choose the constants $(K,q)$ characterizing $C^*(t,x)$ such that $K\geq 2\sqrt{C_1}\vee 1$ and $q\geq \alpha$. So, under \eqref{Norm5} we obtain \[ C_1 n(t,x)^\alpha R (\left\vert x-y\right\vert ^{2}+\left\vert t-s\right\vert ^{2}) \left\vert z\right\vert ^{2} \leq \frac{1}{4}\sum_{k=1}^{m}\left\langle A_{R ,k}(t,x),z\right\rangle ^{2} \] and \[ \left\langle (A_{R }A_{R }^T)(s,y)z,z\right\rangle \geq \frac{1}{4}\sum_{k=1}^{m}\left\langle A_{R ,k}(t,x),z\right\rangle ^{2}=% \frac{1}{4}\left\langle (A_{R }A_{R }^T)(t,x)z,z\right\rangle . \] The converse inequality follows from analogous computations and inequality $(a+b)^{2}\leq 2a^{2}+2b^{2}$. \end{proof} We prove that moving along the skeleton associated to a control $\phi\in L^2([0,T],{\mathbb R}^d)$ for a small time $\delta$, the trajectory remains close to the initial point in the $A_\delta$-norm. To this purpose, we assume the conditions $(H_1)$ and $(H_2)$ in \eqref{Not6}. Notice that these give $(t,x_t(\phi))\in O$ for every $t$. Moreover, in such a case the set ${\mathcal{D}}$ can be replaced by the following class of functions: \begin{equation}\label{AA} \mathcal{A}=\Big\{C\,:\,[0,T]\to {\mathbb R}_+\,:\, C_t=K\Big(\frac{n_t}{\lambda_t}\Big)^q,\mbox{ for some $K,q>0$}\Big\}, \end{equation} $n_t$ and $\lambda_t$ being defined in \eqref{Not6}. We also set $$ 1/\mathcal{A}=\big\{c\,:\,[0,T]\to (0,1]\,:\, 1/c_t\in \mathcal{A}\big\}. $$ \bl{normcontrol} Let $x(\phi)$ be the skeleton path \eqref{skeleton} associated to $\phi\in L^2([0,T],{\mathbb R}^d)$. Assume $(H_1)$ and $(H_2)$ in \eqref{Not6}. Then there exists $\delta^*,\varepsilon^*\in 1/\mathcal{A}$ such that for every $t\in [0,T]$, $\delta_t\leq \delta^*_t$, $\varepsilon_t(\delta_t)\leq \varepsilon^*_t$, $ s\in[0, \delta_t]$ with $t+s\leq T$ and for every $z\in{\mathbb R}^n$ one has \be{correction3} \frac 14\left\vert z\right\vert^2 _{A_{\delta_t}(t,x_{t}(\phi))}\leq \left\vert z\right\vert^2 _{A_{\delta_t}(t+s,x_{t+s}(\phi))}\leq 4\left\vert z\right\vert^2 _{A_{\delta_t}(t,x_{t}(\phi))}. \end{equation} Moreover, there exists $\bar{C}\in \mathcal{A}$ such that \be{normcontrolhat} \sup_{0\leq s \leq \delta_t} |x_{t+s}(\phi)-(x_t(\phi)+b(t,x_t(\phi))s)|_{A_{\delta_t}(t,x_t(\phi))}\leq \bar{C}_t (\varepsilon_t(\delta_t)\vee \sqrt{\delta_t}) \end{equation} where \[ \varepsilon_t(\delta)=\left(\int_t^{t+\delta} |\phi_s|^2ds\right)^{1/2}. \] \end{lemma} \begin{proof} Set $s_t=\inf\{s>0\,:\,|x_{t+s}(\phi)-x_t(\phi)|\geq 1\}$. From \eqref{Not2} and $(H_1)$ in \eqref{Not6}, we have \begin{align*} 1=|x_{t+s_t}(\phi)-x_t(\phi)| &\leq n_t\big(s_t+\sqrt {s_t}\varepsilon_t(s_t)\big). \end{align*} We take $\underline{C}\in \mathcal{A}$ such that $n_t\big(\sqrt{s_t}+\varepsilon_t(s_t)\big)\leq \underline{C}_t^{1/2}$, so that $s_t\geq 1/\underline{C}_t$. Take now $\delta^*\in 1/\mathcal{A}$ such that $\delta^*\leq 1/\underline{C}$. Then if $s\leq \delta_t\leq \delta^*_t$, one has $s\leq s_t$ and again from \eqref{Not2} and $(H_1)$ in \eqref{Not6} we have $$ |x_{t+s}(\phi)-x_t(\phi)|+|s| \leq \sqrt{\delta_t}\big(n_t(\sqrt{\delta^*_t}\, +\varepsilon_t(\delta_t))+\sqrt{\delta^*_t}\big). $$ By continuity, for every $\varepsilon^*\in1/\mathcal{A}$ and for every $t$ there exists $\hat\delta_t$ such that $\varepsilon_t(\hat \delta_t)\leq \varepsilon^*_t$. So, there actually exists $\delta_t\leq \delta^*_t$ for which $\varepsilon_t(\delta_t)\leq \varepsilon^*_t$. And for such a $\delta_t$, we have $$ |x_{t+s}(\phi)-x_t(\phi)|+|s| \leq \sqrt{\delta_t}\big(n_t(\sqrt{\delta^*_t}\, +\varepsilon^*_t)+\sqrt{\delta^*_t}\big). $$ We now choose $\delta^*,\varepsilon^*\in 1/\mathcal{A}$ in order that the last factor in the above right hand side is smaller than $1/C^*(t,x_t(\phi))$, where $C^*(t,x)$ is the function in $\mathcal{D}$ for which Lemma \ref{Norm3} holds. Then \eqref{Norm5} is satisfied with $R=\delta_t$, $x=x_t(\phi)$, $y=x_{t+s}(\phi)$ and $s$ replaced by $t+s$. Hence \eqref{correction3} follows by applying \eqref{Norm6}. \smallskip We prove now \eqref{normcontrolhat}. For the sake of simplicity, we let $x_t$ denote the skeleton path $x_t(\phi)$. We write \begin{align*} J_{t,s}&:=x_{t+s}-x_{t}-b(t,x_{t})s=\int_{t}^{t+s}(\dot x_{u}-b(u,x_{u}))du +\int_{t}^{t+s}(b(u,x_{u})-b(t,x_{t}))du\\ &=\int_{t}^{t+s}\sigma(u,x_u)\phi_udu +\int_{t}^{t+s}(b(u,x_{u})-b(t,x_{t}))ds, \end{align*} so that \begin{align*} |J_{t,s}|^2_{A_{\delta_t}(t,x_t)} &\leq 2s\int_t^{t+s}|\sigma(u,x_u)\phi_u|^2_{A_{\delta_t}(t,x_t)}dt +2s\int_t^{t+s}|b(u,x_{u})-b(t,x_t)|^2_{A_{\delta_t}(t,x_t)}du. \end{align*} In the above right hand side, we apply \eqref{correction3} to the norm in the first term and we use \eqref{Norm3} in the second one. And we obtain: \begin{align*} |J_{t,s}|^2_{A_{\delta_t}(t,x_t)} &\leq 2s\int_t^{t+s} 4|\sigma(u,x_u)\phi_u|^2_{A_{\delta_t}(u,x_{u})}du +2s\int_t^{t+s}\frac{1}{\delta_t^2\lambda_t^2}|b(u,x_u)-b(t,x_t)|^2 du\\ &\leq 8s\int_t^{t+s} |\sigma(u,x_u)\phi_u|^2_{A_{\delta_t}(u,x_u)}du +2\delta_t\int_t^{t+\delta_t}\frac{1}{\delta_t^2\lambda_t^2}\times n_t^2(|u-t|+ |x_u-x_t|)^2 du. \end{align*} We have already proved that, for $u\in[t,t+s]$, $|u-t|+ |x_u-x_t|\leq \sqrt{\delta_t}/C^*_t$, with $C^*\in \mathcal{A}$, so \begin{align*} |J_{t,s}|^2_{A_{\delta_t}(t,x_t)} &\leq 8s\int_t^{t+s} |\sigma(u,x_u)\phi_u|^2_{A_\delta(u,x_u)}du +\bar C_t\delta_t, \end{align*} with $\bar C\in \mathcal{A}$. It remains to study the first term in the above right hand side. For $i=1,\dots,m$, we set $\psi^{(j-1)d+j}=\frac{1}{\sqrt{\delta_t }}\phi^{j}$ for $j=1,\dots,d$, $\psi^i=0$ otherwise. Then, recalling \eqref{Adlucia}, we can write $\sigma(u,x_{u})\phi_u=A_{\delta_t}(u,x_{u}(\phi)) \psi_u$, so that, by \eqref{Norm4}, \[ |\sigma(u,x_u)\phi_u|^2_{A_{\delta_t}(u,x_u)} =|A_{\delta_t}(u,x_u)\psi_u|^2_{A_{\delta_t}(u,x_u)} \leq |\psi_u|^2=\frac 1{\delta_t}\,|\phi_u|^2. \] Hence, for $s\leq \delta_t$, we finally have $|J_{t,s}|^2_{A_{\delta_t}(t,x_t)}\leq 8\varepsilon_t(\delta_t)^2+\bar C_t\delta_t$, and the statement follows. \end{proof} \begin{remark}\label{for-dc} Let us finally discuss an inequality which will be used in next Section \ref{sectioncontrol}. Fix $x\in{\mathbb R}^n$ and let $x(\phi)$ be the skeleton path \eqref{skeleton} associated to $\phi\in L^2([0,T],{\mathbb R}^d)$ with starting condition $x_0(\phi)=x$. Assume simply \eqref{Not2} and recall $\mathcal{D}$ defined in \eqref{DD}. Then looking at the proof of Lemma \eqref{normcontrol}, we have the following result: if $(0,x)\in O$, there exists $\overline{\delta}, \overline{\varepsilon}\in 1/\mathcal{D}$ and $\overline{C}\in\mathcal{D}$ such that if $\delta\leq \overline{\delta}(0,x)$, $\varepsilon_0(\delta)\leq \overline{\varepsilon}(0,x)$ and $s\in[0, \delta]$ then \be{normcontrolhat2} \sup_{0\leq s \leq \delta} |x_{s}(\phi)-(x+b(0,x)s)|_{A_{\delta}(0,x)}\leq \overline{C}(0,x) (\varepsilon_0(\delta)\vee \sqrt{\delta}). \end{equation} \end{remark} \subsection{Proof of Theorem \ref{mttubesS}}\label{sect-proof} This section is organized as follows: the lower bound in Theorem \ref{mttubesS} is proved in next Theorem \ref{th-lower}, whereas the upper bound in Theorem \ref{mttubesS} is studied in next Theorem \ref{th-upper}. As already mentioned, the proof we are going to develop relies on a two-sided bound for the density of equation \eqref{equation} in short time, proved in \cite{BCP1}. The estimate is \emph{diagonal}, meaning that it is local around the drifted initial condition $x_0+b(0,x_0)\delta$, $\delta$ denoting the (small) time at which we are studying the density. But in order to be more precise and self-contained, we briefly recall the result from \cite{BCP1} we are going to strongly use. We will suppose that \begin{equation}\label{hpder} \sum_{0\leq |\alpha| \leq 4} \Big[ \sum_{j=1}^d |\partial_x^{\alpha} \sigma_j(t,x)| + |\partial_x^{\alpha} b(t,x)| + |\partial_x^{\alpha}\partial_t\sigma_j(t,x)| \Big] \leq \kappa, \ \forall t\in [0,T],\,\forall x \in {\mathbb R}^n. \end{equation} Of course \eqref{hpder} is much stronger than \eqref{Not2} but we will see in the sequel that, by a suitable localization, one can reduce to the validity of \eqref{hpder} (see next Remark \ref{2}). We also assume that \begin{equation}\label{ND} \lambda (0,x_0)>0, \end{equation} $x_0$ denoting the starting point of the diffusion $X$ solving \eqref{equation}, and we consider the following set of constants: \begin{equation}\label{CC} \mathcal{D}_0=\Big\{C>0\,:\,C=K \Big(\frac{\kappa}{\lambda(0,x_0)}\Big)^{q},\ \exists\ K,q>0\Big\}. \end{equation} We use the notation $1/\mathcal{D}_0$ for constants $c$ such that $1/c\in\mathcal{D}_0$. We set $p_\delta(x_0,\cdot)$ the density of $X_\delta$ when $X_0=x_0$. We use here the following version of Theorem \ref{m} in \cite{BCP1}: \bt{m} Suppose that \eqref{ND} and \eqref{hpder} hold. Let ${\mathcal{D}}_0$ be defined in \eqref{CC}. Then there exist $r^*,\delta^*\in 1/{\mathcal{D}}_0$, $C\in {\mathcal{D}}_0$ such that for $\delta\leq \delta^*$ and for $|y-x_0-b(0,x_0)\delta|_{A_\delta(0,x_0)} \leq r^*$ one has \[ \frac{1}{ C \delta^{n-\frac{\mathrm{dim}\,\langle \sigma(0,x_0) \rangle}{2} }} \leq p_\delta(x_0,y) \leq \frac{e^C}{ \delta^{n-\frac{\mathrm{dim}\,\langle \sigma(0,x_0) \rangle}{2} }} \] where $\mathrm{dim}\,\langle \sigma(0,x_0) \rangle$ denotes the dimension of the vector space spanned by $\sigma_1(0,x_0),\ldots,$ $\sigma_d(0,x_0)$. \end{theorem} Notice that \eqref{ND} and \eqref{hpder} are, respectively, Assumption 2.2 and Assumption 2.3 in \cite{BCP1}. Therefore, Theorem \ref{m} is actually a re-writing of Theorem 3.7 in \cite{BCP1} (with the constant $C$ specified in Remark 3.8 therein) and Theorem 4.6 in \cite{BCP1}. \begin{remark}\label{est-tx} Of course, Theorem \ref{m} can be written for a general starting condition $(t,x)$ in place of $(0,x_0)$. In such a case, \eqref{ND} and \eqref{CC} have to be replaced by $$ \lambda (t,x)>0\quad\mbox{and}\quad \mathcal{D}_{t,x}=\Big\{C>0\,:\,C=K \Big(\frac{\kappa}{\lambda(t,x)}\Big)^{q},\ \exists\ c,q>0\Big\} $$ respectively. But a closer look to the proofs of Theorem 3.7 and of Theorem 4.6 in \cite{BCP1} shows that the constants $K$ and $q$ in $\mathcal{D}_0$ are universal, that is, they can be taken independently of all the data (the starting point $(0,x_0)$, the diffusion coefficients, the quantities $\lambda (0,x_0)$, $\kappa$ etc.). This means that Theorem \ref{m} can be formulated as follows. Assume that \eqref{hpder} holds and define the (open) set $$ O=\big\{(t,x)\,:\,\lambda(t,x)>0\big\}. $$ Set $$ \mathcal{D}=\Big\{C:O\to{\mathbb R}_+\,:\,C(t,x)=K \Big(\frac{\kappa}{\lambda(t,x)}\Big)^{q},\ \exists\ c,q>0\Big\}. $$ Then there exist $C\in {\mathcal{D}}$, $r^*,\delta^*\in 1/{\mathcal{D}}$ such that for $(t,x)\in O$, $\delta\leq \delta^*(t,x)$ and for every $y$ such that $|y-x-b(t,x)\delta|_{A_{\delta}(t,x)} \leq r^*_t$ one has $$ \frac{1}{ C(t,x) \delta^{n-\frac{\mathrm{dim}\,\langle \sigma(t,x) \rangle}{2} }} \leq p(t,t+\delta,x,y) \leq \frac{e^{C(t,x)}}{ \delta^{n-\frac{\mathrm{dim}\,\langle \sigma(t,x) \rangle}{2} }}, $$ where $p(t,s,x,\cdot)$ denotes the density of the solution $X$ at time $s$ of the equation in \eqref{equation} but with the starting condition $X_t=x$. \end{remark} \begin{remark}\label{C-B} From \eqref{Adlucia} and the Cauchy-Binet formula we obtain (for details see (3.43) in \cite{BCP1}) \be{CB} \frac{1}{C(t,x)} \delta^{n-\frac{\mathrm{dim}\,\langle \sigma(t,x) \rangle}{2}} \leq \sqrt{\det A_\delta A_\delta^T (t,x)}\leq C(t,x) \delta^{n-\frac{\mathrm{dim}\,\langle \sigma(t,x) \rangle}{2}}, \end{equation} so the density bounds above are equivalent to the following ones: \be{denco} \frac{1}{ C(t,x) \sqrt{\det A_{\delta} A_{\delta}^T (t,x)} } \leq p(t,t+\delta,x,y) \leq \frac{e^{C(t,x)}}{ \sqrt{\det A_{\delta} A_{\delta}^T (t,x)}} \end{equation} \end{remark} \br{2} The plan for the proof is the following. Consider first the lower bound (see Theorem \ref{th-lower}). For $\phi\in L^2[0,T]$, let $x(\phi)$ be the skeleton associated to \eqref{equation} is given in \eqref{skeleton}. We set a discretization $0=t_0<t_1<\cdots <t_N=T$ of the time interval $[0,T]$. Then, as $k$ varies, we consider the events \begin{equation}\label{D-Gamma} D_k=\Big\{ \sup_{t_k\leq t \leq t_{k+1}} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \Big\} \mbox{ and } \Gamma_k = \Big\{ y\,:\,|y-x_{t_k}(\phi)|_{A_{R_{t_k}}(t_k,x_{t_k}(\phi))} \leq r_k \Big\}, \end{equation} where $r_k<1$ is a radius that will be suitably defined in the sequel. We denote ${\mathbb P}_k$ the conditional probability \[ {\mathbb P}_k(\cdot)={\mathbb P}\left( \cdot|W_t,t\leq t_k; X_{t_k}\in \Gamma_k \right) \] We will lower bound ${\mathbb P}(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 )$ by computing the product of the probabilities ${\mathbb P}_k \left(D_k\cap \{X_{t_k+1}\in\Gamma_{k+1}\}\right) $, and this computation uses the lower estimate of the densities given in Theorem \ref{m}. Remark that Theorem \ref{m} uses \eqref{hpder}, a condition which asks for a global bound for the derivatives of the coefficients, whereas for the tube estimates we are assuming only $(H_1)$ in \eqref{Not6}, i.e. a bound for the coefficients which is not global but just in a neighborhood of the skeleton. Suppose that we have a process $X$ which, for some external reasons, verifies \eqref{equation} for $t_k\leq t \leq t_{k+1}$, and such that $\sup_{t_k\leq t \leq t_{k+1}} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 $. From $(H_1)$, $n_{t_k}$ bounds the derivatives of $\sigma(t,y)$ and $b(t,y)$ for all $(t_k-1)\vee 0\leq t \leq t_k+1$, and for all $|y-x_{t_k}(\phi)|\leq 1$. Then, for example using the result in \cite{whitney1944}, we can define $\bar{\sigma},\bar{b}$ which coincide with $\sigma,b$ on $[(t_k-1)\vee 0, t_k+1]\times \{y\in {\mathbb R}^n: |y-x_{t_k}(\phi)|\leq 1\}$, are differentiable as many times as $\sigma,b$ but the bound in \eqref{hpder} holds on the whole ${\mathbb R}^+\times {\mathbb R}^n$. Let now $\bar{X}$ be the strong solution to \[ \bar{X}_t=X_{t_k}+\int_{t_k}^t \bar{\sigma}(s,\bar{X}_s)\circ dW_s + \int_{t_k}^t \bar{b}(s,\bar{X}_s) ds, \quad t\in[t_k,t_{k+1}]. \] Now, if we call $\bar D_k$ the sets in \eqref{D-Gamma} with $X$ replaced by $\bar X$, it is clear that \[ {\mathbb P} ( D_k \cap \{X_{t_k+1}\in\Gamma_{k+1}\})= {\mathbb P} ( \bar D_k \cap \{\bar{X}_{t_k+1}\in\Gamma_{k+1}\}) \] and therefore we can equivalently prove our tube estimate supposing that the bound in $(H_1)$ holds globally, that is assuming \eqref{hpder}. This really allows us to apply Theorem \ref{m}. And a similar procedure can be developed for the upper bound (see Theorem \ref{th-upper}). \end{remark} We recall the set $\mathcal{A}$ defined in \eqref{AA}: $$ \mathcal{A}=\Big\{C\,:\,[0,T]\to {\mathbb R}_+\,:\, C_t=K\Big(\frac{n_t}{\lambda_t}\Big)^q,\mbox{ for some $K, q>0$}\Big\}. $$ We also recall $1/\mathcal{A}$ defined as usual. Notice that that, under \eqref{Not6}, $n(t,x_t(\phi))\leq n_t$ and $\lambda(t,x_t(\phi))\geq \lambda_t$. So, any $C(t,x)\in\mathcal{D}$ evaluated in $(t,x_t(\phi))$ is upper bounded by the function $C_t$ in $\mathcal{A}$ written with the same constants $K$ and $q$. For $\mu\geq 1$, $h\in(0,1]$ and $K_*,q_*>0$, we denote \be{Rmax} R_t^*(\phi)=\exp\left(-K_* \left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \mu^{2q_*}\right) \left( h\wedge \inf_{0\leq \delta\leq h} \left\{ \delta \big/ \int_t^{t+\delta} |\phi_s|^2 ds \right\} \right) \end{equation} \begin{theorem}\label{th-lower} Let $\mu\geq 1$, $h\in (0,1]$, $n\,:\,[0,T]\to [1,+\infty)$, $\lambda\,:\,[0,T]\to (0,1]$, $\phi\in L^2([0,T],{\mathbb R}^n)$ and $R\,:\,[0,T]\to (0,1]$ be such that $(H_1)$--$(H_3)$ in \eqref{Not6} hold. Then there exist $\bar K,\bar q>0$ such that \begin{equation}\label{tube-lower} {\mathbb P}\left(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \right) \geq \exp\left(- \int_0^T \bar{K} \left(\frac{\mu n_t}{\lambda_t}\right)^{\bar{q}}\left( \frac 1h+\frac{1}{R_t}+|\phi_t|^2 dt\right) \right). \end{equation} Moreover, if $R_t \leq R_t ^*(\phi)$ for some $K_*,q_*>0$, $R^*(\phi)$ being given in \eqref{Rmax}, then \begin{equation}\label{tube-lower-bis} {\mathbb P}\left(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \right) \geq \exp\left(- \int_0^T 2\bar{K} \left(\frac{\mu n_t}{\lambda_t}\right)^{\bar{q}}\left( \frac{1}{R_t}+|\phi_t|^2 dt\right) \right). \end{equation} \end{theorem} \begin{proof} \noindent \textbf{STEP 1.} We first set-up some quantities which will be used in the rest of the proof. \smallskip We recall $(H_{3})$: $R_.,\,|\phi _{.}|^{2},\,n_{.},\,\lambda_{.} \in L(\mu ,h)$, where $f\in L(\mu ,h)$ if and only if $f(t)\leq \mu f(s)$ for $\left\vert t-s\right\vert \leq h$. We set, for $q_1,K_1>1$ to be fixed in the sequel, \[ f_R(t)= K_1 \left(\frac{\mu n_t}{\lambda_t}\right)^{q_1} \left( \frac{1}{h}+\frac{1}{R_t}+|\phi_t|^2\right). \] Then straightforward computations give that $f_R\in L(\mu^{2 q_1+1},h)$. We define \be{tdelta} \delta(t)=\inf\left\{\delta>0\,:\, \int_t^{t+\delta} f_R(s) ds \geq \frac{1}{\mu^{2 q_1+1}} \right\}. \end{equation} We have \[ \frac{\delta(t)}{h} = \int_t^{t+\delta(t)} \frac{1}{h} ds \leq \int_t^{t+\delta(t)} f_R(s)ds =\frac{1}{\mu^{2 q_1+1}}, \] so $\delta(t)\leq h$. We now prove that $\delta(\cdot)\in L(\mu^{4q_1+1}, h)$. In fact, if $0<t-t'\leq h$, \[ \mu^{2q_1+1} f_R(t) \delta(t) \geq \int_{t}^{t+\delta(t)} f_R (s) ds=\frac 1{\mu^{2q_1+1}} =\int_{t'}^{t'+\delta(t')}f_{R}(s)ds \geq \mu^{-(2q_1+1)} f_{R}(t)\delta(t'), \] so $\delta(t')\leq \mu^{4q_1+2}\delta(t)$. Since the converse holds as well, we get $\delta(\cdot) \in L(\mu^{4q_1+2},h)$. We now prove a further property for $\delta(\cdot)$: we have \[ \frac{1}{\mu^{2 q_1+1}}= \int_t^{t+\delta(t)} f_R(s) ds \geq \int_t^{t+\delta(t)} \frac{f_R(t)}{\mu^{2 q_1+1}} ds \geq \delta(t) \frac{f_R(t)}{\mu^{2 q_1+1}}, \] so \be{Rdelta} \delta(t) \leq \frac{1}{f_R(t)} \leq \frac{R_t}{K_1} \left(\frac{\lambda_t}{\mu n_t}\right)^{q_1} \leq \frac{1}{K_1} \left(\frac{\lambda_t}{\mu n_t}\right)^{q_1} \in1/\mathcal{A} \end{equation} (recall that $R_t,\lambda_t\leq 1$ and $n_t\geq 1$ for every $t$). We also set the energy over the time interval $[t,t+\delta(t)]$: \[ \varepsilon_t(\delta(t))=\left( \int_t^{t+\delta(t)} |\phi_s|^2 ds\right)^{1/2}. \] Since $n,\lambda\in L(\mu, h)$ and $\delta(t)\leq h$, for $s\in (t,t+\delta(t))$ we have $$ f_R(s)\geq K_1\Big(\frac{\mu n_s}{\lambda_s}\Big)^{q_1}|\phi_s|^2 \geq \frac{K_1}{\mu^{2q_1}}\Big(\frac{\mu n_t}{\lambda_t}\Big)^{q_1}|\phi_s|^2. $$ Hence $$ \frac 1{\mu^{2q_1+1}}=\int_t^{t+\delta(t)}f_R(s)ds\geq \frac{K_1}{\mu^{2q_1}}\Big(\frac{\mu n_t}{\lambda_t}\Big)^{q_1}\int_t^{t+\delta(t)}|\phi_s|^2ds, $$ which gives that \be{energy} \varepsilon_t(\delta(t))^2 \leq \frac{1}{K_1} \left(\frac{\lambda_t}{\mu n_t}\right)^{q_1}\in1/\mathcal{A}. \end{equation} \textbf{STEP 2.} We set now some notation and properties that will be used in the ``concatenation'', which is developed in the following steps. \smallskip We define the time grid as \[ t_0=0,\quad t_k=t_{k-1}+\delta(t_{k-1}), \] and introduce the following notation on the grid: \[ \delta_k=\delta(t_k),\ \ \varepsilon_k=\varepsilon_{t_k}(\delta_k),\ \ n_k=n_{t_k},\ \ \lambda_k= \lambda_{t_k},\ \ X_k=X_{t_k},\ \ x_k=x_{t_k}(\phi),\ \ R_k=R_{t_k}. \] Recall that $\delta(t)<h$ for every $t$, so we have \[ R_k/\mu\leq R_t\leq \mu R_k,\quad \mbox{ for } t_k\leq t \leq t_{k+1}. \] We also define \[ \hat{X}_k= X_k + b(t_k,X_k)\delta_k, \quad \hat{x}_k=x_k+b(t_k, x_k)\delta_k, \] and for $t_k\leq t \leq t_{k+1}$, \[ \hat{X}_k(t)=X_k+b(t_k,X_k)(t-t_k), \quad \hat{x}_k(t)=x_k+b(t_k, x_k)(t-t_k). \] Let $r^*\in1/\mathcal{A}$ be the radius-function of Theorem \ref{m}, in the version of Remark \ref{est-tx}, associated to the points $(t,x_t(\phi))$ as $t\in[0,T]$. We set $r^*_k=r^*_{t_k}$. Let us see some properties. For all $t_k\leq t\leq t_{k+1}$, we have $R_t\geq R_k/\mu\geq \delta_k/\mu$ and, by using \eqref{Norm2}, we obtain $$ |\xi|_{A_{R_t}(t,x_t)} \leq \sqrt{\frac{\delta_k}{R_k}}\, |\xi|_{A_{\delta_k/\mu}(t,x_t)} \leq |\xi|_{A_{\delta_k/\mu}(t,x_t)}, $$ last inequality holding because $\delta_k\leq R_k$. Since $\delta_k/\mu\leq \delta_k$, we apply again \eqref{Norm2} to the norm in the right hand side above and we get \begin{equation}\label{magia0} |\xi|_{A_{R_t}(t,x_t)} \leq \mu\, |\xi|_{A_{\delta_k}(t,x_t)}. \end{equation} Taking $\xi=x_t-\hat{x}_k(t)$, we have $$ |x_t-\hat{x}_k(t)|_{A_{R_t}(t,x_t)} \leq \mu\, |x_t-\hat{x}_k(t)|_{A_{\delta_k}(t,x_t)}. $$ By \eqref{Rdelta} and \eqref{energy}, we can choose $q_1, K_1$ large enough such that $\delta(t)\leq \delta^*(t)$, $\varepsilon_t(\delta(t))\leq \varepsilon^*(t)$ where $\delta^*\in 1/\mathcal{A}$ and $\varepsilon^*\in 1/\mathcal{A}$ are the functions in Lemma \ref{normcontrol}. So, we apply \eqref{correction3} to the norm in the above right hand side and we obtain $$ |x_t-\hat{x}_k(t)|_{A_{R_t}(t,x_t)} \leq \mu \times 4|x_t-\hat{x}_k(t)|_{A_{\delta_k}(t_k,x_k)}. $$ We use now \eqref{normcontrolhat}: for some $\bar C\in \mathcal{A}$, we get $$ |x_t-\hat{x}_k(t)|_{A_{\delta_k}(t_k,x_k)} \leq \bar C_k(\varepsilon_k\vee\sqrt{\delta_k}) $$ where $\bar C_k=\bar C_{t_k}$, and, as a consequence of the estimate above, we have also $$ |x_t-\hat{x}_k(t)|_{A_{R_t}(t,x_t)} \leq 4\mu \,\bar C_k(\varepsilon_k\vee\sqrt{\delta_k}), $$ for all $t\in[t_k,t_{k+1}]$ and for all $k$. By recalling that $x_{t_{k+1}}-\hat{x}_k(t_{k+1})=x_{k+1}-\hat{x}_k$, and possibly choosing $K_1$ larger, we can resume by asserting that $\delta_k\leq \delta^*_{t_k}$ in Theorem \ref{m} with initial condition $(t_k,x_k)$ (see its version in Remark \ref{est-tx}) and \begin{align} |x_{k+1}-\hat{x}_k|_{A_{\delta_k}(t_k,x_k)} &\leq r_k^*/4 \mbox{ for all } k,\label{magia1}\\ |\hat{x}_k(t)-x_t|_{A_{R_t}(t,x_t)} &\leq \frac{1}{4} \mbox{ for all } t\in[t_k, t_{k+1}] \mbox{ and for all } k.\label{magia2} \end{align} We have already noticed that, under our settings, \eqref{correction3} holds, so that \[ \frac{1}{2} |\xi|_{A_{\delta_k}(t_k,x_k)}\leq | \xi|_{A_{\delta_k}(t_{k+1},x_{k+1})}\leq 2 |\xi|_{A_{\delta_k}(t_k,x_k)}. \] Since $\delta(\cdot) \in L(\mu^{4q_1+2},h)$, one has $\delta_k/\delta_{k+1}\leq \mu^{4q_1+2}$ and $\delta_{k+1}/\delta_k\leq \mu^{4q_1+2}$. So, using \eqref{Norm2} to the right hand side of the above inequality we easily get \be{fe} \frac{1}{2\mu^{2q_1+1}}|\xi|_{A_{\delta_k}(t_k,x_k)} \leq |\xi|_{A_{\delta_{k+1}}(t_{k+1},x_{k+1})}\leq 2\mu^{2q_1+1} |\xi|_{A_{\delta_k}(t_k,x_k)} \mbox{ for all }k. \end{equation} \textbf{STEP 3.} We are ready to set-up the concatenation for the lower bound. \smallskip We set, for $K_2$ and $q_2$ to be fixed in the sequel, \be{ray} r_k= \frac{1}{K_2 \mu^{2q_1+2q_2+1} }\left(\frac{\lambda_k}{n_k}\right)^{q_2}. \end{equation} Moreover, since $\lambda,n\in L(\mu,h)$ and $\delta_k\leq h$, one easily gets $r_{k+1}/r_k\leq \mu^{2 q_2}$ for every $k$. We define \[ \Gamma_k=\{y\,:\, |y-x_k|_{A_{\delta_k}(t_k,x_k)} \leq r_k \}\mbox{ and } {\mathbb P}_k(\cdot)={\mathbb P}\left( \cdot|W_t,t\leq t_k; X_k\in\Gamma_k \right), \] that is, ${\mathbb P}_k$ is the conditional probability with respect to the knowledge of the Brownian motion up to time $t_k$ and the fact that $X_k\in\Gamma_k$. The aim of this step is to prove that \begin{equation}\label{to-prove1} {\mathbb P}_k(X_{k+1}\in\Gamma_{k+1})\geq 2 \mu^{-4n q_1} \exp(-K_3 (\log \mu + \log n_k -\log \lambda_k ))\mbox{ for all }k. \end{equation} for some constant $K_3$ depending on $K_1$, $K_2$, $q_1$ and $q_2$. We denote $\rho_k(X_k,y)$ the density of $X_{k+1}$ with respect to this probability. We prove that \begin{equation}\label{to-prove2} \Gamma_{k+1} \subset\{y:\,|y-\hat X_k|_{A_{\delta_k}(t_k,X_k)}\leq r^*_k\}. \end{equation} If \eqref{to-prove2} holds, as we will see, then we can apply the lower bound in Remark \ref{est-tx} to $\rho_k(X_k,y)$. More precisely, we use here the version of the estimate given in \eqref{denco}: there exists $\underline{C}\in\mathcal{A}$ such that \be{eqsttt} \rho_k(X_k,y)\geq \frac{1}{\underline{C}_k \sqrt{\det A_{\delta_k} A_{\delta_k}^T(t_k,X_k)}}\mbox{ for all }y\in \Gamma_{k+1}, \end{equation} where $\underline{C}_k=\underline{C}_{t_k}$. Let us show that \eqref{to-prove2} holds. We estimate $$ |y-\hat{X}_{k}|_{A_{\delta_{k}}(t_{k},x_{k})} \leq |y-x_{k+1}|_{A_{\delta_{k}}(t_{k},x_{k})}+|x_{k+1}-\hat{x}_k|_{A_{\delta_{k}}(t_{k},x_{k})} +|\hat{x}_k-\hat{X}_k|_{A_{\delta_{k}}(t_{k},x_{k})} $$ and by using \eqref{magia1} we obtain \be{deco} |y-\hat{X}_{k}|_{A_{\delta_{k}}(t_{k},x_{k})} \leq |y-x_{k+1}|_{A_{\delta_{k}}(t_{k},x_{k})}+\frac{r^*_k}4 +|\hat{x}_k-\hat{X}_k|_{A_{\delta_{k}}(t_{k},x_{k})}. \end{equation} Using \eqref{fe}, the fact that $r_{k+1}/r_k\leq \mu^{2 q_2}$ and recalling that $|y -x_{k+1}|_{A_{\delta_{k+1}}(t_{k+1},x_{k+1})}\leq r_{k+1}$, we obtain \begin{align*} |y-x_{k+1}|_{A_{\delta_{k}}(t_{k},x_{k})} &\leq 2 \mu^{2 q_1+1} |y-x_{k+1}|_{A_{\delta_{k+1}}(t_{k+1},x_{k+1})} \leq 2 \mu^{2 q_1+1} r_{k+1}\\ &\leq 2 \mu^{2 q_1+2 q_2 + 1} r_{k} \leq \frac{2}{K_2}\left(\frac{\lambda_k}{n_k}\right)^{q_2}. \end{align*} \eqref{Not2} also gives $|\hat{x}_k-\hat{X}_k|_{A_{\delta_{k}}(t_{k},x_{k})} \leq C_k |{x}_k-{X}_k|_{A_{\delta_{k}}(t_{k},x_{k})}$, where $C_k=C_{t_k}$ and $C$ is a suitable function in $\mathcal{A}$, and the conditioning with respect to $\Gamma_k$ gives $|\hat{x}_k-\hat{X}_k|_{A_{\delta_{k}}(t_{k},x_{k})} \leq C_k r_k$. Similarly, $|\hat{x}_k(t)-\hat{X}_k(t)|_{A_{R_t}(t,x_t)} \leq C_k|x_k-X_k|_{A_{R_t}(t,x_t)}$ and by using firstly \eqref{magia0} and secondly \eqref{correction3}, we get $$ |\hat{x}_k(t)-\hat{X}_k(t)|_{A_{R_t}(t,x_t)} \leq C_k\times \mu |x_k-X_k|_{A_{\delta_k}(t,x_t)} \leq C_k\mu\times 2 |x_k-X_k|_{A_{\delta_k}(t_k,x_k)} \leq 2\mu C_k r_k, $$ for every $t\in[t_k,t_{k+1}]$. Recalling \eqref{ray}, $K_2$ and $q_2$ (possibly large) such that $|y-x_{k+1}|_{A_{\delta_{k}}(t_{k},x_{k})}\leq r^*_k/8$, $|\hat{x}_k-\hat{X}_k|_{A_{\delta_{k}}(t_{k},x_{k})} \leq r^*_k/8$, and \be{magia3} |\hat{X}_k(t)-\hat{x}_k(t)|_{A_{R_t}(t,x_t)} \leq 1/4,\quad \mbox{ for all } t\in[t_k, t_{k+1}]\mbox{ and for all }k. \end{equation} From \eqref{deco}, this implies $|y-\hat{X}_{k}|_{A_{\delta_{k}}(t_{k},x_{k})} \leq r^*_k/2$. On the event $\Gamma_k$, we also have, from \eqref{Norm3}, $|x_k-X_k|\leq |x_k-X_k|_{A_{\delta_{k}}(t_{k},x_{k})} \lambda^*(A(t_k,x_k))\sqrt{\delta_k}\leq n_{t_k}^\alpha \sqrt{\delta_k}\, r_k$, for some universal constant $\alpha>0$. So, we can fix $K_2$ and $q_2$ in order that Lemma \ref{NORM3} holds with $R=\delta_k$, $x=x_k$, $y=X_k$, $t=t_k$ and $s=0$. Then, we get \[ \frac{1}{2}|\xi|_{A_{\delta_{k}}(t_{k},x_{k})}\leq|\xi|_{A_{\delta_k}(t_k,X_k)}\leq 2|\xi|_{A_{\delta_{k}}(t_{k},x_{k})}. \] These inequalities give two consequences. First, we have $$ |y-\hat{X}_{k}|_{A_{\delta_k}(t_k,X_k)} \leq 2 |y-\hat{X}_{k}|_{A_{\delta_k}(t_k,x_k)}\leq r^*_k, $$ so that \eqref{to-prove2} actually holds and then \eqref{eqsttt} holds as well. As a second consequence, we have that \begin{align*} &\Big\{y:|y -x_{k+1}|_{A_{\delta_k}(t_k,X_k)}\leq \frac{r_{k+1}}{4\mu^{2 q_1+1}} \Big\} \subset \Big\{y:|y -x_{k+1}|_{A_{\delta_{k}}(t_{k},x_{k})}\leq \frac{r_{k+1}}{2\mu^{2 q_1+1} }\Big\}\\ &\quad\subset \{y:|y -x_{k+1}|_{A_{\delta_{k+1}}(t_{k+1},x_{k+1})}\leq r_{k+1}\} =\Gamma_{k+1} , \end{align*} in which we have used \eqref{fe}. Since $ r_{k+1}/(4\mu^{2 q_1+1} )\geq r_k /(4\mu^{2 q_1+ 2q_2+1} )$, we obtain \begin{align*} \Gamma_{k+1} \supset \Big\{y:|y -x_{k+1}|_{A_{\delta_k}(t_k,X_k)}\leq \frac{r_k}{ 4\mu^{2 q_1+ 2q_2+1}} \Big\}. \end{align*} By recalling that $r_k /(4\mu^{2 q_1+ 2q_2+1} )= \frac{1}{4 K_2 \mu^{4 q_1+ 4 q_2+2} }\left(\frac{\lambda_k}{n_k}\right)^{q_2}$, we can write, with $\mathrm{Leb}\,_n$ denoting the Lebesgue measure in ${\mathbb R}^n$, \[ \mathrm{Leb}\,_n(\Gamma_{k+1}) \geq \sqrt{\det(A_{\delta_k} A_{\delta_k}^T(t_k,X_k))} \left(\frac{1}{4 K_2 \mu^{4 q_1+ 4 q_2+2} }\left(\frac{\lambda_k}{n_k}\right)^{q_2}\right)^n. \] So, from \eqref{eqsttt}, \[ {\mathbb P}_k(X_{k+1}\in \Gamma_{k+1})\geq \frac{1}{\underline{C}_k} \left(\frac{1}{4 K_2 \mu^{4 q_1+ 4 q_2+2} }\left(\frac{\lambda_k}{n_k}\right)^{q_2}\right)^n \] where $\underline{C}_k$ is the constant in \eqref{eqsttt}. This implies \eqref{to-prove1}, for some constant $K_3$ depending on $K_2$ and $q_2$. \smallskip \textbf{STEP 4.} We give here the proof of the lower bounds \eqref{tube-lower} and \eqref{tube-lower-bis}. We set \[ D_k=\Big\{ \sup_{t_k\leq t \leq t_{k+1}} |X_t-x_t|_{A_{R_t}(t,x_t)}\leq 1 \Big\} \mbox{ and } E_k=\Big\{\sup_{t_k\leq t\leq t_{k+1}} |X_t-\hat{X}_k(t)|_{A_{R_t}(t,x_t)}\leq \frac{1}{2}\Big\}. \] For $t\in[ t_k, t_{k+1}]$, by using \eqref{magia2} and \eqref{magia3} we have \begin{align*} |X_t-x_t|_{A_{R_t}(t,x_t)} &\leq|X_t-\hat{X}_k(t)|_{A_{R_t}(t,x_t)}+|\hat{X}_k(t)-\hat{x}_k(t)|_{A_{R_t}(t,x_t)} +|\hat{x}_k(t)-x_t|_{A_{R_t}(t,x_t)}\\ &\leq |X_t-\hat{X}_k(t)|_{A_{R_t}(t,x_t)}+\frac 12, \end{align*} so that $E_k\subset D_k$. Moreover, by passing from Stratonovich to It\^o integrals and by using \eqref{Norm3}, we have \[ \begin{split} &|X_t-\hat{X}_k(t)|_{A_{R_t}(t,x_t)} \leq |\sigma(t_k,X_{t_k})(W_t-W_{t_k})|_{A_{R_t}(t,x_t)}\\ &\quad + \Big|\int_{t_k}^{t}\big(\sigma(s,X_s)-\sigma(t_k,X_k)\big) dW_s\Big|_{A_{R_t}(t,x_t)}+ \Big| \int_{t_k}^{t} \big(b(s,X_s)-b(t_k,X_k)\big) ds\Big|_{A_{R_t}(t,x_t)}\\ &\quad+ \sum_{l=1}^d \Big|\int_{t_k}^{t} \nabla \sigma_l(s,X_s)(\sigma_l(s,X_s)-\sigma_l(t_k,X_k)) ds\Big|_{A_{R_t}(t,x_t)}\\ &\quad\leq \Big| \frac{\sqrt{\mu}}{\sqrt{R_k}}\sigma(t_k,X_{t_k})(W_t-W_{t_k})\Big|_{A(t,x_t)}+ \Big|\frac{\mu}{R_k}\int_{t_k}^{t}\big(\sigma(s,X_s)-\sigma(t_k,X_k)\big) dW_s\Big|\\ &\quad+ \Big|\frac{\mu}{R_k} \int_{t_k}^{t} \big(b(s,X_s)-b(t_k,X_k)\big) ds\Big|+ \sum_{l=1}^d \Big|\frac{\mu}{R_k}\int_{t_k}^{t} \frac{\nabla \sigma_l(s,X_s)}{2}(\sigma_l(s,X_s)-\sigma_l(t_k,X_k)) ds\Big|. \end{split} \] We use now the exponential martingale inequality (see also Remark \ref{2}) and we find that \[ {\mathbb P}_k(E_k^c) \leq \exp\left(- \frac{1}{K_4} \left(\frac{\lambda_k}{\mu n_k}\right)^{q_4} \frac{R_k}{\delta_k} \right) \] for some constants $K_4,q_4$. From \eqref{Rdelta}, $R_k/\delta_k\geq K_1(\mu n_k/\lambda_k)^{q_1}$, so by choosing $K_1$ and $q_1$ possibly larger and by recalling \eqref{to-prove1}, we can conclude that \[ {\mathbb P}_k(E^c_k) \leq \mu^{-4 nq_1} \exp(-K_3 (\log \mu + \log n_k -\log \lambda_k )) \leq \frac{1}{2}{\mathbb P}_k(X_{k+1}\in\Gamma_{k+1}). \] Hence, \be{shorttimediff} \begin{split} &{\mathbb P}_k(\{X_{k+1}\in\Gamma_{k+1}\}\cap D_k) \geq {\mathbb P}_k(\{X_{k+1}\in\Gamma_{k+1}\}\cap E_k)\geq {\mathbb P}_k(X_{k+1}\in\Gamma_{k+1})-{\mathbb P}_k(E^c_k)\\ &\quad \geq \frac{1}{2} {\mathbb P}_k(X_{k+1}\in\Gamma_{k+1})\geq \exp\left( -K_5 (\log \mu +\log n_k -\log \lambda_k) \right), \end{split} \end{equation} for some constant $K_5$. Let now $N(T)=\max\{k: t_k\leq T\}$. From definition \eqref{tdelta}, \[ \int_0^T f_R(t)dt \geq \sum_{k=1}^{N(T)} \int_{t_{k-1}}^{t_k} f_R(t)dt = \frac{N(T)}{\mu^{2q_1+1}}. \] From \eqref{shorttimediff}, \begin{align*} &{\mathbb P}\Big(\sup_{t\leq T} |X_t-x_t|_{A_{R_t}(t,x_t)}\leq 1 \Big) \geq {\mathbb P}\Big(\bigcap_{k=1}^{N(T)} \{X_{k+1}\in\Gamma_{k+1}\}\cap D_k\Big) \\ &\quad \geq \prod_{k=1}^{N(T)} \exp(-K_5(\log \mu +\log n_k -\log \lambda_k)) = \exp\Big(-K_5\sum_{k=1}^{N(T)}\big(\log \mu +\log n_k -\log \lambda_k\big)\Big). \end{align*} Since \[ \begin{split} \sum_{k=1}^{N(T)}(\log \mu +\log n_k -\log \lambda_k) &=\mu^{2q_1+1}\sum_{k=1}^{N(T)}\int_{t_k}^{t_k+1} f_R(t)(\log \mu +\log n_k -\log \lambda_k)dt \\ &\leq \mu^{2q_1+1} \int_0^T f_R(t) \log\Big(\frac{\mu^{3} n_t}{\lambda_t}\Big)dt, \end{split} \] the lower bound \eqref{tube-lower} follows. Concerning \eqref{tube-lower-bis}, it is immediate from \eqref{tube-lower} and the fact that $R_t \leq R_t ^*(\phi)\leq h\, \exp\left(-K_* \left(\frac{\mu n_t}{\lambda_t}\right)^{q_*}\right)$. \end{proof} We can now address the problem of the upper bound. \begin{theorem}\label{th-upper} Let $\mu\geq 1$, $h\in (0,1]$, $n\,:\,[0,T]\to [1,+\infty)$, $\lambda\,:\,[0,T]\to (0,1]$, $\phi\in L^2([0,T],{\mathbb R}^n)$ and $R\,:\,[0,T]\to (0,1]$ be such that $(H_1)$--$(H_3)$ in \eqref{Not6} hold. Suppose that, for some $K_*,q_*>0$ and for $R^*(\phi)$ as in \eqref{Rmax}, one has $R_t \leq R_t ^*(\phi)$. Then there exist $\bar K,\bar q>0$ such that \begin{equation}\label{tube-upper} \begin{split} &{\mathbb P}\Big(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \Big)\\ &\qquad\leq \exp\Big(- \int_0^T \bar{K} \Big(\frac{\mu n_t}{\lambda_t}\Big)^{\bar{q}} \Big[\frac{\exp\Big(-K_* \Big(\frac{\mu n_t}{\lambda_t}\Big)^{q_*}\Big)}{R_t}+|\phi_t|^2\Big]dt \Big). \end{split} \end{equation} \end{theorem} \begin{proof} We refer here to notation and arguments already introduced and developed in the proof of Theorem \ref{th-lower}. So, when we recall here STEP 1, 2 and 3, we intend to refer to the same steps developed in the proof of Theorem \ref{th-lower}. We define, with the same $K_1,q_1$ as in STEP 1, \[ g_R(t)= K_1 \left(\frac{\mu n_t}{\lambda_t}\right)^{q_1} \left( \frac{\exp \left( -K_*\left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \mu^{2q_*} \right)}{R_t} +|\phi_t|^2 \right) \] Because of \eqref{Rmax}, for all $t\in[0,T$], \be{rh} \frac{\exp \left( -K_*\left(\frac{\mu n_t}{\lambda_t}\right)^{q_*}\mu^{2q_*} \right)}{R_t} \geq \frac{1}{h} \end{equation} We work here with $\delta(t)$ as in the proof od Theorem \ref{th-lower} but defined from $g_R$: \[ \delta(t)= \inf\Big\{\delta>0\,:\, \int_t^{t+\delta} g_R(s) ds \geq \frac{1}{\mu^{2q_1+1}} \Big\}. \] We set, as before, \[ \varepsilon_t(\delta(t))=\Big( \int_t^{t+\delta(t)} |\phi_s|^2 ds\Big)^{1/2}. \] As in STEP 1, using also \eqref{rh}, we can check estimates similar to \eqref{Rdelta} and \eqref{energy}: we have indeed, \[ \delta(t) \leq \frac{h}{K_1} \Big(\frac{\lambda_t}{\mu n_t}\Big)^{q_1} \leq \frac{1}{K_1} \Big(\frac{\lambda_t}{\mu n_t}\Big)^{q_1}\quad\mbox{ and }\quad \varepsilon_t(\delta(t))^2 \leq \frac{1}{K_1} \Big(\frac{\lambda_t}{\mu n_t}\Big)^{q_1}. \] In particular, $\delta(t)\leq h$. With these definitions we set a time grid $\{t_k: k=0,\dots, N(T)\}$ and all the associated quantities as in STEP 2. As we did for the lower bound, since we estimate the probability of remaining in the tube for any $t\in [t_k,t_{k+1}]$, we can suppose that the bound in \eqref{hpder} holds on $ {\mathbb R}^+\times {\mathbb R}^n$ (recall Remark \ref{2}). The short time density estimate \eqref{denco} holds again. Recall now that $R_{.}\in L(\mu,h)$, and this gives the analogous to \eqref{fe}: \be{fewR} \frac{1}{2 \sqrt{\mu}}|\xi|_{A_{R_k}(t_k,x_k)}\leq |\xi|_{A_{R_{k+1}}(t_{k+1},x_{k+1})} \leq 2 \sqrt{\mu} |\xi|_{A_{R_k}(t_k,x_k)} \end{equation} We define \[ \Delta_k=\{y\,:\, |y-x_k|_{A_{R_k}(t_k,x_k)} \leq 1 \} \mbox{ and } \tilde{{\mathbb P}}_k(\cdot)={\mathbb P}\left( \cdot|W_t,t\leq t_k; X_k\in \Delta_k \right), \] so $\tilde{{\mathbb P}}_k$ is the conditional probability given the Brownian path up to time $t_k$ and the fact that $X_k\in \Delta_k$. Now, since $\delta(t)\leq h$ and $R,\lambda,n\in L(\mu,h)$, we have $$ \int_t^{t+\delta(t)} K_1 \left(\frac{\mu n_s}{\lambda_s}\right)^{q_1} |\phi|_s^2 ds \leq \mu^{2 q_1} K_1 \left(\frac{\mu n_t}{\lambda_t}\right)^{q_1} \int_t^{t+\delta(t)} |\phi|_s^2 ds $$ and \[ \begin{split} &\int_t^{t+\delta(t)} K_1 \left(\frac{\mu n_s}{\lambda_s}\right)^{q_1} \frac{\exp \left( -K_*\left(\frac{\mu n_s}{\lambda_s}\right)^{q_*} \mu^{2q_*}\right) }{R_s} ds\\ &\qquad\qquad\qquad\leq \mu^{2q_1+1} K_1 \left(\frac{\mu n_t}{\lambda_t}\right)^{q_1}\exp \left( -K_*\left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \right) \frac{\delta(t)}{R_t}. \end{split} \] Since \[ R_t\leq R_t^*(\phi)\leq\exp\left(-K_* \left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \mu^{2q_*}\right) \left( \inf_{0\leq \delta\leq h} \left\{ \delta \big/ \int_t^{t+\delta} |\phi_s|^2 ds \right\} \right), \] we have \[ \int_t^{t+\delta(t)} |\phi_s|^2 ds\leq \exp\left(-K_* \left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \right)\frac{\delta(t)}{R_t} \] We obtain \[ 1= \mu^{2q_1+1}\int_t^{t+\delta(t)} g_R(s) ds \leq 2\mu^{4q_1+2} K_1 \left(\frac{\mu n_t}{\lambda_t}\right)^{q_1} \exp \left( -K_*\left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \right) \frac{\delta(t)}{R_t} \] so \be{Rdelta2} \frac{R_t}{\delta(t)} \leq 2\mu^{4q_1+2} K_1 \left(\frac{\mu n_t}{\lambda_t}\right)^{q_1} \exp \left( -K_*\left(\frac{\mu n_t}{\lambda_t}\right)^{q_*} \right) \end{equation} As we did in STEP 3, if $q_*,K_*$ are large enough, $R_{k}$ is small enough and the upper bound for the density holds on $\Delta_{k+1}$. By using \eqref{fewR} and \eqref{Norm2}, we obtain \[ \begin{split} \mathrm{Leb}\,_n( y:|y-x_{k+1}|_{A_{R_{k+1}}(t_{k+1},x_{k+1})} \leq 1) &\leq 2^n \mathrm{Leb}\,_n(y: |y-x_{k+1}|_{A_{R_k}(t_k,x_k)}\leq 1) \\ &= 2^n \sqrt{\det(A_{R_k} A_{R_k}^T(t_k,x_k))} \\ &= C_k \sqrt{\det(A A^T(t_k,x_k))}\, R_k^{n-\frac{\mathrm{dim}\,\langle \sigma(t_k,x_k)\rangle}2}, \end{split} \] in which we have used the Cauchy-Binet formula (see also Remark \ref{C-B}). Now, using the upper estimate for the density in the version of Theorem \ref{m} given in Remark \ref{est-tx}, we obtain \[ \tilde{{\mathbb P}}_k(X_{k+1}\in \Delta_{k+1})\leq e^{\overline{C}_k} \left(\frac{{R_k}}{\delta_k}\right)^{n-\frac{\mathrm{dim}\,\langle \sigma(t_k,x_k)\rangle}2} \] where $\overline{C}_k=\overline{C}_{t_k}$, $\overline{C}\in \mathcal{A}$ (see the constant in the upper bound of \eqref{denco}). Recall \eqref{Rdelta2}, for $t=t_k$ \[ \frac{R_k}{\delta_k} \leq 2\mu^{4q_1+2} K_1 \left(\frac{\mu n_k}{\lambda_k}\right)^{q_1} \exp \left( -K_*\left(\frac{\mu n_k}{\lambda_k}\right)^{q_*} \right) \] so we chose now $K_*,q_*$ large enough to have \[ \tilde{{\mathbb P}}_k(X_{k+1}\in\Delta_{k+1})\leq \exp(-K_2) \] for a constant $K_2>0$. From the definition of $N(T)$ \[ \int_0^T g_R(t)dt = \sum_{k=1}^{N(T)} \int_{t_{k-1}}^{t_k} g_R(t)dt = \frac{N(T)}{\mu^{2q_1+1}}\leq N(T). \] So, we have \begin{align*} & {\mathbb P}\Big(\sup_{t\leq T} |X_t-x_t(\phi)|_{A_{R_t}(t,x_t(\phi))}\leq 1 \Big)\leq E_N^{\eta_N}\Big(\prod_{k=1}^{N(T)} \tilde{{\mathbb P}}_k(\Delta_{k+1})\Big)\\ & \quad \leq \prod_{k=1}^{N(T)} \exp(-K_2)= \exp(-K_2\,N(T))\leq \exp\Big(-K_2\int_0^T g_R(t)\Big) \end{align*} and \eqref{tube-upper} holds. \end{proof} \section{On the equivalence with the control distance}\label{sectioncontrol} We establish here the local equivalence between the norm $% \left\vert \cdot\right\vert _{A_{R}(t,x)}$ and the control (Carath\'e\-odory) distance. We use in a crucial way the alternative characterization given in \cite{NagelSteinWainger:85}. These results hold in the homogeneous case, so we consider now the vector fields $\sigma _{j}(t,x)=\sigma _{j}(x)$, and the associated norm $A_{R}(t,x)=A_{R}(x)$. We assume in this section the following bound on $\sigma$: there exists $\kappa\,:\,{\mathbb R}^n\to [1,+\infty)$ such that \begin{equation}\label{bound} \sup_{|y-x|\leq 1}\sum_{0\leq |\alpha| \leq 4} \sum_{j=1}^d |\partial_x^{\alpha} \sigma_j(y)| \leq \kappa(x), \quad\quad \forall x \in {\mathbb R}^n. \end{equation} So, \eqref{bound} agrees with \eqref{Not2} in the homogeneous case and when $b=0$. We now introduce a quasi-distance $d$ which is naturally associated to the family of norms $\left\vert y\right\vert _{A_{R}(x)}$. We set \[ O=\{x\in {\mathbb R}^n: \lambda_*(A(x))>0\}=\{x:\det(AA^T(x))\neq 0\} \] which is an open set since $x\mapsto \det(AA^T(x))$ is a continuous function. Notice that if $x\in O$ then $\det(A_R A_R^T(x))>0$ for every $R>0$. For $x,y\in O$, we define $d(x,y)$ by \[ d(x,y)<\sqrt{R}\quad \Leftrightarrow \quad \left\vert y-x\right\vert _{A_{R}(x)}<1. \] The motivation for taking $\sqrt{R}$ is the following: in the elliptic case $\left\vert y-x\right\vert _{A_{R}(x)}\sim R^{-1/2}\left\vert y-x\right\vert $ so $\left\vert y-x\right\vert _{A_{R}(x)}\leq 1$ amounts to $\left\vert y-x\right\vert \leq \sqrt{R}$. It is straightforward to see that $d$ is a quasi-distance on $O$, meaning that $d$ verifies the following three properties (see \cite{NagelSteinWainger:85}): \begin{itemize} \item[$i)$] for every $x\in O$ and $r>0$, the set $\{y\in O\,:\,d(x,y)<r\}$ is open; \item[$ii)$] $d(x,y)=0$ if and only if $x=y$; \item[$iii)$] for every compact set $K\Subset O$ there exists $C>0$ such that for every $x,y,z\in K$ one has $d(x,y)\leq C\big(d(x,z)+d(z,y)\big)$. \end{itemize} We recall the definition of equivalence of quasi-distances. Two quasi-distances $d_{1}:\Omega \times\Omega \rightarrow {\mathbb R}^{+}$ and $% d_{2}:\Omega\times \Omega\rightarrow {\mathbb R}^{+}$ are equivalent if for every compact set $K\Subset\Omega$ there exists a constant $C$ such that for every $x,y\in K$ \begin{equation} \frac{1}{C}d_{1}(x,y)\leq d_{2}(x,y)\leq Cd_{1}(x,y). \label{Norm7as} \end{equation} $d_1$ and $d_2$ are locally equivalent if for every $\xi\in \Omega$ there exists a neighborhood $V$ of $\xi$ such that $d_1$ and $d_2$ are equivalent on $V$. We introduce now the control metric. Without loss of generality, we assume $T=1$, For $\psi\in L^2([0,1],{\mathbb R}^d)$, let $u(\psi)$ satisfy the following controlled equation: \be{controlequation} du_{t}(\psi )=\sum_{j=1}^{d}\sigma _{j}(u_{t}(\psi))\psi _{t}^{j}dt. \end{equation} Notice that the equation for $u(\psi )$ is actually the skeleton equation \eqref{skeleton} when the drift $b$ is null. For $x,y\in O$ we denote by $C^2_{\sigma,1}(x,y)$ the set of controls $\psi \in L^{2}([0,1];{\mathbb R}^d)$ such that the corresponding solution $u(\psi)$ of \eqref{controlequation} satisfies $u_0(\psi)=x$ and $u_{1}(\psi )=y$. We define the control (Carath\'eodory) distance as \begin{equation*} d_{c}(x,y)=\inf_{\psi\in C^2_{\sigma,1}(x,y)}\|\psi\|_2. \end{equation*} For $\delta\in(0,1]$, we also denote $C^2_{\sigma,\delta }(x,y)$ the set of controls $\phi \in L^{2}([0,\delta ];{\mathbb R}^d)$ such that the corresponding solution $u(\phi)$ to \eqref{controlequation} satisfies $u_{0}(\phi )=x$ and $u_{\delta }(\phi )=y$. For $\phi\in C_{\sigma,\delta}^2(x,y)$, we set the associated energy \[ \varepsilon_{\phi }(\delta )=\Big(\int_{0}^{\delta }\left\vert \phi _{s}\right\vert ^{2}ds\Big)^{1/2}. \] Notice that \begin{equation}\label{dc-energy} d_{c}(x,y)= \sqrt{\delta } \inf_{\phi\in C^2_{\sigma,\delta}(x,y)}\varepsilon _{\phi }(\delta ). \end{equation} Indeed, for each $x,y\in{\mathbb R}^n$ and $\psi \in C^2_{\sigma,1}(x,y)$, take $\phi_t=\delta ^{-1}\psi (t\delta ^{-1})$ and $\xi_t=u_{t/\delta }(\psi )$. Then, $d\xi_{t}=\sum_{j=1}^{d}\sigma _{j}(\xi_{t})\phi _{t}^{j}dt$ and of course $\xi_0=x$, $\xi_\delta=y$. Moreover, $\|\psi\|_2=\sqrt\delta\,\varepsilon_\phi(\delta)$. Lastly, we define $C^\infty_{\sigma,1}(x,y)$ the set of paths $g \in L^\infty([0,1];{\mathbb R}^d)$ such that the corresponding solution $u(g)$ of \eqref{controlequation} satisfies $u_0(g)=x$ and $u_{1}(g)=y$. Using this set of controls, we define \begin{equation*} d_\infty (x,y)=\inf_{g\in C^\infty_{\sigma,1}(x,y)}\|g\|_\infty. \end{equation*} Under \eqref{bound}, we define \[ {\mathcal{D}}=\Big\{C:O\to {\mathbb R}_+\,:\,C=K \Big(\frac{\kappa(x)}{\lambda(x)}\Big)^{q},\ \exists\ K,q>0\Big\}. \] Notice that ${\mathcal{D}}$ is actually the set in \eqref{DD} in the homogeneous case. \begin{theorem} \label{NORM4} Suppose that \eqref{bound} hold. \smallskip \textbf{A}. There exists $\bar{C}\in {\mathcal{D}}$ such that if $d_c(x,y)\leq 1/\bar{C}^2(x)$ then $d(x,y)\leq 2\bar{C}(x)d_c(x,y)$. \smallskip \textbf{B}. $d$ is locally equivalent to $d_{c}$ on $O$. \smallskip \textbf{C}. For every compact set $K\Subset O $ there exists $r_K$ and $C_K$ such that for every $x,y\leq r_K$ one has $d_c (x,y)\leq C_K d(x,y)$. \end{theorem} \begin{proof} \textbf{A.} Assume that $d_c(x,y)\leq 1/\bar{C}^2(x)$, with $\bar C\in\mathcal{D}$ to be chosen later. We set $\delta(x)=\bar C^2(x) d_c(x,y)^2$. Notice that $\delta(x)\leq 1/\bar C^2(x)$. \eqref{dc-energy} with $\delta=\delta(x)$ gives $$ d_{c}(x,y)= \sqrt{\delta(x) } \inf_{\phi\in C^2_{\sigma,\delta(x)}(x,y)}\varepsilon _{\phi }(\delta(x)) =\bar{C}(x) d_c(x,y) \inf_{\phi\in C^2_{\sigma,\delta(x)}(x,y)}\varepsilon _{\phi }(\delta(x)) $$ and thus, $$ \inf_{\phi\in C^2_{\sigma,\delta(x)}(x,y)}\varepsilon _{\phi }(\delta(x) )=\frac 1{\bar C(x)}<\frac 2{\bar C(x)}. $$ Hence, there exists $\phi_*\in C^2_{\sigma,\delta(x)}(x,y)$ such that $$ \varepsilon _{\phi_*}(\delta(x))<\frac 2{\bar C(x)}. $$ For every fixed $x$, we apply Remark \ref{for-dc} to $\phi_*$ (recall that here $b\equiv 0$): there exists $\bar\delta,\bar\varepsilon\in 1/\mathcal{D}$ and $\bar C\in\mathcal{D}$ such that (with the slightly different notation of the present section) $$ |u_{\delta}(\phi_*)-x|_{A_{\delta}(x)}\leq \overline{C}(x) (\varepsilon_{\phi_*}(\delta)\vee \sqrt{\delta}), $$ for every $\delta$ such that $\delta\leq \bar{\delta}(x)$ and $\varepsilon_{\phi_*}(\delta)\leq \bar{\varepsilon}(x)$. We have just proved that $\delta(x)\leq 1/\bar C^2(x)$ and $\varepsilon_{\phi_*}(\delta(x))\leq 2/\bar C(x)$. So, possibly taking $\bar C$ larger, we can actually use $\delta=\delta(x)$. And since $u_{\delta(x)}(\phi_*)=y$, the above inequality gives $$ |y-x|_{A_{\delta}(x)}\leq \overline{C}(x) (\varepsilon_{\phi_*}(\delta(x))\vee \sqrt{\delta(x)})\leq 2. $$ By \eqref{Norm3}, we obtain $|y-x|_{A_{4\delta}(x)}\leq 1$, that is $d(x,y)\leq \sqrt{4\delta(x)}=2\bar C(x) d_c(x,y)$, and the statement follows. \smallskip \textbf{B.} We prove now the converse inequality. We use a result from \cite{NagelSteinWainger:85}, for which we need to recall the definition of the quasi-distance $d_{\ast }$ (denoted by $\rho _{2}$ in \cite{NagelSteinWainger:85}). The definition we give here is slightly different but clearly equivalent. For $\theta \in {\mathbb R}^m$, consider the equation% \be{Norm15} dv_{t}(\theta )=A(v_t(\theta)) \theta dt. \end{equation} We denote \[ \bar{C}_A(x,y)=\{\theta\in {\mathbb R}^m\,:\, \mbox{ the solution $v(\theta)$ to \eqref{Norm15} satisfies $v_0(\theta)=x$ and $v_{1}(\theta )=y$}\}. \] Notice that $\theta\in\bar{C}_A(x,y)$ is a constant vector, and not a time depending control as in \eqref{controlequation}. Moreover, recalling the definitions \eqref{Al}-\eqref{Alucia} for $A$, \eqref{Norm15} involves also the vector fields $[\sigma_{i},\sigma_{j}]$, differently from \eqref{controlequation}. In both equations the drift term $b$ does not appear. Let $D_R$ be the diagonal matrix in \eqref{DR} and recall that $A_R(x)=A(x)D_R$. We define \[ d_{\ast }(x,y)=\inf\{R>0 \,:\, \mbox{ there exists $\theta\in \bar{C}_A(x,y)$ such that $|D_R^{-1}\theta|<1$}\}. \] As a consequence of Theorem 2 and Theorem 4 from \cite{NagelSteinWainger:85}, $d_{\ast }$ is locally equivalent with $d_\infty$. Since $d_c(x,y)\leq d_\infty(x,y)$ for every $x$ and $y$, one gets that $d_c$ is locally dominated from above by $d_{\ast}$. To conclude we need to prove that $d_{\ast }$ is locally dominated from above by $d$. Let us be more precise: for $x\in O$, we look for $C\in{\mathcal{D}}$ and $R\in 1/{\mathcal{D}}$ such that the following holds: if $0<R \leq R(x)$ and $d(x,y)\leq \sqrt{R }$, then there exists a control $\theta\in \bar{C}_A(x,y)$ such that $|D_R^{-1}\theta|<C(x)$. This implies $d_{\ast}(x,y)\leq C(x) \sqrt{R}$, and the statement holds. Notice that we discuss local equivalence, and that is why we can take $C(x)$ and $R(x)$ depending on $x$. Recall that $d(x,y)\leq \sqrt{R }$ means $|x-y|_{A_R(x)}\leq 1$, and this also implies $|x-y|\leq \lambda^*(A(x))\sqrt{R}$, by \eqref{Norm3}. Let $v(\theta)$ denote the solution to \eqref{Norm15} with $v_0(\theta)=x$. We look for $\theta$ such that $v_1(\theta)=y$. We define \[ \Phi(\theta)= \int_{0}^{1} A(v_s(\theta)) \theta ds = A(x) \theta + r(\theta) \] with $r(\theta)=\int_{0}^{1} (A(v_s(\theta))-A(x))\theta ds$. With this notation, we look for $\theta$ such that $\Phi(\theta)=y-x$. We introduce now the Moore-Penrose pseudoinverse of $A(x)$: $A(x)^+=A(x)^T(A A^T(x))^{-1}$. The idea here is to use it as in the least squares problem, but we need some computations to overcome the fact that we are in a non-linear setting. We use the following properties: $A A(x)^+ = \mathrm{Id}$; $|x-y|_{A(x)}=|A(x)^+\,(x-y)|$. Write $\theta=A(x)^+ \gamma$, $\gamma\in {\mathbb R}^n$. This implies $A(x)\theta =\gamma$, and so we are looking for $\gamma\in{\mathbb R}^n$ such that \[ \gamma + r( A(x)^+ \gamma)=y-x. \] One has $r(0)=0$, $\nabla r(0)=0$ and, as a consequence, $|r(\theta)|\leq C(x) |\theta|^2$, for some $C\in\mathcal{D}$ -- from now on, $C\in {\mathcal{D}}$ will denote a function that may vary from line to line. From the local inversion theorem (in a quantitative form), there exists $l\in {\mathcal{D}}$ such that $\gamma\mapsto\gamma + r( A(x)^+ \gamma)$ is a diffeomorrphism from $B(0,l_x)$ to $B(0,l_x/2)$. Remark that $|x-y|\leq \lambda^*(A(x))\sqrt{R}$. So, taking $R_x$ such that $\lambda^*(A(x))\sqrt{R}=l_x/2$, then for every $R<R_x$ and $|y-x|<\lambda^*(A(x))\sqrt R$ then there exists a unique $\gamma$ such that $\gamma + r( A(x)^+ \gamma)=y-x$ and moreover, $|\gamma|\leq 2|x-y|$. Now, using \eqref{Norm3} \[ |r( A(x)^+ \gamma)|_{A_R(x)}\leq \frac{\lambda^*(A(x)) |r( A(x)^+ \gamma)|}{R} \leq C_x\frac{|A(x)^+ \gamma|^2}{R}\leq C_x\frac{|x-y|^2}{R} \leq C_x |x-y|_{A_R(x)}^2. \] Since $\gamma =x-y-r(A(x)^+ \gamma)$, \[ |\gamma|_{A_R(x)}\leq |x-y|_{A_R(x)}+C_x|x-y|_{A_R(x)}^2\leq C_x, \] (using $|x-y|_{A_R(x)}\leq 1$). We have $|D_R^{-1}\theta|=|D_R^{-1} A(x)^+ \gamma|$. Since $A_R^+ A_R(x)=A_R^+(x) A(x) D_R$ is an orthogonal projection and $AA^+(x)$ is the identity, \[ |D_R^{-1}\theta|\leq |D_R^{-1} A(x)^+ \gamma|\leq A |A_R^+(x) A(x) D_R \,D_R^{-1} A(x)^+ \gamma| =|A_R^+(x) \gamma|= |\gamma|_{A_R(x)}. \] So $|D_R^{-1}\theta|\leq C_x$, and this implies $d_*(x,y)\leq C_x \sqrt{R}$. \smallskip \textbf{C.} The proof immediately follows from the previous items. \end{proof} The proof of Theorem \ref{tubecontrold} is now an immediate consequence of Theorem \ref{mttubesS} and Theorem \ref{NORM4}. The only apparent problem is that in Theorem \ref{NORM4} the global estimate \eqref{bound} is required, whereas in Theorem \ref{tubecontrold} the local estimate $(H_1)$ in \eqref{Not6} holds. But this is not really a problem, since it can be handled as already done for Theorem \ref{tubecontrold} (see Remark \ref{2}). \bibliographystyle{plain}
082e71d83a6ddf724c44f21cdde8c9773f4c8e97
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction.}\label{intro} \section{Introduction} A central modeling primitive in mathematical optimization is the disjunctive constraint: any feasible solution must satisfy at least one of some fixed, finite collection of alternatives. This type of constraint is general enough to capture structures as diverse as boolean satisfiability, complementarity constraints, special ordered sets, and (bounded) integrality. The special case of polyhedral disjunctive constraints corresponds to the form \begin{equation}\label{disconst} x \in \bigcup\nolimits_{i=1}^d P^i, \end{equation} where we have that each $P^i\subseteq \mathbb{R}^n$ is a polyhedron. In this work, we will focus on $\mathscr{V}$-polyhedra; that is, we have a description of the $P^i$ in terms of their extreme points $\operatorname{ext}(P^i)$. We are particularly interested in the case where constraint \eqref{disconst} is \emph{primitive}, or a basic building block for a much more complex optimization problem. For this reason, we are interested in modeling it in a generic, composable way. In particular, if \eqref{disconst} is embedded in a larger, more complex optimization problem \begin{equation} \label{eqn:larger-optimization-problem} \min_{(x,y) \in Q : \eqref{disconst}} f(x,y), \end{equation} we hope for a mathematical description sufficiently structured such that we may use more advanced algorithmic approaches, beyond na\"ive enumeration, to solve \eqref{eqn:larger-optimization-problem}. In particular, $Q$ could be described by any number of different types of constraints: linear inequalities, conic constraints, integrality conditions, or additional disjunctive constraints. In this context, it is well known that merely constructing the convex hull $\operatorname{Conv}(\bigcup_{i=1}^d P^i)$ is not sufficient for solving \eqref{eqn:larger-optimization-problem}; we will need a formulation for $\bigcup_{i=1}^d P^i$ directly. Mixed-integer programming (MIP) has emerged as an incredibly expressive modeling methodology, with advanced computational methods capable of solving many problems of practical interest, often at very large scale \cite{Bixby:2007,Junger:2010}. Constraint \eqref{disconst} with polyhedral sets $P^i$ is particularly well-suited for a mixed-integer programming approach. Indeed, standard formulations for \eqref{disconst} were presented in \cite{Jeroslow:1984}, and are \emph{ideal}, or as strong as possible with respect to their continuous linear programming relaxations (see Section~\ref{ss:formulation-definitions} for a formal definition). However, this formulation requires introducing $d$ auxiliary binary variables, which may be impractically large, especially in the context of the larger problem~\eqref{eqn:larger-optimization-problem}. However, it is sometimes possible to construct ideal formulations with considerably fewer auxiliary variables. In particular, a string of recent work \cite{Adams:2012,Muldoon:2013,Vielma:2016,Vielma:2009a} has presented ideal formulations for certain highly structured constraints such as SOS2 \cite{Beale:1970} with only $\mathscr{O}(\log(d))$ auxiliary binary variables and additional constraints (excluding variable bounds). Moreover, these formulations have proven practically useful, and indeed are the most performant by a significant margin for a large swath of instances of the problem classes to which they have been applied \cite{Vielma:2010}. However, building these formulations is complex and ad-hoc, hindering the construction, analysis, and implementation of such formulations for new constraints. One relatively generic and versatile approach to construct small ideal formulations is the \emph{independent branching} (IB) scheme framework introduced by Vielma and Nemhauser \citep{Vielma:2009a}. The approach is to find some (particularly structured) polyhedra $Q^{1,j}$ and $Q^{2,j}$ such that \eqref{disconst} can be rewritten as \begin{equation}\label{eqn:disjunction-via-IB} \bigcup\nolimits_{i=1}^d P^i = \bigcap\nolimits_{j=1}^t \left(Q^{1,j} \cup Q^{2,j}\right). \end{equation} This represents the disjunctive constraint in term of a series of simple choices between two alternatives. Given such a representation, it is often straightforward to construct a simple, small, and ideal formulation for \eqref{disconst} by formulating each of the $t$ alternatives separately, and then combining them. Furthermore, when the polyhedra $P^i$ are $\mathscr{V}$-polyhedra, the construction of the independent branching scheme-based formulation is purely combinatorial, based on the extreme points that are shared between the different polyhedra $P^i$. As we will see, we can therefore approach formulating \eqref{disconst} combinatorially, by studying the shared structure amongst the extreme points. In this work we generalize and provide a systematic study of the applicability and limitations of the independent branching approach. The contributions of this work can be categorized in the following way. \begin{enumerate} \item We generalize the notion of independent branching schemes to allow for multiple alternatives, and provide an exact characterization of when there exists \emph{any} independent branching representation for \eqref{disconst}, in terms of the graphical representation of the shared extreme points amongst the polyhedra $P^i$. In particular: \begin{enumerate} \item We demonstrate that the widely-used cardinality constraints cannot be expressed by any independent branching scheme with few alternatives. We argue that this negative result provides theoretical justification for the practical observation that both MIP formulations and simple constraint branching schemes struggle with modeling cardinality constraints effectively. \item We show that arbitrary piecewise linear functions in the plane can be modeled with at most three alternatives, and provide a polynomial-time verifiable condition for representability with two alternatives. \item We argue that nonconvex polygonal set avoidance constraints are always representable with two alternatives. \end{enumerate} \item We provide an exact characterization for when there exists a two-alternative independent branching representation for \eqref{disconst} of size $t$, in terms of the classical biclique covering problem. This relation allows the \emph{algorithmic} construction of small independent branching formulations for~\eqref{disconst}. In particular, we study and apply simple properties of biclique covers and their composition to systematically construct explicit descriptions of small independent branching formulations for special structures. \item We apply our framework to a variety of constraints of the form \eqref{disconst} to give an indication of the expressive power of the IB scheme approach and the advantage of using biclique cover techniques to construct formulations. In particular: \begin{enumerate} \item We review and develop several generic properties of biclique covers that lead to systematic construction techniques. Using these techniques we construct explicit, small (i.e. logarithmic in $d$), ideal formulations for generalizations of the special ordered sets of Beale and Tomlin \cite{Beale:1970}, piecewise linear functions over arbitrary 2-dimensional grid triangulations, and outer-approximating discretizations of multilinear terms. This last formulation generalizes the popular logarithmically-sized formulation of Misener et al. \cite{Misener:2012,Misener:2011} for nonconvex quadratic optimization, which we show is not ideal in general. \item We provide matching lower bounds for these constructions, showing that they are asymptotically optimal with respect to the size of \emph{any} possible MIP formulation. \end{enumerate} \end{enumerate} \section{Preliminaries: Definitions, notation, and nomenclature}\label{prelimsec} A (bounded) $\mathscr{V}$-polyhedra (or polyhedra in $\mathscr{V}$-form) is a set $P \subset \mathbb{R}^n$ that can be expressed as \[ P = \operatorname{Conv}(V) \defeq \left\{\sum_{v \in V} \lambda_v v : \lambda \in \Delta^V \right\} \] for some finite set of vectors $V \subset \mathbb{R}^n$, where $\Delta^V \defeq \{\lambda \in \mathbb{R}_{+}^V : \sum_{v \in V} \lambda_v = 1\}$ is the standard simplex. According to the celebrated Minkowski-Weyl Theorem (e.g. \cite[Corollary 3.14]{Conforti:2014}), any polyhedral disjunctive constraint \eqref{disconst} can be expressed as the union of $\mathscr{V}$-polyhedra in terms of their extreme points $\operatorname{ext}(P^i)$\footnote{For the moment we are assuming that the $P^i$ are bounded; the unbounded case is more delicate, as we will discuss shortly.}. When the disjunctive constraint \eqref{disconst} is a union of $\mathscr{V}$-polyhedra, it suffices to consider only the combinatorial structure of the extreme points of the polyhedra $P^i$. To see why, consider $J = \bigcup_{i=1}^d \operatorname{ext}(P^i)$ as the \emph{ground set} and take $\mathscr{S} = \{\operatorname{ext}(P^i)\}_{i=1}^d \subseteq 2^J$ as the collection of extreme points for each of the polyhedra. We can then define a corresponding disjunctive constraint that is purely combinatorial on the sets $\mathscr{S}$. \begin{definition}\label{CDCDef} A \emph{combinatorial disjunctive constraint} (CDC) induced by the sets $\mathscr{S}$ is \[ \lambda \in \operatorname{CDC}(\mathscr{S}) \defeq \bigcup_{S \in \mathscr{S}} Q(S), \] where $Q(S) \defeq \{\lambda \in \Delta^J : \lambda_{J \backslash S} \leq 0\}$ is the face that $S \subseteq J$ induces on the standard simplex. \end{definition} Combinatorial disjunctive constraints may also appear as natural primitive constraints that do not explicitly arise from unions of $\mathscr{V}$-polyhedra, as we will see in Sections~\ref{soskintrosec} and \ref{cardintrosec}. However, when they do arise from unions of $\mathscr{V}$-polyhedra, it is straightforward to construct a corresponding formulation for \eqref{disconst} as \begin{equation} \label{eqn:disaggregated-CDC-to-DC} \left\{\sum_{v \in J} \lambda_v v \: : \: \lambda \in \operatorname{CDC}(\mathscr{S}) \right\}. \end{equation} One advantage of this approach is that formulation \eqref{eqn:disaggregated-CDC-to-DC} allows us to divorce the problem-specific data (i.e. the values $v \in J$) from the underlying combinatorial structure encapsulated in $\operatorname{CDC}(\mathscr{S})$. As such, we can construct a single, strong formulation for a given structure $\operatorname{CDC}(\mathscr{S})$ and this formulation will remain valid for transformations of the data, so long as this transformation sufficiently preserves the combinatorial structure of $\operatorname{CDC}(\mathscr{S})$. For instance, if $\set{P^i}_{i=1}^d$ are the polyhedra for the original constraint represented by $\operatorname{CDC}(\mathscr{S})$, and $\{\hat{P}^i\}_{i=1}^d$ are those associated with the new data, then a sufficient condition for the formulation of $\operatorname{CDC}(\mathscr{S})$ yielding a valid formulation for $\bigcup_{i=1}^d \hat{P}^i$ is the existence of a bijection $\pi:J \to \hat{J}$ (with $J = \bigcup\nolimits_{i=1}^d \operatorname{ext}(P^i)$ and $\hat{J} = \bigcup\nolimits_{i=1}^d \operatorname{ext}(\hat{P}^i)$) such that \begin{equation}\label{biyectioncond} v \in \operatorname{ext}(P^i) \Longleftrightarrow \pi(v) \in \operatorname{ext}(\hat{P}^i) \quad \forall i \in \llbracket d \rrbracket, \: v \in J, \end{equation} where $\llbracket d \rrbracket\defeq \{1,\ldots,d\}$. In this way, we can construct a single small, strong formulation for $\operatorname{CDC}(\mathscr{S})$, and use it repeatedly for many different ``combinatorially equivalent'' instances of the same constraint. We note that one subtle disadvantage of this data-agnostic approach is that, even if condition \eqref{biyectioncond} is satisfied, the resulting formulation for $\bigcup_{i=1}^d \hat{P}^i$ may be larger than necessary. An extreme manifestation of this would be when the new polyhedra $\{\hat{P}^i\}_{i=1}^d$ are such that $\hat{P}_i\subseteq \hat{P}_1$ for all $i\in \llbracket d \rrbracket$. In this case, $\bigcup\nolimits_{i=1}^d \hat{P}^i = \hat{P}^1$, and so formulating this does not require a MIP formulation at all. Less pathological cases could occur where some subset of the disjunctive sets become redundant after changing the problem data. However, we note that in many of the applications considered in this work, the combinatorial representation leads to redundancy of this form only in rare pathological cases (e.g. Sections~\ref{sos2introsec} and \ref{gridintrosec}). In the remaining cases we will take care to consider, for example, the geometric structure of the data before constructing the disjunctive constraint (e.g. Section~\ref{ss:partitions-of-the-plane}). Finally, we also note that if we wish to model the case where the polyhedra $P^i$ are unbounded, a result of Jeroslow and Lowe \cite{Jeroslow:1984} \cite[Proposition 11.2]{Vielma:2015} tells us that we may only construct a (binary) MIP formulation for \eqref{disconst} if the recession cones coincide for each $P^i$. In the case this condition is met, we may formulate \eqref{disconst} with \begin{equation} \label{eqn:disaggregated-CDC-to-DC-with-rays} \left\{\sum_{v \in J} \lambda_v v + \sum_{r \in R} \mu_r r \: : \: \lambda \in \operatorname{CDC}(\mathscr{S}), \: \mu \in \mathbb{R}^R_{+} \right\}, \end{equation} where $R$ is the shared set of extreme rays for each of the $P^i$. Therefore, we will restrict our attention to the case where each of the $P^i$ are bounded, as formulating the unbounded case is a straightforward extension. In the remainder of the paper, we will make the following assumptions on $\mathscr{S}$ that are without loss of generality. \begin{assumption}\label{basicassumption} We assume the following about $\mathscr{S}$. \begin{itemize} \item $\mathscr{S}$ is irredundant: there do not exist distinct $S,T \in \mathscr{S}$ such that $S \subseteq T$. \item $\mathscr{S}$ covers the ground set: $\bigcup_{S \in \mathscr{S}} S = J$. \end{itemize} \end{assumption} We will say that a set $S \subseteq J$ is a \emph{feasible set} with respect to $\operatorname{CDC}(\mathscr{S})$ if $Q(S) \subseteq \operatorname{CDC}(\mathscr{S})$ (equivalently, if $S \subseteq T$ for some $T \in \mathscr{S}$) and that it is an \emph{infeasible set} otherwise. \section{Motivating examples} \label{sec:motivating-examples} As mentioned, we can represent any polyhedral disjunctive constraint \eqref{disconst} as the union of $\mathscr{V}$-polyhedra. However, there are many disjunctive constraints for which the $\mathscr{V}$-form of \eqref{disconst} is especially natural. We now present some as running examples that we will return to throughout. One motif to appear repeatedly will be the graph of a continuous piecewise linear function. That is, given some bounded domain $\Omega \subset \mathbb{R}^n$ and some polyhedral partition $\bigcup_{i=1}^d P^i = \Omega$ (i.e. the relative interiors do not overlap\footnote{Formally, $\operatorname{relint}(P^i) \cap \operatorname{relint}(P^j) = \emptyset$ for each $i \neq j$.}), we are interested in modeling a continuous function $f : \Omega \rightarrow \mathbb{R}$ such that $x \in P^i \Longrightarrow f(x) = a^i \cdot x + b_i$ for some appropriate $a^i \in \mathbb{R}^n$ and $b \in \mathbb{R}$. In order to model the graph $\operatorname{gr}(f;\Omega) \defeq \{(x,f(x)) : x \in \Omega\}$, we can construct a formulation for $\operatorname{CDC}(\mathscr{S})$ (with $\mathscr{S} = \{\operatorname{ext}(P^i)\}_{i=1}^d$) and express \begin{equation} \label{eqn:pwl-graph} \operatorname{gr}(f) = \left\{ \sum_{v \in J} \lambda_v (v,f(v)) : \lambda \in \operatorname{CDC}(\mathscr{S}) \right\}, \end{equation} where we will use the notation $\operatorname{gr}(f) \equiv \operatorname{gr}(f;\Omega)$ when $\Omega$ is clear from context.\footnote{We note that the results to follow can potentially be extended to certain discontinuous piecewise linear functions by working instead with the epigraph of $f$; we point the interested reader to \cite{Vielma:2010,Vielma:2008a} for further discussion.} \subsection{Univariate piecewise linear functions and the SOS2 constraint}\label{sos2introsec} Consider a univariate (nonconvex) piecewise linear function $f$ characterized by $N$ breakpoints $x^1 < x^2 < \ldots < x^N$. We may model the graph of this function via \eqref{eqn:pwl-graph}, where $\mathscr{S} = \{\{x^j,x^{j+1}\} : j \in \llbracket N-1 \rrbracket\}$. As long as the ordering of the breakpoints is preserved, it is easy to see that condition \eqref{biyectioncond} will be satisfied for any transformations of the problem data. Furthermore, the only case in which knowledge of the specific data $\{(x^j,f(x^j))\}_{j=1}^n$ allows the simplification of the original disjunctive representation of $\operatorname{gr}(f)$ is when $f$ is affine in on two adjacent intervals, e.g. affine over $[x^j,x^{j+2}]$ for some $j \in \llbracket N-2 \rrbracket$. Therefore, the potential disadvantage of disregarding the specific data when formulating the constraint occurs only in rare pathological cases which are easy to detect. For this reason, we will strip out the problem data and instead express the constraint with respect to the indices $j$ of the vertices $\{x^j\}_{j=1}^N$. That is, we take $J = \llbracket N \rrbracket $ and write $\mathscr{S} = \{\{\tau,\tau+1\} : \tau \in \llbracket N-1 \rrbracket\}$ to emphasize the data independence of the constraint and highlight the combinatorial structure. In this form, we can recognize the special ordered set of type 2 (SOS2$(N)$) constraint of Beale and Tomlin \cite{Beale:1970}, which requires that at most two components of $\lambda$ may be nonzero, and that these nonzero components must be consecutive in the ordering on $J$. \subsection{SOS$k$}\label{soskintrosec} A generalization of the special ordered sets considers the case where at most $k$ consecutive components of $\lambda$ may be nonzero at once. In particular, if $J = \llbracket N \rrbracket$, we have $\mathscr{S} = \{\{\tau,\tau+1,\ldots,\tau+k-1\} : \tau \in \llbracket N - k + 1 \rrbracket\}$. This constraint may arise, for example, in chemical process scheduling problems, where an activated machine may only be on for $k$ consecutive time units and must produce a fixed quantity during that period \cite{Floudas:2005,Kondili:1993}. \subsection{Cardinality constraints}\label{cardintrosec} An extremely common constraint in optimization is the cardinality constraint of degree $\ell$, where at most $\ell$ components of $\lambda$ may be nonzero. This corresponds to $\mathscr{S} = \{I \subseteq J : |I| = \ell\}$. A particularly compelling application of the cardinality constraint is in portfolio optimization \cite{Bertsimas:2007a,Bienstock:1996,Chang:2000,Vielma:2008}, where it is often advantageous to limit the number of investments to some fixed number $\ell$ to minimize transaction costs, or to allow differentiation from the performance of the market as a whole. \subsection{Discretizations of multilinear terms}\label{multilinearsec} Consider a multilinear function $f(x_1,\ldots,x_\eta) = \prod_{i=1}^\eta x_i$ defined over some box domain $\Omega \defeq [l,u] \subset \mathbb{R}^\eta$. This function appears often in optimization models \cite{Foulds:1992}, but is nonconvex, and often leads to problems which are difficult to solve to global optimality in practice \cite{Androulakis:1995,Quesada:1995,Wicaksono:2008}. As a result, computational techniques will often ``relax'' the graph of the function $\operatorname{gr}(f)$ with a convex outer approximation, which is easier to optimize over \cite{Sahinidis:1996}.\footnote{Note that these relaxations are useful in the context of global optimization, coupled with algorithmic techniques such as spatial branch-and-bound.} For the bilinear case ($\eta=2$), the well-known McCormick envelope \cite{McCormick:1976} describes the convex hull of $\operatorname{gr}(f)$. Although traditionally stated in an inequality description, we may equivalently describe the convex hull via its four extreme points, which are readily available in closed form. For higher-dimensional multilinear terms, the convex hull has $2^\eta$ extreme points, and can be constructed in a similar manner (e.g. see equation (3) in~\cite{luedtke2012some} and the associated references). Misener et al. \cite{Misener:2012,Misener:2011} propose a computational technique for optimizing problems with bilinear terms where, instead of modeling the graph over a single region $\Omega = [l,u] \subset \mathbb{R}^2$, they discretize the region in a regular fashion and apply the McCormick envelope to each subregion. They model this constraint as a union of polyhedra, where each subregion enjoys a tighter relaxation of the bilinear term. Additionally, they propose a logarithmically-sized formulation for the union. However, it is not ideal (see Appendix~\ref{app:misener}), it only applies for bilinear terms ($\eta=2$), and it is specialized for a particular type of discretization (namely, only discretizing along one component $x_1$, and with constant discretization widths $\{h^1_{j+1}-h^1_{j}\}_{j=1}^{d_1-1}$). For a more general setting, we have that the extreme points of the convex hull of the graph $\operatorname{Conv}(\operatorname{gr}(f))$ are given by $\{(x,f(x)) : x \in \operatorname{ext}(\Omega)\}$ \cite[equation (3)]{luedtke2012some}, where it is easy to see that $\operatorname{ext}(\Omega) = \prod_{i=1}^\eta \{l_i,u_i\}$. Consider a grid imposed on $[l,u] \subset \mathbb{R}^\eta$; that is, along each component $i \in \llbracket \eta \rrbracket$, we partition $[l_i,u_i]$ along the points $l_i \equiv h_1^i < h_2^i < \cdots < h_{d_i-1}^i < h_{d_i}^i \equiv u_i$. This yields $\prod_{i=1}^\eta \bra{d_i-1}$ subregions; denote them by $\mathscr{R} \defeq \left\{\prod_{i=1}^\eta [h^i_{k_i},h^i_{k_i+1}] : k \in \prod_{i=1}^\eta \llbracket d_i -1 \rrbracket\right\}$. We can then take the polyhedral partition of $\Omega$ given by $P^R \defeq \operatorname{Conv}(\operatorname{gr}(f;R))$ for each subregion $R$, the sets as $\mathscr{S} = \left\{\operatorname{ext}(P^R)\right\}_{R \in \mathscr{R}}$, and the ground set as $J = \bigcup\{S \in \mathscr{S}\}$. In particular, we have that $J = \prod_{i=1}^\eta \{h^i_1,\ldots,h^i_{d_i}\}$. Analogously to the notational simplification we took with the SOS2 constraint, for the remainder we will take $J=\prod_{i=1}^\eta \llbracket d_i \rrbracket$ and $\mathscr{S}=\set{\prod_{i=1}^\eta \set{k_i,k_i+1} : k \in \prod_{i=1}^\eta \llbracket d_i -1 \rrbracket}$. We also note that condition \eqref{biyectioncond} is satisfied as long as the ordering of the discretization is respected along each dimension. \subsection{Piecewise linear functions in the plane and grid triangulations}\label{gridintrosec} Consider a (potentially nonconvex) region $\Omega \subset \mathbb{R}^2$. We would like to model a (also potentially nonconvex) piecewise linear function $f$ with domain over $\Omega$. Take $\{P^i\}_{i=1}^d$ as the set of pieces of the domain, and the corresponding ground set as $J = \bigcup_{i=1}^d \operatorname{ext}(P^i)$ and sets as $\mathscr{S} = \{\operatorname{ext}(P^i)\}_{i=1}^d$. We may then model the piecewise linear function via the graph representation \eqref{eqn:pwl-graph}. An important special case occurs when the function $f$ is affine over a triangulation of a grid similar to the one used for multilinear terms. A description of the associated CDC that emphasizes its combinatorial structure can be obtained through the same simplification from Section~\ref{multilinearsec} as follows. Consider a rectangular region in the plane $\Omega = [1,M] \times [1,N]$, and the regular grid points $J = \{1,\ldots,M\} \times \{1,\ldots,N\}$. A \emph{grid triangulation} $\mathscr{S}$ of $\Omega$ is then a set $\mathscr{S}$ where: \begin{itemize} \item Each $S \in \mathscr{S}$ is a triangle: $|S|=3$. \item $\mathscr{S}$ partitions $\Omega$: $\bigcup_{S \in \mathscr{S}} \operatorname{Conv}(S) = \Omega$ and $\operatorname{relint}(\operatorname{Conv}(S)) \cap \operatorname{relint}(\operatorname{Conv}(T)) = \emptyset$ for each distinct $S, T \in \mathscr{S}$. \item $\mathscr{S}$ is on a regular grid: $S \subset J$ for each $S \in \mathscr{S}$, and $||v-w||_\infty \leq 1$ for each $v,w \in S$. \end{itemize} As concrete examples, consider Figure~\ref{fig:triangulations}, where we depict three different triangulations with $M=N=3$. Grid triangulations are often used to model bivariate, non-separable, piecewise linear functions using \eqref{eqn:pwl-graph}~\cite{Vielma:2010,Vielma:2009a}. Furthermore, any formulation constructed for a given grid triangulation can be readily applied to any other grid triangulation obtained by shifting the grid points in the plane, so long as the resulting triangulation is strongly isomorphic to, or compatible with, the original triangulation \cite{aichholzer2003towards}. \begin{figure}[htpb] \centering \includegraphics[width=.25\linewidth]{Fig1a.pdf} \includegraphics[width=.25\linewidth]{Fig1b.pdf} \includegraphics[width=.25\linewidth]{Fig1c.pdf} \caption{Three grid triangulations of $\Omega = [1,3]\times[1,3]$: the Union Jack (J1) \cite{Todd:1977} (Left), the K1 \cite{Kuhn:1960} (Center), and a more idiosyncratic construction (Right).} \label{fig:triangulations} \end{figure} \subsection{Obstacle avoidance}\label{obstacleavoidancesec} Consider an unmanned aerial vehicle (UAV) which you would like to navigate through an area with fixed obstacles. At any given time, you wish to impose the constraint that the location of the vehicle $x \in \mathbb{R}^2$ must lie in some (nonconvex) region $\Omega \subset \mathbb{R}^2$, which is the plane, less any obstacles in the area. MIP formulations of this constraint has received interest as a useful primitive for path planning \cite{Bellingham:2002,Deits:2015,Prodan:2016,Mellinger:2012}. We may model $x \in \Omega$ by partitioning the region $\Omega$ with polyhedra such that $\Omega = \bigcup_{i=1}^d P^i$. Traditional approaches to modeling constraints \eqref{disconst} of this form use a linear inequality description for each of the polyhedra $P^i$ and construct a corresponding big-$M$ formulation \cite{Prodan:2012,Prodan:2016}, which will not be ideal in general. In the $\mathscr{V}$-polyhedra framework, we will instead take the $P^i$ as $\mathscr{V}$-polyhedra that partition $\Omega$, and be able to construct small, ideal formulations. \section{MIP formulations for combinatorial disjunctive constraints} Using standard MIP formulation techniques, we now present formulations for $\operatorname{CDC}(\mathscr{S})$ for comparison with our approach. In particular, we will argue that the framework we will present later can lead to formulations that are smaller (in terms of the number of auxiliary variables), and enjoy other favorable properties we enumerate in Section \ref{ss:IB-schemes}. \subsection{MIP formulations: Definitions, size, and strength}\label{ss:formulation-definitions} Formally, we say that a (binary) MIP formulation $F$ for a constraint $x \in Q \subseteq \mathbb{R}^{n_1}$ is the composition of linear inequalities (the \emph{linear programming (LP) relaxation}, or just \emph{relaxation} in the context of this work) \[ R \defeq \left\{ (x,y,z) \in [l^x,u^x] \times [l^y,u^y] \times [0,1]^{n_3} : Ax + By + Cz \leq d \right\} \] with (binary) integrality conditions \begin{equation}\label{eqn:generic-MIP-formulation} F \defeq R \cap \left(\mathbb{R}^{n_1} \times \mathbb{R}^{n_2} \times \{0,1\}^{n_3}\right) \end{equation} such that $\operatorname{Proj}_x(F) = Q$.\footnote{To handle the unbounded case, we allow the variable bounds $l^x \leq x \leq u^x$ and $l^y \leq y \leq u^y$ to take infinite values.} Finally, we assume $R$ is line-free (i.e. has at least one extreme point), which is satisfied by essentially all practical formulations. Throughout, we will be interested in ways of understanding both the strength of a given formulation, as well as in quantifying the size or complexity of a formulation. We say that a formulation is \emph{ideal} if each extreme point of the relaxation naturally satisfies the integrality conditions, i.e. $\operatorname{Proj}_z(\operatorname{ext}(R)) \subseteq \{0,1\}^{n_3}$. The choice of name is apt, as this is the strongest possible MIP formulation we can expect.\footnote{An ideal formulation is also \emph{sharp}, i.e. its relaxation projects down to the convex hull of the set we are formulating ($\operatorname{Proj}_x(R) = \operatorname{Conv}(Q)$).} As a measure of the complexity of the formulation, we count the number of auxiliary continuous variables $y$ and continuous binary variables $z$ used by the formulation, as well as the number of inequalities in our description of $R$. We ignore the size of $x$, since this is intrinsic to the constraint we wish to model. We will say that a formulation is \emph{extended} if there are auxiliary continuous variables $y$ in the representation $F$ (that is, $n_2 > 0$) and \emph{non-extended} otherwise ($n_2 = 0$). Furthermore, as suggested by the definition of $R$, we distinguish between variable bounds (e.g. $l^y \leq y \leq u^y$) and general inequalities ($Ax + By + Cz \leq d$), as modern MIP solvers are able to incorporate variable bounds with minimal extra computational cost. Finally, we note that from now on (with a single exception), any statements regarding formulation properties or size will be with respect to the combinatorial disjunctive constraint $\operatorname{CDC}(\mathscr{S})$, rather than the $\mathscr{V}$-polyhedral disjunctive constraints itself. In other words, all statements are with respect a fixed combinatorial structure, with no possible simplifications from the specific data as discussed in Section~\ref{prelimsec}. The sole exception will be in Section~\ref{ss:partitions-of-the-plane}, where we will make use of some geometric properties inherent in a given realization of the data. \subsection{Existing formulations for combinatorial disjunctive constraints} A standard formulation for $\operatorname{CDC}(\mathscr{S})$ adapted from Jeroslow and Lowe \cite{Jeroslow:1984} is \begin{subequations} \label{eqn:jeroslow-formulation} \begin{align} \lambda_v = \sum_{S \in \mathscr{S} : v \in S} \gamma^S_v &\quad\quad \forall v \in J \\ z_S = \sum_{v \in S} \gamma^S_v &\quad\quad \forall S \in \mathscr{S} \\ \sum_{S \in \mathscr{S}} z_S = 1 &\\ \gamma^S \in \Delta^S &\quad\quad \forall S \in \mathscr{S} \\ (\lambda,z) \in \Delta^J \times \{0,1\}^\mathscr{S}. \end{align} \end{subequations} This formulation has $d=|\mathscr{S}|$ auxiliary binary variables, $\sum_{S \in \mathscr{S}} |S|$ auxiliary continuous variables, and no general inequalities. Additionally, it is ideal. Using Proposition 9.3 from \cite{Vielma:2015}, we can construct an ideal MIP formulation with fewer auxiliary binary variables: \begin{subequations} \label{eqn:prop-9.3} \begin{align} \lambda_v = \sum_{S \in \mathscr{S} : v \in S} \gamma^S_v &\quad\quad \forall v \in J \\ \sum_{S \in \mathscr{S}} \sum_{v \in S} \gamma^S_v = 1 & \\ \sum_{S \in \mathscr{S}} \sum_{v \in S} h^S \gamma^S_v = z & \\ \gamma^S \geq 0 &\quad\quad \forall S \in \mathscr{S} \\ z \in \{0,1\}^r,& \end{align} \end{subequations} where $\{h^S\}_{S \in \mathscr{S}} \subseteq \{0,1\}^r$ is some set of distinct binary vectors. This formulation is actually a generalization of \eqref{eqn:jeroslow-formulation}, which we recover if we take $h^S = \textbf{e}^S \in \mathbb{R}^\mathscr{S}$ as the canonical unit vectors. If instead we take $r$ to be as small as possible (while ensuring that the vectors $\{h^S\}_{S \in \mathscr{S}}$ are distinct), we recover $r = \lceil \log_2(d) \rceil$. Therefore, formulation \eqref{eqn:prop-9.3} yields an ideal extended formulation for \eqref{disconst} with $\lceil \log_2(d) \rceil$ auxiliary binary variables, $\sum_{S \in \mathscr{S}} |S|$ auxiliary continuous variables, and no general inequalities. The following corollary shows that this is the smallest number of auxiliary binary variables we may hope for. \begin{proposition} \label{prop:log-bound} If the sets $\mathscr{S}$ are irredundant, then any binary MIP formulation for $\operatorname{CDC}(\mathscr{S})$ must have at least $\lceil\log_2(d)\rceil$ auxiliary binary variables. \end{proposition} \proof{\textbf{Proof}} See Appendix~\ref{app:log-bound}. \Halmos\endproof The formulations thus far have been extended formulations, as they are constructed by formulating each polyhedra $Q\bra{S}$ separately and then aggregating them, rather than working with the combinatorial structure underlying the shared extreme points. Therefore, each of these formulations requires a copy of the multiplier $\gamma^S_v$ for each set $S \in \mathscr{S}$ for which $v \in S$, and so $\sum_{i=1}^d |\mathscr{S}|$ auxiliary continuous variables total. In contrast, we can construct non-extended formulations for $\operatorname{CDC}(\mathscr{S})$ that work directly on the $\lambda$ variables and the underlying combinatorial structure of $\mathscr{S}$. An example of a non-extended formulation for CDC is the widely used ad-hoc formulation (see \cite[Section 6]{Vielma:2015} and the references therein) given by \begin{subequations} \label{eqn:disaggregated} \begin{align} \lambda_v \leq \sum_{S \in \mathscr{S}: v \in S} z_S &\quad\quad \forall v \in J \label{eqn:disaggregated-1} \\ \sum_{S \in \mathscr{S}} z_S = 1 \\ (\lambda,z) \in \Delta^J \times \{0,1\}^{\mathscr{S}}. \end{align} \end{subequations} This formulation is not necessarily ideal, and it requires no auxiliary continuous variables, $d$ auxiliary binary variables, and $|J|$ general inequalities. In summary, we have seen an ideal extended formulation \eqref{eqn:prop-9.3} for $\operatorname{CDC}(\mathscr{S})$ with relatively few auxiliary binary variables, but relatively many auxiliary continuous variables. On the other end of the spectrum, we have a non-extended formulation \eqref{eqn:disaggregated} with no auxiliary continuous variables, but which requires relatively many auxiliary binary variables and which may fail to be ideal. However, we know that in special cases we can construct ideal, non-extended formulations with only $\mathscr{O}(\log(d))$ auxiliary variables and constraints (e.g. SOS1, SOS2, and particular 2-dimensional grid triangulations \cite{Vielma:2016,Vielma:2009a}). This work provides a framework for constructing such \emph{small, strong}, non-extended MIP formulations for $\operatorname{CDC}(\mathscr{S})$, which are automatically ideal, and in the best case will have $\mathscr{O}(\log(d))$ auxiliary binary variables and general inequality constraints. \section{Independent branching schemes} \label{ss:IB-schemes} Vielma and Nemhauser \cite{Vielma:2009a} introduced the notion of an \emph{independent branching scheme} as a natural framework for constructing formulations for combinatorial disjunctive constraints. The independent branching scheme is a logically equivalent way of expressing a CDC in terms of a conjunction of dichotomies: that is, as a series of choices between two (simple) options. This approach is parsimonious: if you are given an independent branching scheme for a particular CDC, it is straightforward to construct an ideal formulation whose size is on the order of the number of dichotomies. For our purposes, we present a generalized notion, where we allow potentially more than two alternatives. \begin{definition} A \emph{$k$-way independent branching scheme} for $\operatorname{CDC}(\mathscr{S})$ is given by a family of sets $(L^j_1,\ldots,L^j_k)$ (where each $L^j_i \subseteq J$) for $j \in \llbracket t \rrbracket$, where \begin{equation} \label{eqn:multi-way-CDC} \operatorname{CDC}(\mathscr{S}) = \bigcap_{j=1}^t \left( \bigcup_{i=1}^k Q(L^j_i) \right). \end{equation} \end{definition} We say that such an IB scheme has \emph{depth $t$}, and that each $j \in \llbracket t \rrbracket$ yields a corresponding \emph{level} of the IB scheme $\bigcup_{i=1}^k Q(L^j_i)$, given by the $k$ \emph{alternatives} $Q(L^j_i)$. An equivalent way of understanding these representations, which we will be using for the remainder of this work, is by eschewing the polyhedra $Q(L^j_i)$ and working directly on the underlying set $L^j_i$. That is, a valid $k$-way IB scheme satisfies the condition that \[ T \subseteq J \text{ is a feasible set} \Longleftrightarrow \forall j \in \llbracket t \rrbracket, \: \exists i \in \llbracket k \rrbracket \text{ s.t. } T \subseteq L^j_i. \] First, we observe that, due to our assumption that $\mathscr{S}$ covers the ground set, we have that for each element $v \in J$ and level $j \in \llbracket t \rrbracket$, there will be at least one alternative $i \in \llbracket k \rrbracket$ such that $v \in L^j_i$. We will use this extensively in the analysis to come, as it simplifies some otherwise tedious case analyses. Second, we see that this definition can capture potential schemes with a variable number of alternatives in each level by adding empty alternatives $L^j_i=\emptyset$, provided we take $k$ as the maximum number of alternatives for all levels. For notational simplicity, we say that a 2-way IB scheme is a \emph{pairwise IB scheme}, and in this case we write the sets as $\{(L^{j},R^{j})\}_{j=1}^t$ as in \cite{Vielma:2009a}. In contrast, we will call the case with $k > 2$ a \emph{multi-way IB scheme}. In this form, we have replaced the monolithic constraint $\operatorname{CDC}(\mathscr{S})$ by $t$ constraints, each of which require the selection between $k$ alternatives. We may then use standard techniques to construct a corresponding mixed-integer formulation. \begin{proposition} Given an independent branching scheme $\{(L_1^j,\ldots,L_k^j)\}_{j=1}^t$ for $\operatorname{CDC}(\mathscr{S})$, the following is a valid formulation for $\operatorname{CDC}(\mathscr{S})$: \begin{subequations} \label{eqn:multiway-formulation} \begin{align} \sum_{v \not\in L^j_i} \lambda_v \leq 1-z^j_i &\quad \forall j \in \llbracket t \rrbracket, \forall i \in \llbracket k \rrbracket \label{eqn:multiway-formulation-1} \\ \sum_{i=1}^k z^j_i = 1 &\quad \forall j \in \llbracket t \rrbracket \label{eqn:multiway-formulation-2} \\ \lambda \in \Delta^J &\\ z^j \in \{0,1\}^k &\quad \forall j \in \llbracket t \rrbracket. \end{align} \end{subequations} \end{proposition} The formulation is known to be ideal for $k=2$ \cite{Vielma:2010,Vielma:2009a}. It has no auxiliary continuous variables, $k t$ auxiliary binary variables, and $k t$ general inequalities. \subsection{Constraint branching via independent branching-based formulations}\label{constraintbranchingsec} The canonical algorithmic technique for solving mixed-integer programming problems is some variation of branch-and-bound \cite{Land:1960}, which implicitly enumerates all possible values for the binary variables. In its simplest form, a sequence of problems are solved, starting with the relaxation of the MIP formulation, after which a binary variable $z_i$ is chosen for \emph{branching}. That is, the current problem is branched into two subproblems: one with the additional constraint $z_i \leq 0$, another with $z_i \geq 1$. Repeating this procedure, the subproblems form a (binary) tree whose leaves correspond to all $2^{n_3}$ possible values for the $n_3$ binary variables in formulation~\eqref{eqn:generic-MIP-formulation}. At any given subproblem, the augmented relaxation to be solved is described by the set of binary variables fixed to zero, and the set of those fixed to one. The spirit of constraint branching is to allow richer branching decisions. For example, a branching decision might be between $k$ alternatives of the form $\{Q^i\}_{i=1}^k$, where each $Q^i$ is formed by adding a general inequality constraint to the existing relaxation at the current node. This concept has significant overlap with the broader field of constraint programming \cite{Apt:2003,Jaffar:1994}, which has been recognized and exploited in the mixed-integer programming literature \cite{Achterberg:2008,Appleget:2000,Hooker:2002,Ostrowski:2009,Ryan:1981}. More complex constraint branching can often lead to a more balanced branch-and-bound tree, which can significantly improve computational performance (see, for example, \cite[Section 8]{Vielma:2015} and \cite{Yildiz:2013} for more discussion). Combinatorial disjunctive constraints are a natural setting to apply constraint branching directly on the continuous $\lambda$ variables \cite{Beale:1970,DeFarias01,keha04,keha06,martin06}. Indeed, the classical examples of the SOS1 and SOS2 constraints~\cite{Beale:1970} show that we do not necessarily require a MIP formulation (or the auxiliary binary variables $z$) for modeling combinatorial disjunctive constraints, as the disjunction can be enforced directly through constraint branching on the $\lambda$ variables. These constraint branching approaches without auxiliary binary variables can be implemented in an ad-hoc branch-and-bound procedure, or through branching callbacks available in some MIP solvers, such as CPLEX. In theory, this approach should outperform a MIP formulation like \eqref{eqn:disaggregated} that introduces additional variables and constraints. However, realizing this performance advantage in practice can require significant effort and technical expertise. For instance, Vielma et al.~\cite{Vielma:2010,Vielma:2008a} observe that the basic formulation \eqref{eqn:disaggregated} clearly outperformed the SOS2 branching implementation in CPLEX v9.1. However, CPLEX v11 implemented an optimized version of SOS2 branching that used the advanced branch selection techniques available for variable branching, reversing this performance gap with respect the MIP formulation approach. One way to avoid re-implementing the advanced branching selection techniques for a new constraint branching approach is by constructing a MIP model that automatically inherits the advanced constraint branch selection, but using the solvers traditional variable branching~\cite{Appleget:2000,Vielma:2009a}. For simplicity, assume that the constraint branching approach has $t$ branching options, each of which creates $k$ branches, and that each constraint added has support on the $\lambda$ variables, with variable coefficients in $\set{0,1}$ and a zero right-hand-side. That is, branch $i \in \llbracket k \rrbracket$ of branching option $j \in \llbracket t \rrbracket$ adds a constraint of the form $\sum_{v \not\in L^j_i} \lambda_v \leq 0$. This is equivalent to a \emph{multi-variable} branching approach that fixes groups of variables to zero, and it includes as special cases most constraint branching approaches, including SOS1/SOS2 branching. Then \eqref{eqn:multiway-formulation} is a MIP formulation for this multi-variable constraint branching scheme, as variable branching on $\{z^j_i\}_{i=1}^k$ enforces constraint branching option $j$ on the $\lambda$ variables. This connection highlights the natural theoretical equivalence between a multi-variable branching and an independent branching formulation. A practical difference between the two is that direct multi-variable branching must implement an explicit branch selection and implementation routine, while an independent branching formulation inherits the variable branching selection and implementation routines of the MIP solver. The upshot of this is that the independent branching formulation must provide a complete catalog of all possible branching options up-front (i.e. through the formulation), while direct multi-variable branching can have a large catalog of branching options that are implicitly defined by the branching routines. We finally note that, as discussed by Vielma and Nemhauser~\cite[Section 3]{Vielma:2009a}, variable branching on a non-IB formulation such as \eqref{eqn:prop-9.3} can fix components of $\lambda$ to zero in a fashion that is \emph{dependent} on previous branching decisions. This is not the case with independent branching formulations, hence the name. We review this independence property in detail in Appendix~\ref{IBappendix}. \section{Independent branching scheme representability} To start, we observe that the independent branching approach is not sufficiently general to capture every possible formulation for $\operatorname{CDC}(\mathscr{S})$. In particular, there is the restriction that each alternative $Q(L^j_i)$ restricts the $\lambda$ variables to lie on a \emph{single} face of the standard simplex. A natural first question is then: given a family of sets $\mathscr{S}$, do any $k$-way IB schemes exist for $\operatorname{CDC}(\mathscr{S})$? We provide an answer, based on a graphical characterization of the constraint. \begin{definition} Let $H\defeq (J,\mathcal{E})$ be a hypergraph with hyperedge set $\mathcal{E}\subseteq 2^J$. \begin{itemize} \item The \emph{rank} of $H$ is $r\bra{H} \defeq \max\set{\abs{E}\,:\, E\in \mathcal{E}}$. \item A \emph{(weakly) independent set} of $H$ is a set $U\subseteq J$ that does not contain any element of $\mathcal{E}$ as a subset. \item The \emph{conflict hypergraph} of $\mathscr{S}$ is $H^c_\mathscr{S} \defeq (J,\mathcal{E}_\mathscr{S})$, where $\mathcal{E}_\mathscr{S} \defeq \{E\subseteq J : \text{ $E$ is a minimal infeasible set}\}$. \end{itemize} \end{definition} \begin{lemma} The maximal independent sets $S$ in $H^c_\mathscr{S}$ are exactly the sets $S \in \mathscr{S}$. \end{lemma} \proof{\textbf{Proof}} If $S \in \mathscr{S}$, it is obviously a feasible set, and so we have immediately that $S$ is an independent set in $H^c_\mathscr{S}$. If it is not maximal, then we could add some $v \in J \backslash S$ and maintain feasibility, which would violate our irredundancy assumption (i.e. $S \cup \{v\} \subseteq S' \in \mathscr{S}$, $S\in \mathscr{S}$, and $ S \subsetneq S \cup \{v\} \subseteq S'$). If $S$ is a maximal independent set in $H^c_\mathscr{S}$, then it must be a feasible set with respect to $\mathscr{S}$ as well. As it is maximal, there is no set $\hat{S} \in \mathscr{S}$ with $S \subsetneq \hat{S}$, and so we must have $S \in \mathscr{S}$ as well. \Halmos\endproof \begin{theorem}\label{thm:representability} A $k$-way IB scheme exists for $\operatorname{CDC}(\mathscr{S})$ if and only if $r\bra{H^c_\mathscr{S}} \leq k$. In particular, if $\mathcal{E}_\mathscr{S}=\set{E^j = \{e^j_1,\ldots,e^j_{|E^j|}\}}_{j=1}^t$ is the hyperedge set for the conflict hypergraph $H^c_\mathscr{S}$, then an $r\bra{H^c_\mathscr{S}}$-way IB scheme for $\operatorname{CDC}(\mathscr{S})$ is given by \begin{equation}\label{basicIB} L^j_i = \begin{cases} J \backslash \{e^j_i\} & i \leq \abs{E^j} \\ \emptyset & \text{o.w.} \end{cases} \quad \forall i \in \llbracket r\bra{H^c_\mathscr{S}} \rrbracket, \: j\in \llbracket t \rrbracket. \end{equation} \end{theorem} \proof{\textbf{Proof}} To show the ``if'' direction, it suffices to show the validity of \eqref{basicIB}. First note that every minimally infeasible set $E^j \in \mathcal{E}_\mathscr{S}$ is rendered infeasible by level $j$, which implies that every infeasible set is rendered infeasible as well. Then note that for any $S \in \mathscr{S}$ and for any $j\in \llbracket t \rrbracket$, we have $E^j \not\subseteq S$, so there exists $i \in \llbracket \abs{E^j} \rrbracket $ such that $e^j_i\in E^j \setminus S$. Hence, $S \in L^j_i$ and $S$ is feasible for level $j$. To show the ``only if'' direction, assume for a contradiction that there exists a $k$-way IB scheme with $k\leq r\bra{H^c_\mathscr{S}}-1$. Take a minimal infeasible set $E =\{e_1,\ldots,e_r\} \in \mathcal{E}_\mathscr{S}$, where $r = r(H_\mathscr{S}^c)$. Then take $j \in \llbracket t \rrbracket$ as a level of the IB scheme that renders $E$ infeasible. By the minimality of $E$, we have that, for all $\ell \in \llbracket r \rrbracket$, there exists some $i(\ell) \in \llbracket k \rrbracket$ such that $E(\ell) \defeq E \setminus \set{e_\ell} \subseteq L^j_{i(\ell)}$. As $k < r$, we may apply the pigeonhole principle to see that there must exist some distinct $\ell^1,\ell^2 \in \llbracket r \rrbracket$ such that $i(\ell^1)=i(\ell^2)$, and such that $E(\ell^1) \subseteq L^j_{i(\ell^1)}$ and $E(\ell^2) \subseteq L^j_{i(\ell^2)}$. As $E = E(\ell^1) \cup E(\ell^2)$ and $L^j_{i(\ell^1)} = L^j_{i(\ell^2)}$, this implies that $E \subseteq L^j_{i(\ell^1)}$, which contradicts our supposition that level $j$ rendering $E$ infeasible. \Halmos\endproof Throughout, we will say that $\operatorname{CDC}(\mathscr{S})$ is $k$-way IB-representable (or pairwise IB-representable for $k=2$) if it admits a $k$-way IB scheme. \subsection{Cardinality constraints} \label{ss:cardinality} Our first application of Theorem~\ref{thm:representability} is to derive a strong restriction on the existence of multi-way IB schemes for the cardinality constraint. \begin{corollary} \label{prop:cardinality} A cardinality constraint of degree $\ell$ is $k$-way IB-representable if and only if $k > \ell$. \end{corollary} \proof{\textbf{Proof}} Direct from Theorem~\ref{thm:representability} by observing that the conflict hypergraph has rank $\ell+1$. \Halmos\endproof We observe that the IB scheme \eqref{basicIB}, when applied to the cardinality constraint, is a natural MIP formulation for the ``conjunctive normal form'' \cite{Balas:1985}, and is unlikely to be practical for even moderately large $\ell$. In addition, both specialized constraint branching schemes for cardinality constraints \cite{Ismael-R.-de-Farias:2003} and the binary variable branching induced by standard formulations for cardinality constraints are quite imbalanced. The existence of a pairwise independent branching scheme for cardinality constraints would likely have finally produced the sought-after balanced constraint branching. However, Corollary~\ref{prop:cardinality} implies that such a balanced constraint branching cannot be produced via IB schemes, or equivalently by constraint branchings that do not use general inequalities (i.e. are only multi-variable branchings). \subsection{Polygonal partitions of the plane} \label{ss:partitions-of-the-plane} Consider a (nonconvex) bounded region in the plane $\Omega \subset \mathbb{R}^2$ that describes all possible locations for a UAV, as described in Section~\ref{obstacleavoidancesec}. Assume that $\Omega$ can be partitioned into polyhedra $\{P^i\}_{i=1}^d$ such that $\bigcup_{i=1}^d P^i = \Omega$ and $\operatorname{relint}(P^i) \cap \operatorname{relint}(P^j) = \emptyset$ for each distinct $i,j \in \llbracket d \rrbracket$. We note that this partition will not, in general, be unique, and its selection can have a significant effect on questions of representability or formulation size. Figure~\ref{fig:partitioning} illustrates this for a convex region with a ``hole.'' The figure shows three ways to partition the resulting nonconvex region into convex polyhedra. Once this partition is fixed, we describe the associated polyhedra in $\mathscr{V}$-form, and so the corresponding combinatorial disjunctive constraint is given by $\mathscr{S} = \{\operatorname{ext}(P^i)\}_{i=1}^d$ and $J = \bigcup\{S \in \mathscr{S}\}$. We additionally forbid polyhedra with ``internal vertices'' by requiring that \begin{equation}\label{intvercon} v \in P^i \Longleftrightarrow v \in \operatorname{ext}(P^i) \quad \forall i \in \llbracket d \rrbracket, v \in J, \end{equation} so that $\mathscr{S}$ corresponds to the maximal elements of a polyhedral complex \cite[Section 5.1]{Ziegler:2007}. For example, the second and third partitions in Figure~\ref{fig:partitioning} satisfy this condition, while the first does not. In this setting, minimal infeasible sets have a natural characterization. \begin{theorem} \label{thm:partitioning} Take bounded $\Omega \subset \mathbb{R}^2$ and a polyhedral partition $\{P^i\}_{i=1}^d$ of $\Omega$ satisfying the internal vertex condition \eqref{intvercon}. If $\mathscr{S} = \{\operatorname{ext}(P^i)\}_{i=1}^d$, then $r\bra{H^c_\mathscr{S}}\leq 3$. \end{theorem} \proof{\textbf{Proof}} Take some minimal infeasible hyperedge $E \in \mathcal{E}_\mathscr{S}$ of $H^c_\mathscr{S}$, assuming for contradiction that $r = |E| > 3$, and label the points $E = \{v^i\}_{i=1}^r$. First, we show that the points may not be in general position, i.e. that w.l.o.g. $v^r \in \operatorname{Conv}(\{v^i\}_{i=1}^{r-1})$. Then, we argue that the points not being in general position implies that $\{v^i\}_{i=1}^{r-1}$ is also an infeasible set, violating the minimality condition. Assume for contradiction that the points are in general position; that is, that none can be written as a convex combination of the others. This implies that $\operatorname{ext}(\operatorname{Conv}(E)) = E$. Assume that the ordering $\{v^1,\ldots,v^r\}$ forms a path around the edges of $\operatorname{Conv}(E)$; that is, $v^i$ and $v^j$ both lie on an edge of $\operatorname{Conv}(E)$ if and only if $|i-j|=1$ or $\{i,j\}=\{1,r\}$. Choose some set $S^1 \in \mathscr{S}$ and some $2 < j < r$ such that $v^1,v^j \in S^1$ and $v^2 \not\in S^1$; the associated polyhedron is $P^1$. Such a set exists, else $E$ is not a minimal infeasible set (choose instead $E \backslash \{v^2\}$). Now choose $S^2 \in \mathscr{S}$ such that $v^2, v^r \in S^2$; the associated polyhedron is $P^2$. Such as set exists, as $\{v^2,v^r\} \subsetneq E$ and $E$ is minimal. As the nodes $v^1,v^2,v^j,v^r$ are interlaced along the boundary of $\operatorname{Conv}(E)$, we have that $\operatorname{Conv}(\{v^1,v^j\}) \cap \operatorname{Conv}(\{v^2,v^r\}) \subset \operatorname{Conv}(E)$ is nonempty. As each of the four points is on the boundary of $\operatorname{Conv}(E)$, and the points are in general position, it follows that $\operatorname{Conv}(\{v^1,v^j\}) \cap \operatorname{Conv}(\{v^2,v^r\}) = \operatorname{relint}(\operatorname{Conv}(\{v^1,v^j\})) \cap \operatorname{relint}(\operatorname{Conv}(\{v^2,v^r\}))$. Therefore, there must exist some point $y$ with $y \in \operatorname{relint}(\operatorname{Conv}(\{v^1,v^j\})) \subseteq \operatorname{relint}(P^1)$ and $y \in \operatorname{relint}(\operatorname{Conv}(\{v^2,v^r\})) \subseteq \operatorname{relint}(P^2)$. However, this implies that $\operatorname{relint}(P^1) \cap \operatorname{relint}(P^2) \neq \emptyset$, which contradicts the assumption that our sets partition the region $\Omega$. Finally, it just remains to show that $\{v^i\}_{i=1}^{r-1}$ is also an infeasible set, and therefore $\{v^i\}_{i=1}^r$ cannot be a minimal infeasible set. Assume for contradiction that it is not: i.e. that there exists some $j$ such that $\{v^i\}_{i=1}^{r-1} \subseteq \operatorname{ext}(P^j)$. But this implies that $v^r \in J$ and $v^r \in \operatorname{Conv}(E) \subseteq P^j$, yet $v^r \not\in \operatorname{ext}(P^j)$, a contradiction of the internal vertices assumption. \Halmos\endproof In other words, every polyhedral partition of the plane is $3$-way independent branching-representable, and pairwise IB representability can be checked in polynomial time (for example, by enumerating the subsets of $J$ of cardinality 3). To illustrate, in Figure~\ref{fig:partitioning} we depict the three possible cases for a partition with respect to Theorem~\ref{thm:partitioning}: 1) it does not satisfy the internal vertices condition, 2) it admits a pairwise IB scheme ($r(H^c_\mathscr{S})=2$), or 3) it does not admit a pairwise IB scheme, but does admit a 3-way IB scheme ($r(H^c_\mathscr{S})=3$). \begin{figure}[htpb] \centering \includegraphics[width=.32\linewidth]{Fig2a} \includegraphics[width=.32\linewidth]{Fig2b} \includegraphics[width=.32\linewidth]{Fig2c} \caption{Partitions of a nonconvex region in the plane obtained by removing a central non-convex portion from a convex polyhedron. The first partition does not satisfy the internal vertices condition \eqref{intvercon} (Left), the second partition admits a pairwise IB scheme (Center), and the third partition admits a 3-way IB scheme but not a pairwise one (Right).} \label{fig:partitioning} \end{figure} Furthermore, we can argue that we can always represent a obstacle avoidance constraint in such a way that it admits a pairwise IB scheme. Inspecting Figure~\ref{fig:partitioning}, we see that the region $\Omega$ is the same in each, and it is only the partition of $\Omega$ that can potentially lead to constraints that are not pairwise IB-representable. Therefore, the obstacle avoidance constraint is invariant to the specification of the partition, and if any polyhedral partitioning exists, then it is always possible to construct one that satisfies the conditions of Theorem~\ref{thm:partitioning}.\footnote{Note that this result does not carry over to piecewise linear functions over $\Omega$, as the choice of the partition is intimately connected with the values the function may take.} \subsection{SOS2} \label{ss:SOS2} Our first example of a constraint that is always pairwise IB-representable is the SOS$2(N)$ constraint. Recall that $J = \llbracket N \rrbracket$ and $\mathscr{S} = \{\{\tau,\tau+1\}: \tau \in \llbracket N-1 \rrbracket\}$ for SOS2. Then $\mathcal{E}_\mathscr{S}=\set{\{\tau,\tau+t\}: \tau,\tau+t \in \llbracket N \rrbracket, \: t \geq 2}$, $r\bra{H^c_\mathscr{S}}=2$, and formulation \eqref{basicIB} has depth $t=\binom{N}{2}-N+1$. However, Vielma and Nemhauser \cite{Vielma:2009a} construct a pairwise IB scheme for SOS$2(N)$ constraints of depth logarithmic in $N$. The construction is built around a Gray code \cite{Savage:1997}, or sequence of distinct binary vectors $\{h^i\}_{i=1}^{N-1} \subseteq \{0,1\}^{\lceil \log_2(N-1) \rceil}$ where each adjacent pair $(h^i,h^{i+1})$ differs in exactly one component. Notationally, here and throughout, take $h^0 \defeq h^1$ and $h^N \defeq h^{N-1}$. The pairwise IB scheme is then given by \[ L^j = \left\{\tau \in \llbracket N \rrbracket : h^{\tau-1}_j = 1 \text{ or } h^{\tau}_j = 1 \right\}, \quad R^j = \left\{\tau \in \llbracket N \rrbracket : h^{\tau-1}_j = 0 \text{ or } h^{\tau}_j = 0 \right\} \quad\quad \forall j \in \llbracket \lceil \log_2(N-1) \rceil \rrbracket. \] We observe that the resulting formulation matches the lower bound from Proposition~\ref{prop:log-bound} with respect to the number of auxiliary binary variables and is significantly smaller than formulation \eqref{basicIB}. Indeed, formulation \eqref{basicIB} is likely to be unnecessarily large for the other constraints we consider as well, so we turn our attention to finding smaller IB schemes in Section~\ref{PISEC}. We end this section by applying Theorem~\ref{thm:representability} to succinctly prove the pairwise IB-representability of two other constraints. \subsection{Other pairwise IB-representable constraints} \paragraph{SOS$k$} In this case $\mathcal{E}_\mathscr{S}=\set{\{\tau,\tau+t\}: \tau,\tau+t \in \llbracket N \rrbracket, \: t\geq k+1}$ and $r\bra{H^c_\mathscr{S}}=2$. \paragraph{Grid triangulations} We show that $r\bra{H^c_\mathscr{S}}=2$ by seeing that for any infeasible set $T \subseteq J$ there exist some distinct $v,w\in T$ such that $\set{v,w}$ is infeasible. Indeed, if there are some $v,w \in T$ such that $||v - w||_\infty > 1$, then there does not exist any triangle on the grid that contains both, so $\{v,w\}$ is also an infeasible set. Otherwise, we have that $T \subset \{r,r+1\}\times\{s,s+1\}$ for some $r,s$, and that $T$ contains elements in both of the triangles in this square. For each of the two triangles, we can select an element of $T$ that is not contained in the other triangle, which yields an infeasible pair contained in $T$. Therefore, any grid triangulation is pairwise IB-representable. \paragraph{Discretization of multilinear terms} Similarly as for the grid triangulation case, each infeasible set $T \subseteq J$ must necessarily contain two elements $v,w \in T$ with $||v - w||_\infty > 1$, and so we have that $T$ can be reduced to the infeasible pair $\{v,w\}$. Therefore, $r(H^c_\mathscr{S})=2$, and so any discretization of this form is pairwise IB-representable. \section{Pairwise independent branching schemes}\label{PISEC} The pairwise independent branching scheme framework was initially introduced by Vielma and Nemhauser \cite{Vielma:2009a}, where it was used to model particularly structured piecewise linear functions. In the remainder of this work, we will focus on pairwise IB schemes and offer a complete picture of their expressive powers, along with an algorithmic framework for constructing them. \subsection{Graphical representations of pairwise IB-representable CDCs} \label{ss:graph-representation} From our covering assumption $J = \bigcup\set{S\in \mathscr{S}}$, we can see that $\abs{E}\geq 2$ for each $E\in\mathcal{E}_\mathscr{S}$. By applying Theorem~\ref{thm:representability}, we then immediately have that $\operatorname{CDC}(\mathscr{S})$ is pairwise IB-representable if and only if $H^c_\mathscr{S}$ is (equivalent to) a graph. Along this line, for any constraint we may define a \emph{conflict graph} for any $\operatorname{CDC}(\mathscr{S})$ as $G^c_\mathscr{S} \defeq (J,\bar{E})$, where $\bar{E} = \bar{E}_\mathscr{S} \defeq \{\set{u,v} \in [J]^2 : u \neq v, \: \{u,v\} \text{ is an infeasible set}\}$ is the set of all infeasible pairs of elements of $J$. Checking for pairwise IB-representability then reduces to verifying if $\mathcal{E}_\mathscr{S}=\bar{E}_\mathscr{S} $. The following corollary of Theorem~\ref{thm:representability} shows that this can also be verified by only working with $G^c_\mathscr{S}$. \begin{corollary} \label{thm:cliques} $\operatorname{CDC}(\mathscr{S})$ is pairwise IB-representable if and only if the sets $\mathscr{S}$ are exactly the maximal independent sets of $G_\mathscr{S}^c$. \end{corollary} \proof{\textbf{Proof}} If $\operatorname{CDC}(\mathscr{S})$ is pairwise IB-representable, then $G_\mathscr{S}^c$ is equivalent to $H^c_\mathscr{S}$. By applying Theorem~\ref{thm:representability}, the maximal independent sets of $G_\mathscr{S}^c$ are exactly the elements of $\mathscr{S}$. For the converse, assume for a contradiction that $\mathscr{S}$ is exactly the maximal independent sets of $G_\mathscr{S}^c$, but that there exists some $E\in\mathcal{E}_\mathscr{S}$ with $\abs{E}\geq 3$. By the minimal infeasibility of $E$, we have that $\set{r,s}\notin \bar{E}_\mathscr{S}$ for any distinct $r,s\in E$, and therefore $E$ is an independent set in $G_\mathscr{S}^c$. This implies that $E$ is contained in a maximal independent $S$. By assumption, $S\in \mathscr{S}$, which contradicts the infeasibility of $E$. \Halmos\endproof Therefore, verifying general pairwise IB-representability reduces to enumerating the maximal independent sets of $G_\mathscr{S}^c$ and identifying them to exactly the sets $\mathscr{S}$. As an example, we can see that, for cardinality constraint of degree $\ell$ with $2 \leq \ell < |J|$, the only maximal independent set of $G_\mathscr{S}^c$ is the entire ground set $J$, which certainly cannot be identified with $\mathscr{S} = \{S \subset J : |S| = \ell \}$. \subsection{Representation at a given depth} Once a CDC has been shown to be pairwise IB-representable, a natural next question is: what is the smallest possible depth at which we may construct an IB scheme? That is, we ask if there exists a pairwise IB scheme for $\operatorname{CDC}(\mathscr{S})$ of some given depth $t$. The answer to this question reduces to the existence of a graphical decomposition of the conflict graph $G_\mathscr{S}^c$. \begin{definition} A \emph{biclique cover} of the graph $G = (J,E)$ is a collection of complete bipartite subgraphs $\left\{G^j\defeq (A^j\cup B^j,E^j)\right\}_{j=1}^t$ of $G$ that cover all the edges of $G$. Formally, this means that for each $j \in \llbracket t \rrbracket$, $\emptyset \subsetneq A^j,B^j \subsetneq J$, $A^j \cap B^j = \emptyset$, and $E^j =A^j * B^j\defeq \set{\set{a,b}\,:\, a\in A^j,\: b\in B^j}$, and that $\bigcup_{j=1}^t E^j = E$. For notational simplicity, we will often refer to the sets $\{(A^j,B^j)\}_{j=1}^t$ as a biclique cover, as we can recover the graphs $G^j$ directly. \end{definition} The following theorem formalizes the equivalence between biclique covers and pairwise IB schemes. \begin{theorem} \label{thm:biclique-IBSneq} If $\{(A^j,B^j)\}_{j=1}^t$ is biclique cover of the conflict graph $G_\mathscr{S}^c$ for pairwise IB-representable $\operatorname{CDC}(\mathscr{S})$, then a pairwise IB scheme for $\operatorname{CDC}(\mathscr{S})$ is given by \begin{equation} \label{eqn:biclique-to-IBnew} L^j = J \backslash A^j, \quad R^j = J \backslash B^j \quad\quad \forall j \in \llbracket t \rrbracket. \end{equation} Conversely, if $\{(L^j,R^j)\}_{j=1}^t$ is a pairwise IB scheme for $\operatorname{CDC}(\mathscr{S})$, then a biclique cover of the conflict graph $G_\mathscr{S}^c$ is given by \begin{equation} \label{eqn:biclique-to-IBnewconv} A^j = J \backslash L^j, \quad B^j = J \backslash R^j \quad\quad \forall j \in \llbracket t \rrbracket. \end{equation} \end{theorem} \proof{\textbf{Proof}} For the first part, take $\bar{E}$ as the edge set of $G_\mathscr{S}^c$. To see that any $S \in \mathscr{S}$ is feasible for the IB scheme \eqref{eqn:biclique-to-IBnew}, note that if $S\not\subseteq L^j$ and $S\not\subseteq R^j$, then there exist some $u\in A^j \cap S$ and $v\in B^j \cap S$. However, this implies that $\set{u,v}\in A^j * B^j \subseteq \bar{E}$, which is a contradiction of feasibility as $\set{u,v}\subseteq S$ and $S \in \mathscr{S}$. Furthermore, as $\{(A^j,B^j)\}_{j=1}^t$ is a biclique cover of $G^c_\mathscr{S}$, for every $\set{u,v}\in \bar{E}$ we have that there exists some level $j\in \sidx{t}$ such that w.l.o.g. $u\in A^j$ and $v\in B^j$. This implies that $u\notin L^j$ and $v\notin R^j$ by their construction, and as $\operatorname{CDC}(\mathscr{S})$ is pairwise IB-representable, then any infeasible set for $\operatorname{CDC}(\mathscr{S})$ is also infeasible for the proposed IB scheme. Therefore, \eqref{eqn:biclique-to-IBnew} is a valid pairwise IB scheme. For the second part, note that $A^j \cap B^j = \emptyset$ for all $j \in \llbracket t \rrbracket$, and that the covering portion of Assumption~\ref{basicassumption} implies that $L^j\cup R^j=J$. Therefore, it only remains to show that $\bar{E} = \bigcup_{j=1}^t \bar{E}^j$, where $\bar{E}^j = A^j * B^j$. For that, first note that as $L^j\cup R^j=J$, we have that $A^j=R^j\setminus L^j$ and $B^j=L^j\setminus R^j$. The containment $\bar{E} \subseteq \bigcup_{j=1}^t \bar{E}^j$ then follows by noting that, as $\{(L^j,R^j)\}_{j=1}^t$ is a valid pairwise IB scheme, each minimal infeasible set $\{u,v\}\in \bar{E}$ has some level $j \in \llbracket t \rrbracket $ such that $\{u,v\}\not\subseteq L^j$ and $\{u,v\}\not\subseteq R^j$. Then, as $L^j\cup R^j=J$, we have (w.l.o.g.) that $u \in L^j \backslash R^j \equiv B^j$ and $b \in R^j \backslash L^j \equiv A^j$, and so $\{a,b\} \in \bar{E}^j$. For the reverse containment $\bigcup_{j=1}^t \bar{E}^j \subseteq \bar{E}$, take some arbitrary $j \in \llbracket t \rrbracket$ and some edge $\{a,b\} \in \bar{E}^j$. From the definition of our biclique cover, we have that w.l.o.g. $a \in A^j \equiv R^j \backslash L^j$ and $b \in B^j \equiv L^j \backslash R^j$. Therefore, $\{a,b\}$ is an infeasible set for the IB scheme, and thus for $\operatorname{CDC}(\mathscr{S})$ as well, and so $\{a,b\} \in \bar{E}$. \Halmos\endproof We can now naturally frame the problem of finding a minimum depth pairwise IB scheme as the minimum biclique cover problem \cite{Fishburn:1996,Garey:1979}. Unfortunately, the decision version of this problem is known to be NP-complete \cite{Orlin:1977} and inapproximable within a factor of $|J|^{1/3-\epsilon}$ if $P \neq NP$~\cite{Gruber:2007}, even for bipartite graphs. However, we note that it is simple to construct a MIP feasibility problem for finding a pairwise IB scheme of a given depth $t$, which gives us a way to algorithmically find the smallest pairwise IB scheme for a specific (fixed) CDC. We present such a formulation in Proposition~\ref{prop:feasibility-IP} in Appendix~\ref{app:feasibility-IP}. Additionally, Cornaz and Fonlupt \citep{Cornaz:2006} present a MIP formulation (with an exponential number of constraints that can be efficiently separated) to find the minimum level biclique cover of a graph. Furthermore, we can restate the MIP formulation from \cite{Vielma:2009a} (which is a special case of \eqref{eqn:multiway-formulation} with $k=2$) in terms of biclique covers of $G_\mathscr{S}^c$. \begin{proposition}[Theorem 5, \cite{Vielma:2009a}; Theorem 1, \cite{Vielma:2010}] \label{prop:ideal-formulation-for-IBS} If $\operatorname{CDC}(\mathscr{S})$ is pairwise independent branching-representable and $\{(A^j,B^j)\}_{j=1}^t$ is a biclique cover for $G_\mathscr{S}^c$, then the following is an ideal formulation for $\operatorname{CDC}(\mathscr{S})$: \begin{subequations} \label{eqn:ideal-formulation-for-IBS} \begin{alignat}{2} \sum_{v \in A^j} \lambda_v &\leq z_j &\quad \forall j \in \llbracket t \rrbracket \label{eqn:ideal-formulation-for-IBS-1} \\ \sum_{v \in B^j} \lambda_v &\leq 1 - z_j &\quad \forall j \in \llbracket t \rrbracket \label{eqn:ideal-formulation-for-IBS-2} \\ (\lambda,z) &\in \Delta^J \times \{0,1\}^t. \label{eqn:ideal-formulation-for-IBS-3} \end{alignat} \end{subequations} \end{proposition} We end the section by noting that the relation between biclique covers and independent sets has also been exploited in the study of boolean functions, particularly in the equivalence between posiforms and maximum weighted stable sets (e.g. \cite[Theorem 13.16]{crama2011boolean}). In fact, formulation \eqref{eqn:ideal-formulation-for-IBS} is reminiscent of formulation (13.45--13.50) in \cite[Theorem 13.13]{crama2011boolean}. The main difference between these formulations is that in the context of \cite{crama2011boolean} the $\lambda$ variables will be binary variables not constrained to lie in the unit simplex. For this reason inequalities (\ref{eqn:ideal-formulation-for-IBS-1}--\ref{eqn:ideal-formulation-for-IBS-2}) appear disaggregated in \cite[Theorem 13.13]{crama2011boolean} in the form $\lambda_v \leq z_j$ for all $v\in A^j$, $j \in \llbracket t \rrbracket$. However, the resulting formulation is not ideal (See \cite[Section 5]{Vielma:2009a} for more details). Still, the combinatorial aspects of this connection could prove useful for constructing small IB schemes. In the next section, we will explore instances where we can, in closed form, construct small (asymptotically optimal) IB schemes for families of particularly structured CDCs. \section{Illustrative examples} \label{s:examples} With a framework to construct pairwise independent branching schemes for arbitrary pairwise IB-representable CDCs, we now return to some of our motivating examples. We will apply our methodology to these specific structures, and produce small, closed-form IB schemes. In particular, this allows us to construct novel, small MIP formulations for these constraints. \subsection{A simple IB scheme and its limitations} To start, we show that \emph{any} pairwise IB-representable CDC admits an IB scheme of depth $|J|$. If $|J|$ is smaller than $|\mathscr{S}|$, this already offers a drop in size from \eqref{eqn:disaggregated}. This IB scheme covers all edges incident to node with the simple biclique corresponding to the star centered at that node. \begin{proposition}[Covering with Stars] \label{prop:simple-pIBS} For pairwise IB-representable $\operatorname{CDC}(\mathscr{S})$, a biclique cover for $G_\mathscr{S}^c$ is given by: \begin{subequations} \label{eqn:simple-pIBS} \[ A^v = \{v\},\quad\quad B^v = \left\{u \in J : \set{u,v} \in \bar{E}\right\} \quad\quad \forall v \in J. \] \end{subequations} \end{proposition} \proof{\textbf{Proof}} By construction of the sets, we see that each $\set{r,s} \in \bar{E}^v \equiv A^v * B^v$ corresponds to an infeasible edge: that is, $\bar{E}^v \subseteq \bar{E}$ for each $v$, and so $\bigcup_{v \in J} \bar{E}^v \subseteq \bar{E}$. Furthermore, each infeasible edge $\set{r,s} \in \bar{E}$ is infeasible for levels $r$ and $s$, and so $\bar{E} \subseteq \bigcup_{v \in J} \bar{E}^v$. Therefore, this construction forms a valid biclique cover of the conflict graph. \Halmos\endproof This gives us an upper bound of $|J|$ on the minimum depth for any pairwise IB-representable CDC. However, if we exploit the specific structure of a CDC, we can typically get much smaller formulations. For instance, consider the following two instances of the SOS$3(N)$ constraint for small values of $N$. First, consider the instance with $N=6$, where $\abs{J}=6$ and $\mathscr{S} = \{\{1,2,3\},\{2,3,4\},\{3,4,5\},\{4,5,6\}\}$. Therefore, $|\mathscr{S}|=4$, yielding a lower bound of depth $\log_2(4)=2$ from Proposition~\ref{prop:log-bound}. However, there does not exist a biclique cover of depth 2 (which can be verified via Proposition \ref{prop:feasibility-IP}), though one of depth 3 does exist: \begin{alignat*}{3} A^1 &= \{1\}, \quad &B^1 &= \{4,5,6\} \\ A^2 &= \{1,2\}, \quad &B^2 &= \{5,6\} \\ A^3 &= \{1,2,3\}, \quad &B^3 &= \{6\}. \end{alignat*} We can see the proposed IB scheme on the left side of Figure \ref{fig:sos3}. For clarity, the associated MIP formulation for the CDC from Proposition~\ref{prop:ideal-formulation-for-IBS} is \begin{alignat*}{2} \lambda_1 &\leq z_1 \quad\quad \lambda_4 + \lambda_5 + \lambda_6 &\leq 1 - z_1 \\ \lambda_1 + \lambda_2 &\leq z_2 \quad\quad \lambda_5 + \lambda_6 &\leq 1 - z_2 \\ \lambda_1 + \lambda_2 + \lambda_3 &\leq z_3 \quad\quad \lambda_6 &\leq 1 - z_3 \\ (\lambda,z) &\in \Delta^6 \times \{0,1\}^3. \end{alignat*} Next, we consider $N=10$, where we also cannot attain the $\log_2(8)=3$ lower bound. However, a biclique for this the conflict graph of this constraint is \begin{subequations} \label{eqn:sos3-6-biclique} \begin{alignat}{3} A^1 &= \{1,8,9,10\}, \quad &B^1 &= \{4,5\} \\ A^2 &= \{1,2,10\}, \quad &B^2 &= \{5,6,7\} \\ A^3 &= \{1,2,3,9,10\}, \quad &B^3 &= \{6\} \\ A^4 &= \{1,2,3,4\}, \quad &B^4 &= \{7,8,9,10\}, \end{alignat} \end{subequations} as seen on the right side of Figure \ref{fig:sos3}. The corresponding MIP formulation is \begin{alignat*}{2} \lambda_1 + \lambda_8 + \lambda_9 + \lambda_{10} &\leq z_1 \quad\quad \lambda_4 + \lambda_5 &\leq 1 - z_1 \\ \lambda_1 + \lambda_2 + \lambda_{10} &\leq z_2 \quad\quad \lambda_5 + \lambda_6 + \lambda_7 &\leq 1 - z_2 \\ \lambda_1 + \lambda_2 + \lambda_3 + \lambda_9 + \lambda_{10} &\leq z_3 \quad\quad \lambda_6 &\leq 1 - z_3 \\ \lambda_1 + \lambda_2 + \lambda_3 + \lambda_4 &\leq z_4 \quad\quad \lambda_7 + \lambda_8 + \lambda_9 + \lambda_{10} &\leq 1 - z_4 \\ (\lambda,z) &\in \Delta^{10} \times \{0,1\}^4. \end{alignat*} \begin{figure}[htpb] \centering \includegraphics[width=.37\linewidth]{Fig3a} \hspace{2em} \includegraphics[width=.55\linewidth]{Fig3b} \\ \includegraphics[width=.37\linewidth]{Fig3c} \hspace{2em} \includegraphics[width=.55\linewidth]{Fig3d} \\ \includegraphics[width=.37\linewidth]{Fig3e} \hspace{2em} \includegraphics[width=.55\linewidth]{Fig3f} \\ \includegraphics[width=.37\linewidth]{Fig3g} \hspace{2em} \includegraphics[width=.55\linewidth]{Fig3h} \caption{Visualizations of the biclique covers presented in the text for SOS$3(6)$ (Left) and SOS$3(10)$ (Right). Each row corresponds to some level $j$, and the elements of $A^j$ and $B^j$ are the squares and diamonds, respectively.} \label{fig:sos3} \end{figure} The ad-hoc construction for SOS$3(6)$ suggests a more general construction for SOS$k(N)$ when $k \leq N/2$ (assume for convenience that $N$ is even). Consider the sets given by \[ A^j = \{1,\ldots,j\} \cup \{j+N/2+k,\ldots,N\}, \quad\quad\quad B^j = \{j+k,\ldots,j+N/2\} \] for each $j \in \llbracket N/2 \rrbracket$. It is straightforward to see that this yields a biclique cover of the conflict graph for SOS$k(N)$ of depth $N/2$. Therefore, with this simple operation, we have constructed an ideal formulation for SOS$k(N)$ with size strictly smaller than $N$, the size of the na\"ive non-extended formulation \eqref{eqn:disaggregated}. Based on the second example~\eqref{eqn:sos3-6-biclique}, we know that this construction is, in general, not the smallest possible. In Section~\ref{ss:sosk}, we will see how we can systematically construct small biclique covers (and MIP formulations) for SOS$k$($N$) with arbitrary $k$ and $N$, using techniques we will now develop. \subsection{Systematic construction of biclique covers} As discussed in Section~\ref{ss:SOS2}, there exists an IB scheme for the SOS2 constraint of optimal depth that can be constructed using a Gray code. The following proposition shows how the validity of this scheme can easily be proven by reinterpreting it via a biclique cover. \begin{proposition} \label{prop:sos2-IB-scheme} Take a Gray code $\{h^i\}_{i=1}^{N-1} \subseteq \{0,1\}^{\lceil \log_2(N-1) \rceil}$ and let $h^0 \defeq h^1$ and $h^N \defeq h^{N-1}$. If $G_\mathscr{S}^c$ is the conflict graph of SOS2($N$), then a biclique cover for $G_\mathscr{S}^c$ of depth $\lceil \log_2(N-1) \rceil$ is given by \begin{equation}\label{SOS2biclique} A^{j} = \left\{\tau \in \llbracket N \rrbracket : h^{\tau-1}_j = h^{\tau}_j = 0 \right\},\quad B^{j} = \left\{\tau \in \llbracket N \rrbracket : h^{\tau-1}_j = h^{\tau}_j = 1 \right\} \quad \forall j \in \llbracket \lceil \log_2(N-1) \rceil \rrbracket. \end{equation} \end{proposition} \proof{\textbf{Proof}} For the SOS2($N$) constraint we have that $\bar{E}_\mathscr{S}=\set{\{r,s\} \in \llbracket N \rrbracket^2 : r+2 \leq s}$. Take any infeasible pair $\{r,s\}\in \bar{E}_\mathscr{S}$. As $r + 2 \leq s$, we conclude that $r-1<r<s-1<s$, and so it must be that $h^{r-1}, h^r \neq h^{s-1},h^s$. The set of components which flip values between the two pairs of adjacent codes $(h^{r-1},h^r)$ and $(h^{s-1},h^s)$ is $I = \{ j \in \llbracket \lceil \log_2(N-1) \rceil \rrbracket : h^{r-1}_j \neq h^r_j \text{ or } h^{s-1}_j \neq h^s_j\}$, and $|I| \leq 2$ as we have selected a Gray code. Now it must be the case that there is some component $j \in \llbracket \lceil \log_2(N-1) \rceil \rrbracket \backslash I$ wherein $h^{r-1}_j=h^r_j \neq h^{s-1}_j=h^s_j$, else we conclude that two of the vectors $h^i=h^{\ell}$ coincide for some $i \in \{r-1,r\}$ and $\ell \in \{s-1,s\}$, a contradiction of their uniqueness. Then $\{r,s\} \in E^j$, i.e. it is covered by the $j$-th level of the biclique. Furthermore, we observe that no edges of the form $\{r,r+1\}$ will be contained in the biclique cover, as it is not possible that $h_j^{r-1} = h_j^r = 0$ (resp. $=1$) and $h_j^r = h_j^{r+1} = 1$ (resp. $=0$) simultaneously. \Halmos\endproof \begin{figure}[htpb] \centering \includegraphics[width=.57\linewidth]{Fig4.pdf} \caption{The recursive construction for biclique covers for SOS2. The first row is a single biclique that covers the conflict graph for SOS2(3) ($A^{1,1}$ are squares, $B^{1,1}$ are diamonds). The second row shows the construction which duplicates the ground set $\{1,2,3\}$ and inverts the ordering on the second copy. The third row shows the identification of the nodes that yields a valid biclique for SOS2(5). This biclique is then combined with a second that covers all edges between nodes identified with the first copy and those identified with the second, giving a biclique cover for SOS2(5) with two levels.} \label{fig:sos2-recursive} \end{figure} Interestingly, we can also view this construction recursively if we use a specific Gray code known as the binary reflected Gray code~\cite{Savage:1997}. For SOS2($2^k$), we will take $E^k$ as the edge set for the corresponding conflict graph. First, with $k=1$, $d=2^k=2$, $N=2^{k}+1=3$, then $E^1 = \{\{1,3\}\}$. A complete biclique cover is given by the single biclique $A^{1,1}=\{1\}$ and $B^{1,1} = \{3\}$. As we see in Figure~\ref{fig:sos2-recursive}, we can construct a biclique cover for SOS2 with $k=2$ by stitching together two copies of the biclique. We construct two copies of the node set for $k=1$, invert the second, and identify the last node from the first set with the first node with the second set. Then we can readily construct a mapping of the biclique $(A^{1,1},B^{1,1})$ for $k=1$ to a biclique for $k=2$, using the node identification, as $A^{2,1} = \{1,5\}$ and $B^{2,1} = \{3\}$. This will cover all edges in $E^2$ with both incident nodes in the first half of the nodes, or both in the second half of the nodes (along with some other edges in $E^2$, as well). To cover all edges with one adjacent node in the first half, and the other in the second half, we construct a second biclique of the form $A^{2,2} = \{1,2\}$ and $B^{2,2} = \{4,5\}$. We can repeat this construction with $k=3$ to get the three level biclique cover \begin{alignat*}{3} A^{3,1} &= \{1,5,9\}, \quad\quad &B^{3,1} &= \{3,7\} \\ A^{3,2} &= \{1,2,8,9\}, \quad\quad &B^{3,2} &= \{4,5,6\} \\ A^{3,3} &= \{1,2,3,4\}, \quad\quad &B^{3,2} &= \{6,7,8,9\}. \end{alignat*} Iterating this construction gives a biclique cover for $E^{k+1}$ as $\{(A^{k+1,i},B^{k+1,i})\}_{i=1}^{k+1}$, where \begin{alignat*}{3} A^{k+1,i} &= \bigcup_{u \in A^{k,i}} \left\{u, 2^{k+1}+2-u\right\}, \quad\quad &B^{k+1,i} =& \bigcup_{v \in B^{k,i}} \left\{v,2^{k+1}+2-v\right\} \quad \forall i \in \llbracket k \rrbracket \\ A^{k+1,k+1} &= \left\{1,\ldots,2^k\right\}, \quad\quad &B^{k+1,k+1} =& \left\{ 2^k + 2, \ldots, 2^{k+1}+1 \right\}. \end{alignat*} In fact, we can readily state this recursive construction in a more general form, where we adapt a biclique cover for one graph into a biclique cover for another graph that is created in some specific way. \begin{lemma}\label{lem:recursive-cover} Take some graph $G=(\llbracket m+1 \rrbracket,E)$, and define $G^2 = (\llbracket 2m+1 \rrbracket, E^2)$ for \[E^2 =E \cup \bigl\{\{2m+2-u,2m+2-v\}\,:\, \set{u,v}\in E\bigr\} \cup\bigl(\sidx{m}*\sidx{m+2,2m+1}\bigr) \] where $\sidx{a,b}\defeq \set{a,\ldots,b}$. If $\{(A^j,B^j)\}_{j=1}^t$ is a biclique cover of $G$, then $\{(\tilde{A}^j,\tilde{B}^j)\}_{j=1}^{t+1}$ is a biclique cover of $G^2$, where \begin{alignat*}{3} \tilde{A}^j &= \bigcup_{u \in A^j} \{u, 2m+2-u\}, \quad\quad &\tilde{B}^j =& \bigcup_{v \in B^j} \{v, 2m+2-v\} \quad \forall j \in \llbracket t \rrbracket \\ \tilde{A}^{t+1} &= \{1,\ldots,m\}, \quad\quad &\tilde{B}^{t+1} =& \{m+2,\ldots,2m+1\}. \end{alignat*} \end{lemma} In the remainder of this work, we will see how we may apply similar graphical results to systematically construct small biclique covers for the conflict graphs of constraints by exploiting their specific structure. \subsection{Biclique covers for graph products and discretizations of multilinear terms} Consider the discretization of multilinear terms described in Section~\ref{multilinearsec}, given by $J=\prod_{i=1}^\eta \llbracket d_i \rrbracket$ and $\mathscr{S}=\set{\prod_{i=1}^\eta \set{k_i,k_i+1} : k \in \prod_{i=1}^\eta \llbracket d_i -1 \rrbracket}$. We can interpret this constraint as a $\eta$-dimensional version of the SOS2 constraint, or as the Cartesian product of $\eta$ SOS2 constraints. This can be formalized through the following definition and straightforward lemma. \begin{definition} The \emph{(disjunctive) graph product} of a family of graphs $\{G^i = (J^i,E^i)\}_{i=1}^\eta$ is $\bigvee_{i=1}^\eta G^i \defeq (J_P,E_P)$, where $J_P = \prod_{i=1}^\eta J^i$ and \[ E_P = \left\{ \{u,v\} \in J_P * J_P : \exists i \in \llbracket \eta \rrbracket \text{ s.t. } \{u_i,v_i\} \in E^i \right\}. \] \end{definition} \begin{lemma} \label{lemma:graph-productmulti} Let $J=\prod_{i=1}^\eta \llbracket d_i \rrbracket$ and $\mathscr{S}=\set{\prod_{i=1}^\eta \set{k_i,k_i+1} : k \in \prod_{i=1}^\eta \llbracket d_i -1 \rrbracket}$ be a $\eta$-dimensional discretization of multilinear terms, and $G_\mathscr{S}^c$ be the corresponding conflict graph. If $G^i$ is the conflict graph of SOS2($d_i$) for each $i\in \llbracket \eta \rrbracket$, then $G_\mathscr{S}^c= \bigvee_{i=1}^\eta G^i$. \end{lemma} Using this characterization, we can easily construct an IB scheme for discretizations of multilinear terms by taking the graph products of IB schemes for the SOS2 constraint. \begin{lemma} \label{lemma:graph-product} Take a family of graphs $\{G^i = (J^i,E^i)\}_{i=1}^\eta$, and a biclique cover $\{(\tilde{A}^{i,j},\tilde{B}^{i,j})\}_{j=1}^{t_i}$ for each $G^i$. Then a biclique cover for $\bigvee_{i=1}^\eta G^i$ is given by $\bigcup_{i=1}^\eta \{(A^{i,j},B^{i,j})\}_{j=1}^{t_i}$, where \[ A^{i,j} = \left(\prod_{\ell=1}^{i-1} J^\ell\right) \times \tilde{A}^{i,j} \times \left(\prod_{\ell=i+1}^\eta J^\ell \right), \quad\quad B^{i,j} = \left(\prod_{\ell=1}^{i-1} J^\ell\right) \times \tilde{B}^{i,j} \times \left(\prod_{\ell=i+1}^\eta J^\ell \right) \quad\quad \forall i \in \llbracket \eta \rrbracket, \: j \in \llbracket t_i \rrbracket. \] \end{lemma} \begin{corollary}\label{prop:multilinear-IB-scheme} Let $J=\prod_{i=1}^\eta \llbracket d_i \rrbracket$ and $\mathscr{S}=\set{\prod_{i=1}^\eta \set{k_i,k_i+1} : k \in \prod_{i=1}^\eta \llbracket d_i -1 \rrbracket}$ describe a $\eta$-dimensional discretization of multilinear terms, and take $G_\mathscr{S}^c$ as its conflict graph. If for each $i \in \llbracket \eta \rrbracket$ we have a biclique cover $\{(\tilde{A}^{i,j},\tilde{B}^{i,j})\}_{j=1}^{t_i}$ for the conflict graph of SOS2$(d_i)$, then a biclique cover for $G_\mathscr{S}^c$ of depth $\sum_{i=1}^\eta t_i$ is given by $\bigcup_{i=1}^\eta \{(A^{i,j},B^{i,j})\}_{j=1}^{t_i}$, where \[ A^{i,j} = \left\{x \in J : x_i \in \tilde{A}^{i,j} \right\}, \quad B^{i,j} = \left\{x \in J : x_i \in \tilde{B}^{i,j} \right\} \quad\quad \forall i \in \llbracket \eta \rrbracket, j \in \llbracket t_i \rrbracket. \] In particular, if we take $\{h^{i,j}\}_{j=1}^{d_i-1} \subseteq \{0,1\}^{\lceil \log_2(d_i-1) \rceil}$ as a Gray code for each $i \in \llbracket \eta \rrbracket$, where $h^{i,0} \defeq h^{i,1}$ and $h^{i,d_i} \defeq h^{i,d_i-1}$, then a biclique cover for $G_\mathscr{S}^c$ of depth $\sum_{i=1}^\eta \lceil \log_2(d_i-1) \rceil$ is given by: \begin{align*} \hat{A}^{i,j} = \left\{x \in J : \exists \gamma \text{ s.t. } x_i = \gamma, \: h^{i,\gamma-1}_j = h^{i,\gamma}_j = 0 \right\},\quad \hat{B}^{i,j} = \left\{x \in J : \exists \gamma \text{ s.t. } x_i = \gamma, \: h^{i,\gamma-1}_j = h^{i,\gamma}_j = 1 \right\} \end{align*} for each $i \in \llbracket \eta \rrbracket$ and $j \in \llbracket \lceil \log_2(d_i-1) \rceil \rrbracket$. \end{corollary} We note that, since $|\mathscr{S}| = \prod_{i=1}^\eta (d_i-1)$, by Proposition \ref{prop:log-bound} this construction yields a formulation that is asymptotically optimal (with respect to number of auxiliary binary variables) for any possible MIP formulation, up to an additive factor of at most $\eta$. Furthermore, we can specialize this to the bilinear case studied by Misener et al. \cite{Misener:2011}. \begin{corollary} There exists a biclique cover for a grid discretization of a bilinear function ($\eta=2$) with $d_1=m+1$ and $d_2=1$ of depth $\lceil \log_2(m) \rceil$. \end{corollary} This result yields an ideal MIP formulation for the outer-approximation of bilinear terms with $\lceil \log_2(m) \rceil$ auxiliary binary variables, $2(m+1)$ auxiliary continuous variables (the $\lambda$ variables, one for element in $J$), and $2\lceil \log_2(m) \rceil$ general inequality constraints. In contrast, the logarithmic formulation from Misener et al. \cite{Misener:2011} has $\lceil \log_2(m) \rceil$ auxiliary binary variables, $2\lceil \log_2(m) \rceil + 1$ auxiliary continuous variables, at least $2\lceil \log_2(m) \rceil + 6$ general inequality constraints, and is not ideal in general (see Appendix~\ref{app:misener}). Therefore, we gain an ideal formulation with a naturally induced constraint branching at the price of a modest number of additional auxiliary continuous variables. Furthermore, our formulation generalizes readily to a discretization along the second dimension ($d_2 > 1$), for non-uniform discretizations, and for higher dimensional multilinear functions ($\eta > 2$). \subsection{Completing biclique covers via graph unions} Another useful graphical technique for our heuristic constructions will be to combine together biclique covers, each of which is designed to cover a substructure of the constraint. For example, the conflict graph of a grid triangulation of the plane is equivalent to the conflict graph of a $2$-dimensional grid discretization of multilinear terms, with one extra edge added for each subrectangle in the grid. Therefore, a biclique cover of a grid triangulation can be obtained from a biclique cover of a $2$-dimensional discretization of multilinear terms (i.e. from Corollary~\ref{prop:multilinear-IB-scheme}) by completing it with some number additional bicliques that cover those extra edges. This construction can be formalized in the following way. \begin{definition} The \emph{graph union} of a family of graphs $\set{G^i=(J^i,E^i)}_{i=1}^\eta$ is $\bigcup_{i=1}^\eta G^i\defeq (J_U,E_U)$, where $J_U=\bigcup_{i=1}^\eta J^i$ and $E_U=\bigcup_{i=1}^\eta E^i$. \end{definition} \begin{lemma} \label{lemma:graph-union} Take a family of graphs $\{G^i = (J^i,E^i)\}_{i=1}^\eta$ and a corresponding biclique cover $\{(A^j_i,B^j_i)\}_{j=1}^{t_i}$ of $G^i$ for each $i \in \llbracket \eta \rrbracket$. Then $\bigcup_{i=1}^\eta \{(A^j_i,B^j_i)\}_{j=1}^{t_i}$ is a biclique cover of $\bigcup_{i=1}^\eta G^i$. \end{lemma} We can apply Lemma~\ref{lemma:graph-union} to construct biclique covers for the grid triangulations depicted in Figure~\ref{fig:triangulations}. First, we apply the biclique cover construction from Corollary~\ref{prop:multilinear-IB-scheme} to cover all edges not sharing a subrectangle. This is depicted in the first two subfigures of each row in Figure~\ref{fig:triangulation-IBS}. To cover the remaining 4 edges created by the triangulation, we see that the number of additional levels needed is dependent on the combinatorial structure. Additionally, in all three cases we can verify through Proposition~\ref{prop:feasibility-IP} that the resulting biclique cover is of the smallest possible depth. \begin{figure}[htpb] \includegraphics[width=.185\linewidth]{Fig5a.pdf} \includegraphics[width=.185\linewidth]{Fig5b.pdf} \includegraphics[width=.185\linewidth]{Fig5c.pdf} \\ \includegraphics[width=.185\linewidth]{Fig5d.pdf} \includegraphics[width=.185\linewidth]{Fig5e.pdf} \includegraphics[width=.185\linewidth]{Fig5f.pdf} \includegraphics[width=.185\linewidth]{Fig5g.pdf} \\ \includegraphics[width=.185\linewidth]{Fig5h.pdf} \includegraphics[width=.185\linewidth]{Fig5i.pdf} \includegraphics[width=.185\linewidth]{Fig5j.pdf} \includegraphics[width=.185\linewidth]{Fig5k.pdf} \includegraphics[width=.185\linewidth]{Fig5l.pdf} \caption{Independent branching schemes for the three triangulations presented in Figure \ref{fig:triangulations}, each given its own row.The sets $A^j$ and $B^j$ are given by the blue squares and green diamonds, respectively, in the $j$-th subfigure of the corresponding row.} \label{fig:triangulation-IBS} \end{figure} The first example is the ``Union Jack'' triangulation \cite{Todd:1977} for $N=M=3$ and the results in \cite{Vielma:2009a} show that for this triangulation the biclique cover from Corollary~\ref{prop:multilinear-IB-scheme} can be completed with a single additional biclique cover for any $N$ and $M$. The second triangulation is a K1 triangulation \cite{Kuhn:1960} for $N=M=3$, and an early version of \cite{Vielma:2016} showed that for this triangulation the biclique can always be completed with two additional bicliques (See Section~\ref{chromaticsec} for a generalization of these results). In contrast, for generic triangulations such as the third one, it was not previously known if the biclique cover can always be completed with fewer than the trivial $(M-1)(N-1)$ levels needed to cover each ``diagonal'' edge with its own additional biclique. First, we can adapt Proposition~\ref{prop:simple-pIBS} to cover the extra edges with stars, but in general this will result in $\Theta(M\cdot N)$ stars, and hence the same number of additional levels. To reduce this, we need a way to \emph{stick} the stars together into more complicated bicliques. In section~\ref{chromaticsec} we will see how we may use graph colorings for a broad class of triangulations (subsuming the Union Jack and K1 triangulations as special cases), to cover the extra edges with either one or two additional bicliques. In general, it turns out that we may cover the remaining edges for any grid triangulation with a constant number of additional levels by applying the simple following lemma. \begin{lemma}\label{stickinglemma} Let $\{(A^j,B^j)\}_{j=1}^{t}$ be a family of bicliques of a graph $G$. If $(A^k,B^\ell)$ is also a biclique of $G$ for each $k,\ell\in\sidx{t}$, then $\bra{\bigcup_{j=1}^t A^j,\bigcup_{j=1}^t B^j}$ is a biclique of $G$. \end{lemma} The strength of Lemma~\ref{stickinglemma} comes from the fact that many CDCs of practical interest have a local structure (i.e. sets in $\mathscr{S}$ have small cardinality, or, equivalently, the minimum degree of the conflict graph is close to the total number of nodes). In this case, the condition of Lemma~\ref{stickinglemma} will hold for families of stars centered at nodes that are located ``sufficiently far apart.'' \subsubsection{Grid triangulations of the plane} \label{ss:grid-triangulations} We may now present a biclique cover construction for generic grid triangulations, with no further assumptions on the structure of the triangles such as in \cite{Vielma:2010,Vielma:2009a}, whose depth scales like $\log_2(M) + \log_2(N) + \mathscr{O}(1)$. In the same way as depicted in Figure~\ref{fig:triangulation-IBS}, we construct the biclique cover by using Lemma~\ref{lemma:graph-union} to complete the construction of Corollary~\ref{prop:multilinear-IB-scheme}. For this, we will use the following corollary of Lemma~\ref{stickinglemma} that shows how to combine certain stars centered at sufficiently separated nodes. \begin{corollary}\label{stencil1coro} Take a regular grid $J = \llbracket M \rrbracket \times \llbracket N \rrbracket$, let $\mathscr{S}$ be a grid triangulation of $[1,M]\times[1,N]$, and take $G_\mathscr{S}^c=(J,\bar{E})$ as its conflict graph. For all $w\in J$, define $A(w)\defeq\set{w}$ and $B(w) \defeq \set{w+v\,:\, v\in \set{-1,1}^2, \set{w,w+v}\in\bar{E} }$. Then \[ \left( \bigcup_{w \in J \cap (u+3\mathbb{Z}^2)} A(w), \bigcup_{w \in J \cap (u+3\mathbb{Z}^2)} B(w) \right) \] is a biclique of $G_\mathscr{S}^c$ for any $u\in J$. \end{corollary} \proof{\textbf{Proof}} Direct from Lemma~\ref{stickinglemma} by taking $u \in J$ and the family of bicliques $\{(A(w),B(w)\}_{w \in J \cap (u + 3\mathbb{Z}^2)}$ and noting that, if $u,v\in J \cap (u+3\mathbb{Z}^2)$, then $||u-v||_\infty \geq 3$, and so $(A(u),B(v))$ is also a biclique for $G^c_\mathscr{S}$. \Halmos\endproof Figure~\ref{fig:pwl} shows two possible bicliques that can be obtained from Corollary~\ref{stencil1coro}. \begin{figure}[htpb] \centering \includegraphics[width=.49\linewidth]{Fig6a.pdf} \includegraphics[width=.49\linewidth]{Fig6b.pdf} \caption{Two bicliques constructed via Corollary~\ref{stencil1coro} for a grid triangulation with $M=N=8$. On the left the construction follows by taking $u=(1,1)$; on the right, with $u=(2,3)$. For each level, the sets $A(u)$ and $B(u)$ are given by the squares and diamonds, respectively.} \label{fig:pwl} \end{figure} We can now use Lemma~\ref{lemma:graph-union} and Corollary~\ref{stencil1coro}, along with the biclique cover derived in Corollary~\ref{prop:multilinear-IB-scheme}, to obtain a biclique cover for any triangulation with an asymptotically optimal number of levels. \begin{theorem} \label{thm:pwl} Take $J = \llbracket M \rrbracket \times \llbracket N \rrbracket$, and let $\mathscr{S}$ be a grid triangulation of $[1,M]\times[1,N]$. Take $G^c_\mathscr{S}=(J,\bar{E})$ as its conflict graph. Presume that $\{(\tilde{A}^{1,j},\tilde{B}^{1,j})\}_{j=1}^{t_1}$ and $\{(\tilde{A}^{2,j},\tilde{B}^{2,j})\}_{j=1}^{t_2}$ are biclique covers for the conflict graphs of the SOS2($M$) and SOS2($N$) constraints, respectively. Furthermore, define \begin{align*} A^{3,u} &= J \cap (u + 3\mathbb{Z}^2) \\ B^{3,u} &= \bigcup_{w\in J\cap\bra{u+3\mathbb{Z}^2}}\set{w+v\,:\,v\in \set{-1,1}^2,\: \set{w,w+v}\in\bar{E} } \end{align*} for each $u \in \{0,1,2\}^2$. Then $\{(A^{1,j},B^{1,j})\}_{j=1}^{t_1} \cup \{(A^{2,j},B^{2,j})\}_{j=1}^{t_2} \cup \{(A^{3,u},B^{3,u})\}_{u \in \{0,1,2\}^2}$ is a biclique cover for $G^c_\mathscr{S}$, where \begin{alignat*}{2} A^{1,j} &= \tilde{A}^{1,j} \times \llbracket N \rrbracket, \quad\quad B^{1,j} &= \tilde{B}^{1,j} \times \llbracket N \rrbracket, \\ A^{2,j'} &= \llbracket M \rrbracket \times \tilde{A}^{2,j'}, \quad\quad B^{2,j'} &= \llbracket M \rrbracket \times \tilde{B}^{2,j'}, \end{alignat*} for each $j \in \llbracket t_1 \rrbracket$ and $j' \in \llbracket t_2 \rrbracket$. In particular, if $\{h^{1,i}\}_{i=1}^{M-1} \subseteq \{0,1\}^{\lceil \log_2(M-1) \rceil}$ and $\{h^{2,i}\}_{i=1}^{N} \subseteq \{0,1\}^{\lceil\log_2(N-1)\rceil}$ are Gray codes, where $h^{1,0} \defeq h^{1,1}$, $h^{1,M} \defeq h^{1,M-1}$, $h^{2,0} \defeq h^{2,1}$, and $h^{2,N} \defeq h^{2,N-1}$, then a biclique cover of $G_\mathscr{S}^c$ of depth $\lceil \log_2(M-1) \rceil + \lceil \log_2(N-1) \rceil + 9$ is given by: \begin{subequations}\label{generalPWL} \begin{align} A^{1,j} &= \big\{ (x,y) \in J : h^{1,x-1}_j = h^{1,x}_j = 1 \big\} \\ B^{1,j} &= \big\{ (x,y) \in J : h^{1,x-1}_j = h^{1,x}_j = 0 \big\} \\ A^{2,j'} &= \big\{ (x,y) \in J : h^{2,y-1}_{j'} = h^{2,y}_{j'} = 1 \big\} \\ B^{2,j'} &= \big\{ (x,y) \in J : h^{2,y-1}_{j'} = h^{2,y}_{j'} = 0 \big\} \\ A^{3,u} &= J \cap (u + 3\mathbb{Z}^2) \\ B^{3,u} &= \bigcup_{w\in J\cap\bra{u+3\mathbb{Z}^2}}\set{w+v\,:\,v\in \set{-1,1}^2,\: \set{w,w+v}\in\bar{E} } \end{align} \end{subequations} for all $j \in \llbracket \lceil \log_2(M-1) \rceil \rrbracket$, $j' \in \llbracket \lceil \log_2(N-1) \rceil \rrbracket$, and $u \in \{0,1,2\}^2$. \end{theorem} \proof{\textbf{Proof}} Let $G^x \defeq (\llbracket M \rrbracket,E^x)$ and $G^y \defeq (\llbracket N \rrbracket,E^y)$ be the conflict graphs for SOS2($M$) and SOS2($N$), respectively. Furthermore, let \[ G^3 \defeq \bigcup_{u \in \{0,1,2\}^2}(J,A^{3,u} * B^{3,u})=\left(J, \bigcup_{u \in \{0,1,2\}^2}(A^{3,u} * B^{3,u})\right). \] Then we see that $G_\mathscr{S} = (G^x \times G^y) \cup G^3$ by noting that all \emph{diagonal} edges of $E$ (i.e. those of the form $\set{w,w+v}\in E$ for $w\in J$ and $v\in \set{-1,1}^2$) are included in $G^3$, and observing that $G^3$ is a subgraph of $G^c_\mathscr{S}$. The result then follows from Lemma~\ref{lemma:graph-product}, Lemma~\ref{lemma:graph-union}, and Corollary~\ref{stencil1coro}. \Halmos\endproof By referring to Proposition \ref{prop:log-bound}, we recover a $\lceil \log_2(2(M-1)(N-1)) \rceil \geq \lceil\log_2(M-1)\rceil + \lceil\log_2(N-1)\rceil$ lower bound on the depth of any biclique cover for a grid triangulation, and see that our construction yields a MIP formulation that is within a constant additive factor of the smallest possible. Finally, we note that similarly to the results in \citep{Vielma:2010,Vielma:2009a} for the Union Jack triangulation, formulation \eqref{eqn:ideal-formulation-for-IBS} for biclique cover \eqref{generalPWL} can provide a significant computational advantage for general grid triangulations \cite{Huchette:2017}. \subsubsection{SOS$k$}\label{ss:sosk} In this subsection, we will see how we may use the graph union construction to produce an IB scheme for SOS$k$($N$) of depth $\log_2(N/k)+\mathscr{O}(k)$, for any $k$ and $N$. Similar to the construction for grid triangulations, we first construct an initial family of bicliques based on the SOS2 constraint. Next, we expand this onto a larger node set by the graph product construction. Finally, we complete the biclique cover by combining a family of sufficiently separated stars. For grid triangulations, this approach meant applying SOS2 constraints horizontally and vertically, and taking a graph product of the two. One way to interpret this is as an SOS2 constraint applied to groups of aggregated nodes in the ground set (e.g. when SOS2 is applied horizontally, we group all elements with the same horizontal coordinate into a single group). For the SOS$k$($N$) constraint, we will apply the SOS2 constraint to the groups obtained by partitioning the $N$ original ground elements into $\lceil N/k \rceil$ subsets of $k$ consecutive elements. The following simple lemma shows how this grouping can also be represented through a graph product. For the remainder of the section, we assume that $N/k$ is integer; if this is not true, we artificially introduce $\lceil N/k \rceil k - N$ nodes such that this is the case, construct the formulation in Theorem~\ref{thm:sosk}, and remove the artificial nodes from the formulation afterwards. \begin{lemma}\label{soskproduct} Let $J = \llbracket N \rrbracket$, $k\leq N $, $\mathscr{S}$ correspond to the SOS$k(N)$ constraint, and $G_\mathscr{S}^c=(J,\bar{E})$ be the corresponding conflict graph. Let $G^1 = (\llbracket N/k \rrbracket,E^1)$ be the conflict graph for SOS2($N/k$) and $G^2 = (\{0,\ldots,k-1\},\emptyset)$ be the empty graph on $k$ nodes. Then $G^1 \times G^2$ is isomorphic to a subgraph $\hat{G}=(J,\hat{E})$ of $G^c_\mathscr{S}$ wherein $\hat{E} \subseteq \bar{E}$, and each edge $\{u,v\} \in \bar{E}$ with $|u-v| \geq 2k$ is contained $\{u,v\} \in \hat{E}$. \end{lemma} \proof{\textbf{Proof}} Let $G^1 \times G^2 = (J',E')$. Consider the bijection $f:J\to J'$ given by $f(u)=(\div(u,k),\operatorname{mod}(u,k))$, where $\div(u,k)=\lfloor u/k \rfloor$ and $\operatorname{mod}(u,k)=u-k \div(u,k)$ are the quotient and remainder of the division of $u$ by $k$, so that $f^{-1}\bra{m,r}=km+r$. We have that $\set{(m,r),(m',r')} \in E'$ if and only if $\set{m,m'} \in E^1$, which in turn is equivalent to $|m-m'| \geq 2$. Therefore, for any $\set{(m,r),(m',r')} \in E'$, we have \[ \abs{f^{-1}(m,r) -f^{-1}(m',r')}= |(km+r)-(km'+r')| = |k(m-m')+(r-r')| \geq k|m-m'| + |r-r'| \geq 2k, \] and hence $\set{f^{-1}(m,r),f^{-1}(m',r')} \in \bar{E}$, i.e. $\hat{E} \subseteq \bar{E}$. For the second condition, see that if $u,v\in J$ are such that $\abs{u-v}\geq 2k$, then $|\div(u)-\div(v)| \geq 2$, and therefore $\set{f(u),f(v)} \in E'$. \Halmos\endproof We can then cover the remaining edges with the following bicliques obtained by stitching together families of sufficiently separated stars. \begin{corollary}\label{stencil2coro} Let $J = \llbracket N \rrbracket$, $k\leq N $, $\mathscr{S}$ correspond to the SOS$k(N)$ constraint, and $G_\mathscr{S}^c$ be the corresponding conflict graph. For all $w\in J$, define $A(w) \defeq \set{w}$ and $B(w) \defeq \set{u\in J\,:\, k\leq \abs{u-w}< 2k }$. Then \[ \bra{\bigcup_{w\in J\cap\bra{u+3k\mathbb{Z}}} A(w),\bigcup_{w\in J\cap\bra{u+3k\mathbb{Z}}} B(w)} \] is a biclique of $G_\mathscr{S}^c$ for any $u\in J$. \end{corollary} \proof{\textbf{Proof}} Direct from Lemma~\ref{stickinglemma} by considering the family of bicliques $\{(A(w),B(w)\}_{w \in J \cap (u+3k\mathbb{Z})}$ and noting that, for distinct $u,v\in J \cap (u+3k\mathbb{Z})$, $|u-v| \geq 3k$, and so $(A(u),B(v))$ is also a biclique for $G^c_\mathscr{S}$. \Halmos\endproof Finally, we can combine both classes of bicliques with Lemma~\ref{lemma:graph-union} to construct a complete biclique cover for SOS$k(N)$. See Figure~\ref{fig:sos3-26} for an example of the resulting construction. \begin{theorem} \label{thm:sosk} Let $J = \llbracket N \rrbracket$, $k\leq N $, $\mathscr{S}$ correspond to the SOS$k(N)$ constraint on $J$, and $G_\mathscr{S}^c$ be the corresponding conflict graph. Let $\{(\tilde{A}^{1,j},\tilde{B}^{1,j})\}_{j=1}^{t_1}$ be a biclique cover for the conflict graph of the SOS2($N/k$) constraint, and take \begin{align*} A^{2,j'} &= \bigcup_{i=0}^{\left\lceil \frac{N}{3k} \right\rceil} \left\{ \tau \in J : \tau = j' + (3i-3)k \right\} \\ B^{2,j'} &= \bigcup_{i=0}^{\left\lceil \frac{N}{3k} \right\rceil} \left\{ \tau \in J : j'+(3i-2)k \leq \tau \leq j'+(3i-1)k \right\} \end{align*} for all $j' \in \llbracket 3k \rrbracket$. Then $\{(A^{1,j},B^{1,j})\}_{j=1}^{t} \cup \{(A^{2,j'},B^{2,j'})\}_{j'=1}^{\llbracket 3k \rrbracket} $ is a biclique cover for $G_\mathscr{S}^c$, where \begin{alignat*}{2} A^{1,j} &= \set{ \tau \in J\,:\, \lceil \tau/k \rceil \in \tilde{A}^{1,j} }, \quad\quad B^{1,j} &=\set{ \tau \in J\,:\, \lceil \tau/k \rceil \in\tilde{B}^{1,j}}, \end{alignat*} for each $j \in \llbracket t \rrbracket$. In particular, if $\{h^i\}_{i=1}^{\lceil N/k \rceil-1} \subseteq \{0,1\}^{\lceil \log_2(\lceil N/k \rceil-1) \rceil}$ is a Gray code where $h^0 \defeq h^1$ and $h^{\lceil N/k \rceil } \defeq h^{\lceil N/k \rceil-1}$, then a biclique cover of $G_\mathscr{S}^c$ of depth $\lceil \log_2(\lceil N/k \rceil-1)\rceil+3k$ is given by $\{(A^{1,j},B^{1,j})\}_{j=1}^{\lceil \log_2(\lceil N/k \rceil -1)\rceil} \cup \{(A^{2,j'},B^{2,j'})\}_{j'=1}^{\llbracket 3k \rrbracket} $, where \begin{align*} A^{1,j} = \left\{ \tau \in J : h^{\lceil \tau / k \rceil-1}_j = h^{\lceil \tau / k \rceil}_j = 0 \right\},\quad \quad B^{1,j} = \left\{ \tau \in J : h^{\lceil \tau / k \rceil-1}_j = h^{\lceil \tau / k \rceil}_j = 1 \right\} \end{align*} for all $j \in \llbracket \lceil \log_2(\lceil N/k \rceil-1) \rceil \rrbracket$. \end{theorem} \proof{\textbf{Proof}} Take $G^1 \defeq (\llbracket N/k \rrbracket,E^1)$ as the conflict graph for SOS2($N/k$), $G^2 \defeq (\{0,\ldots,k-1\},\emptyset)$ as the empty graph on $k$ nodes, and $G^3 \defeq \bra{J, \bigcup_{j'=1}^{3k} A^{2,j'} * B^{2,j'}}$. Let $\hat{G}$ be the subgraph of $G_\mathscr{S}^c$ from Lemma~\ref{soskproduct}, which is isomorphic to $G^1 \times G^2$ through the bijection $g:\llbracket N/k \rrbracket\times \{0,\ldots,k-1\}\to J$ with $g\bra{m,r}=km+r$. Then we have that $G^c_\mathscr{S}= \hat{G}\cup G^3$, after applying Lemma~\ref{soskproduct} and using the fact that the edges of $G^3$ contain the edges of $G^c_\mathscr{S}$ not included in $\hat{G}$. The result then follows from Lemma~\ref{lemma:graph-product}, Lemma~\ref{lemma:graph-union}, Lemma~\ref{soskproduct}, and Corollary~\ref{stencil1coro}. \Halmos\endproof \begin{figure}[htpb] \centering \includegraphics[width=\linewidth]{Fig7a.pdf} \includegraphics[width=\linewidth]{Fig7b.pdf} \includegraphics[width=\linewidth]{Fig7c.pdf} \includegraphics[width=\linewidth]{Fig7d.pdf} \includegraphics[width=\linewidth]{Fig7e.pdf} \includegraphics[width=\linewidth]{Fig7f.pdf} \includegraphics[width=\linewidth]{Fig7g.pdf} \includegraphics[width=\linewidth]{Fig7h.pdf} \includegraphics[width=\linewidth]{Fig7i.pdf} \includegraphics[width=\linewidth]{Fig7j.pdf} \includegraphics[width=\linewidth]{Fig7k.pdf} \includegraphics[width=\linewidth]{Fig7l.pdf} \caption{Visualizations of the biclique cover from the proof of Theorem~\ref{thm:sosk} for SOS$3(26)$. Each row corresponds to some level $j$, and the sets $A^j$ and $B^j$ are the squares and diamonds, respectively. The first three rows correspond the the ``first stage'' of the biclique cover $\{(A^{1,j},B^{1,j})\}_{j=1}^3$, and the second nine correspond to the ``second stage'' $\{(A^{2,j},B^{2,j})\}_{j=1}^9$.} \label{fig:sos3-26} \end{figure} We note that, when $k = \mathscr{O}(\log(N))$, this biclique cover yields a MIP formulation that is asymptotically tight (with respect to the number of auxiliary binary variables) with our lower bound of $\lceil \log_2(N-k+1) \rceil$ from Proposition~\ref{prop:log-bound}. We can also show an absolute lower bound of depth $k$ for any biclique cover for SOS$k$. This implies that when $k = \omega(\log(N))$, although the formulation from Theorem~\ref{thm:sosk} is not tight with respect to the lower bound from Proposition~\ref{prop:log-bound}, it is asymptotically the smallest possible formulation in the pairwise IB framework. \begin{proposition} \label{prop:sosk-lowerbound} Any biclique cover for the conflict graph of SOS$k(N)$ must have depth at least $\min\{k,N-k\}$. \end{proposition} \proof{\textbf{Proof}} Define $\gamma \defeq \min\{k, N-k\}$ and consider any possible biclique cover $\{(A^j,B^j)\}_{j=1}^t$. The biclique cover must separate the edges $\{(\tau,\tau+k)\}_{\tau=1}^{\gamma}$. Consider a level $j$ of the biclique cover that contains edge $\{\tau,\tau+k\}$ for some $\tau \in \llbracket \gamma \rrbracket$; w.l.o.g., $\tau \in A^j$ and $\tau+k \in B^j$. Consider the possibility that the same level $j$ separates another such edge in the set, e.g. $(\tau',\tau'+k)$ for $\tau' \in \llbracket \gamma \rrbracket$, where w.l.o.g. $\tau < \tau'$. That would imply that either $\tau' \in A^j$ or $\tau' \in B^j$. In the case that $\tau' \in A^j$, we have that $\bar{E}^j$ contains the edge $\{\tau',\tau+k\}$. However, since $|(\tau+k)-\tau'| = \tau+k - \tau' < \tau+k-\tau = k$, this implies that the biclique cover separates a feasible edge, a contradiction. In the case where $\tau' \in B^j$, we have that $\bar{E}^j$ contains the edge $\{\tau,\tau'\}$, and as $\tau'-\tau < k \leq \gamma$ from the definition of our set of edges, a similar argument holds. Therefore, each edge $\{\{\tau,\tau+k\}\}_{\tau=1}^{\gamma}$ must be uniquely contained in some level of the biclique cover, giving the result. \Halmos\endproof Furthermore, when $k = \lfloor N/2 \rfloor$, this proposition gives a lower bound on the depth of a biclique cover that is asymptotically tight with the upper bound of $N$ from Proposition~\ref{prop:simple-pIBS}. In other words, in this particular regime, we have that the SOS$k$ constraint admits a pairwise IB-based formulation, but only one that is relatively large ($\Omega(|\mathscr{S}|) = \Omega(|J|)$ auxiliary binary variables and constraints). \subsection{Covering edges with a chromatic characterization of bicliques}\label{chromaticsec} Our final contribution is to adapt a result of Cornaz and Fonlupt~\cite{Cornaz:2006} that gives a chromatic characterization of the set of edges that can be covered by a biclique. \begin{theorem}[\cite{Cornaz:2006}]\label{cornaztheo} Take the graph $G=(J,\bar{E})$, along with some edge subset $\bar{F} \subseteq \bar{E}$. Define $V(\bar{F}) = \bigcup\{\{u,v\} \in \bar{F}\}$ as all nodes incident to $\bar{F}$, and take $F = [(\bar{F} * \bar{F}) \backslash \bar{E}$ as all pairs incident to the edges $V(\bar{F})$ not contained in $\bar{E}$. Define both $E' = F \cup \bar{F}$ and $p:E'\to \set{0,1}$ such that $p(e)= \mathds{1}[e \in \bar{E}]$\footnote{Where $\mathds{1}[e \in \bar{E}]=1$ if $e \in \bar{E}$ and $\mathds{1}[e \in \bar{E}]=0$ otherwise.}. Finally, take $\mathcal{C}\bra{E'}$ as the family of all cycles in $G'$. Then the following are equivalent: \begin{enumerate} \item There exists a biclique $(A,B)$ of $G$ covering $(V(\bar{F}),\bar{F})$. \item For all $C\in \mathcal{C}\bra{E'}$, $\sum\nolimits_{u\in C} p(u)$ is even. \item There exists some $f:V(\bar{F})\to \set{0,1}$ such that \begin{itemize} \item $f(u)=f(v)$ for all $\set{u,v}\in F$, \item $f(u)\neq f(v)$ for all $\set{u,v}\in \bar{F}$, and \item $\bra{\set{u\in V(\bar{F})\,:\, f(u)=0},\set{u\in V(\bar{F})\,:\, f(u)=1}}$ is a biclique of $G$ covering $(V(\bar{F}),\bar{F})$. \end{itemize} \end{enumerate} \end{theorem} Using this result, we can sometimes exploit the structure of a grid triangulations to obtain biclique covers that are smaller than those derived in Theorem~\ref{thm:pwl}. \begin{theorem} \label{thm:pwl2} Take a regular grid $J = \llbracket M \rrbracket \times \llbracket N \rrbracket$, the sets $\mathscr{S}$ given by a grid triangulation of $[1,M]\times[1,N]$ with $\bigcup\{S \in \mathscr{S}\} = J$, and take $G_\mathscr{S}^c=(J,\bar{E})$ to be the corresponding conflict graph. Furthermore, let $\{(\tilde{A}^{1,j},\tilde{B}^{1,j})\}_{j=1}^{t_1}$ and $\{(\tilde{A}^{2,j},\tilde{B}^{2,j})\}_{j=1}^{t_2}$ be biclique covers for the conflict graphs of the SOS2($M$) and SOS2($N$) constraints, respectively. Also define $J_{even} \defeq \set{u\in J\,:\, u_1 \equiv u_2 \mod 2}$ and $J_{odd}\defeq\set{u\in J\,:\, u_1 \not\equiv u_2 \mod 2}$ as the set of nodes whose sum of components is either even and odd, respectively. For each $s \in \{even,odd\}$, let $E_s\defeq \set{\set{u,v} \in [J_s]^2 : \left\|u-v\right\|_\infty=1}$, $\bar{F}_s\defeq E_s\cap\bar{E}$, and $F_s\defeq ([V\bra{\bar{F}_s}]^2 \cap E_s) \backslash \bar{E}$. If, for each $s \in \{even,odd\}$, there exists $f_s:J_s\to \set{0,1}$ such that \begin{equation}\label{pwlcolorcond} f_s(u)=f_s(v) \quad \forall \set{u,v}\in F_s\quad \quad \text{ and } \quad\quad f_s(u)\neq f_s(v)\quad \forall\set{u,v}\in \bar{F}_s, \end{equation} then $\{(A^{1,j},B^{1,j})\}_{j=1}^{t_1} \cup \{(A^{2,j},B^{2,j})\}_{j=1}^{t_2} \cup \{(A^{3,i},B^{3,i})\}_{i\in\sidx{2}}$ is a biclique cover for $G_\mathscr{S}^c$, where \begin{alignat*}{2} A^{1,j} &= \tilde{A}^{1,j} \times \llbracket N \rrbracket, \quad\quad& B^{1,j} &= \tilde{B}^{1,j} \times \llbracket N \rrbracket, \\ A^{2,j'} &= \llbracket M \rrbracket \times \tilde{A}^{2,j'}, \quad\quad& B^{2,j'} &= \llbracket M \rrbracket \times \tilde{B}^{2,j'},\\ A^{3,s} &= \set{u\in J_s\,:\, f_s(u)=0} , \quad\quad& B^{3,s} &=\set{u\in J_s\,:\, f_s(u)=1} \end{alignat*} for each $j \in \llbracket t_1 \rrbracket$, $j' \in \llbracket t_2 \rrbracket$, and $s \in \{even,odd\}$. If we select $\{(\tilde{A}^{1,j},\tilde{B}^{1,j})\}_{j=1}^{t_1}$ and $\{(\tilde{A}^{2,j},\tilde{B}^{2,j})\}_{j=1}^{t_1}$ to correspond to the Gray code construction for SOS2 in Proposition~\ref{prop:sos2-IB-scheme}, then the resulting biclique cover has depth $\lceil \log_2(M-1) \rceil + \lceil \log_2(N-1) \rceil + 2$. Furthermore, if for some $s \in \{even,odd\}$ we have that $f_s(u)=\alpha$ for all $u\in J_s$ and some constant $\alpha$, we may reduce the depth to $\lceil \log_2(M-1) \rceil + \lceil \log_2(N-1) \rceil + 1$. Finally, if we fix $s \in \{even,odd\}$ and $r \in \{even,odd\} \backslash \{s\}$, a sufficent condition for the existence of $f_s:J_s\to \set{0,1}$ satisfying \eqref{pwlcolorcond} is that \begin{equation}\label{odddegreecond} d_{F_{r}}(u) \defeq \left|\left\{ e \in F_{r} : \exists v\in J_{r}\text{ s.t. } e=\set{u,v} \right\}\right| \text{ is even } \quad \forall u\in J_{r}\cap \llbracket 2,M-1 \rrbracket \times \llbracket 2,N-1 \rrbracket. \end{equation} \end{theorem} \proof{\textbf{Proof}} Let $G^x \defeq (\llbracket M \rrbracket,E^x)$ and $G^y \defeq (\llbracket N \rrbracket,E^y)$ be the conflict graphs for the SOS2($M$) and SOS2($N$) constraints, respectively. Furthermore, let $G^{3} \defeq (\llbracket M \rrbracket \times \llbracket N \rrbracket,(A^{3,even} * B^{3,even}) \cup (A^{3,odd} * B^{3,odd}))$. Then we may see that $G^c_\mathscr{S} = (G^x \times G^y) \cup G^3$ by noting that all \emph{diagonal} edges of $\bar{E}$ (i.e. those of the form $\set{w,w+v}\in \bar{E}$ for $w\in J$ and $v\in \set{-1,1}^2$) are included in $G^3$, and that $G^3$ is a subgraph of $G^c_\mathscr{S}$. The first part of the theorem then follows from Lemma~\ref{lemma:graph-product}, Lemma~\ref{lemma:graph-union}, and Theorem~\ref{cornaztheo}. For the sufficient condition, w.l.o.g. consider the case where $s=even$ and $r=odd$. Define $p_{even}:E_{even}\to \set{0,1}$ as $p_{even}(e) = \mathds{1}[e \in \bar{E}]$. The result will follow from Theorem~\ref{cornaztheo} by showing that \eqref{odddegreecond} satisfies condition 2 in the equivalence of the theorem. Let $E'_{even}=F_{even}\cup\bar{F}_{even}$, and assume for contradiction that there exists $C\in \mathcal{C}\bra{E'_{even}}$ such that $\sum\nolimits_{u\in C} p_{even}(u)$ is odd. If $\abs{C}=4$, we may assume without loss of generality that $V(C)=\set{u,u+(1,1),u+(1,-1),u+(2,0)}\subset J_{even}$ for some $u\in J_{even}$. Then $v=u+(0,1)\in J_{odd}$ is such that $d_{E_{odd}}(v)$ is odd, a contradiction of \eqref{odddegreecond}. If $\abs{C}>4$, note that $C\in \mathcal{C}\bra{E_{even}}$ and that there exists $e\in C$ such that $p_{even}(e)=1$. In addition, there exists $C_0\in \mathcal{C}\bra{E_{even}}$ such that $e\in C_0$, $\abs{C_0}=4$, $C_1=\bra{C_0\cup C}\setminus \bra{C_0\cap C}\in \mathcal{C}\bra{E_{even}}$ and $\operatorname{Conv}(C_1) \subsetneq \operatorname{Conv}(C)$. If $\sum\nolimits_{u\in C_0} p_{even}(u)$ is odd, we may make the same argument above as $|C_0|=4$ to derive a contradiction of \eqref{odddegreecond}. If not, then $C_1$ and $C$ have the same parity, and therefore $\sum\nolimits_{u\in C_1} p_{even}(u)$ is odd. We may then repeat this shrinking procedure recursively on $C_1$ until either the $4$-cycle $C_0$ has odd parity, or $C_1$ is itself a $4$-cycle. In either case, we have a $4$-cycle with odd parity, which implies some $u\in J_{odd}$ that violates \eqref{odddegreecond}, giving the result. \Halmos\endproof We note that we may use this coloring characterization to recover the biclique covers for both the Union Jack and K1 triangulation example in Figure~\ref{fig:triangulation-IBS}. For the Union Jack example, we have that \[ \bar{F}_{even} = \emptyset, \quad\quad \bar{F}_{odd} = \left\{\{(1,2),(2,1)\},\:\{(2,1),(3,2)\},\:\{(1,2),(2,3)\},\:\{(2,3),(3,2)\}\right\}, \] and so we may apply the simplification in Theorem~\ref{thm:pwl2} to construct a biclique cover of depth $\log_2(2)+\log_2(2)+1=3$. Indeed, the original formulation of Vielma and Nemhauser~\cite{Vielma:2009a} for the Union Jack triangulation can be reinterpreted analogously through the chromatic characterization of Theorem~\ref{thm:pwl2}. For the K1 triangulation example, we have \[ \bar{F}_{even} = \left\{ \{(1,3),(2,2)\},\:\{(2,2),(3,1)\} \right\}, \quad\quad \bar{F}_{odd} = \left\{\{(1,2),(2,1)\},\:\{(2,3),(3,2)\} \right\}, \] giving a biclique cover of depth $\log_2(2)+\log_2(2)+2=4$. Furthermore, we close by noting that the sufficient condition in Theorem~\ref{thm:pwl2} is, in general, not necessary. For example, in Figure~\ref{fig:coloring-example} we see a grid triangulation that does not satisfy \eqref{odddegreecond}, but for which there exists a coloring given by $\{f_{even},f_{odd}\}$ that satisfies \eqref{pwlcolorcond}. That is, $(2,3)$ has odd degree (i.e. $d_{F_{odd}}((2,3)) = 1$), and so the sufficient condition is not satisfied. However, a coloring satisfying the conditions of Theorem~\ref{thm:pwl2} exists. This offers a generalization of the result that originally appeared in a preliminary version of \cite{Vielma:2016}, which only showed the sufficient condition \eqref{odddegreecond}. \begin{figure}[htpb] \centering \includegraphics[width=.32\linewidth]{Fig8a.pdf} \includegraphics[width=.32\linewidth]{Fig8b.pdf} \includegraphics[width=.32\linewidth]{Fig8c.pdf} \caption{(Left) A grid triangulation that does not satisfy the sufficient condition of Theorem~\ref{thm:pwl2}, as $(2,3)$ (circle) has odd degree. However, there a coloring of the form described in Theorem~\ref{thm:pwl2}, leading to two bicliques (Center and Right) that cover all ``diagonal'' edges of the conflict graph.} \label{fig:coloring-example} \end{figure} \section*{Acknowledgements} This material is based upon work supported by the National Science Foundation Graduate Research Fellowship under Grant No. 1122374, and by the National Science Foundation under Grant No. CMMI-1351619. The authors would like to thank Yves Crama for pointing out the relation between biclique covers, independent sets, and boolean functions. \bibliographystyle{ormsv080}
1d180dc0d0d38bd255ea09094e7793ee2e19bd8a
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Recently, we proposed Prototypes as a way to represent knowledge on the web~\cite{researchPaper}\footnote{ Please use that paper as a reference for all definitions.}. That paper has its focus on theoretical aspects and analysis. In this resource paper we describe the tools we developed to deploy prototypes. First, in \cref{sec:standalone} we present the implementation of a knowledge base, based on the implementation of a Java interface, which can be used for storing prototypes. Then, we reuse this system to show how the knowledge base can be used in remote and distributed settings (\cref{sec:distributed}). Each of the sections includes some amount of benchmarking to give the reader an impression of the practical re-usability of the provided solutions. We do assume that the reader has some familiarity with the ideas behind prototypes. The implementation and the code used for the benchmarks is licensed under the LGPLv3 license and can be downloaded from \url{https://github.com/miselico/knowledgebase}. \section{A Standalone Knowledge Base} \label{sec:standalone} A prototype knowledge base (KB) consists of a collection of prototypes. To mirror this, the IKnowledge interface, which we define as the basis for a KB, has only one method which must\footnote{Other methods are Java 8 default methods.} be implemented. The signature of this method is \lstinline$Optional<? extends Prototype> isDefined(ID id);$. The provided Java source code contains five implementations of this interface, namely \begin{description} \item[EmptyKnowledgeBase] is a KB without any content. However, as per the definition, it still contains the empty prototype $\emptyProtoID$. \item[PredefinedKB] is a KB containing string and integer constants and is described in more details in this section. \item[KnowledgeBase] stores prototypes. It can be constructed using another \lstinline$IKnowledgeBase$ as a basis. The underlying basis will be queried in case the requested prototype is not directly defined in the \lstinline$KnowledgeBase$. \item[RemoteKB] gets its prototypes from a remote KB. This implementation is described further in~\cref{sec:remote}. \item[ChainedKB] is an \lstinline$IknowledgeBase$ which connects multiple \lstinline$IKnowledgeBase$s together. The KBs are checked in turn until one where the Prototype is defined is found. If none is found, an empty \lstinline$Optional$ is returned, indicating that no Prototype could be found. \end{description} Each prototype consists of four components, namely \begin{inparaenum}[\itshape 1\upshape)] \item its own ID, \item the ID of its base, \item the change set for adding parts, and \item the change set for removing parts~\cite{researchPaper}. \end{inparaenum} This structure is closely mimicked in our implementation. The IDs are essentially represented using String types and the change sets using multimaps (i.e., maps which can associate multiple values for a given key). The formal definition allows the creation of a prototype which would remove all values for a given property. According to the theoretical definition, that computation would involve the set of all possible IDs by enumeration, which is unfeasible. Hence % our implementation has two distinct changeset implementations and treats the `remove all' as a special case which does not require enumeration. Another aspect which a concrete implementation should cover is the use of literals. At its current state the formal Prototype KB definition does not support literals as the value of a property. Instead, one has to represent a literal by using an agreed prototype. Therefore, we designed \lstinline$PredefinedKB$ which acts like a KB which implicitly contains all possible string and integer literals encoded as prototypes. The extension of the supported literal types to any type is facilitated. \subsection{Consistency Checking} When a KB is created out of a set of prototypes, it should be checked that the result is in accordance with our Prototype Knowledge Base definition \cite[Definition 4]{researchPaper}. In this paper we say that the KB must be checked for consistency. This consistency check is performed in the \lstinline$KnowledgeBase$ implementation. First, all IDs and property names must be valid absolute IRIs, which is enforced by the type system. Next, if there is any prototype with a definition involving an $ID$ which cannot be found in the KB, then the creation will be refused. Then, it is checked whether all inheritance chains eventually (recursively) end up at the empty prototype. If also that is the case then a check is performed to ensure that no ID is used twice (this includes checking the underlying KB). In practice, one might want to remove this last check and the issue of duplicates could be resolved in favor of prototypes in a given KB. This is possible using the \lstinline$ChainedKB$. The design of our software helps to build KB which are consistent. The KB provides a builder class which can be used for construction. Further, the KB itself is completely immutable. Changes are made by creating a new KB. This ensures the consistency at any point in time. \subsection{Fixpoint Computation} Given a knowledge base $KB_{o}$ we implemented a method to compute its interpretation $KB_{n}$. This interpretation contains for each prototype definition with ID $id$ in $KB_o$ a new prototype definition of the form $(id, (\text{ \url{PROTO:P_0} }, ADD, \emptyset))$. Where $ADD$ is such that under an interpretation $I_{KB}(KB_o) = I_{KB} (KB_n)$. This boils down to computing the fixpoint for each of the prototype expressions. However, a direct implementation of the definition would not work since there is a universal quantification over all IDs (an infinite set). Hence, we implement this such that simple change expressions are created only for these IDs which are actually used. We implemented both the consistency check and the fixpoint computation in a scalable fashion. For both the consistency check and the computation of the fixpoints, the implementation is optimized such that it will not compute things twice. When the fixpoint for a prototype $P_1$ has already been computed, then the computation of the fixpoint for a prototype $P_2$ with $base$ $P_1$ will reuse this result. Similarly during the consistency check for recursive derivation from $\emptyProtoID$: if it is already known that a prototype $P_1$ derives recursively from $\emptyProtoID$, then we reuse this information to conclude that $P_2$ with base $P_1$ derives from $\emptyProtoID$. Next, we will introduce the data sets and the benchmarks in which they are used. \subsection{Data Sets and Benchmarks} \label{sec:datasets} Since the prototype system is new there are no existing real-world dataset which make use of its features. It would be possible to use existing RDF datasets as prototypes, but it would result in a KB which does not use the inheritance feature specific to prototypes. Therefore, we decided to use synthetic data sets for our benchmarks. We created three types of data sets in three different sizes, resulting in nine data sets altogether. Note that we do not use the remove capabilities in our data sets. This is not required since a remove will only reduce the burden on the system. An overview of the datasets can be found in \cref{tab:datasets}. \begin{table*}[t] \caption{An overview of the data sets. The numbers between brackets indicate the different size parameters used to generate the data sets (see \cref{sec:datasets} for more information). The table shows the amount of prototypes and the average number (and st. dev.) of properties in the add set of the prototypes. Below we will refer to the data sets with their initial letters only. } \label{tab:datasets} \centering \setlength{\tabcolsep}{4pt} \makebox[\textwidth][c]{ % \begin{tabular}{lcccccc} \toprule Data set & \multicolumn{3}{c}{prototypes} & \multicolumn{3}{c}{properties per prototype} \\ \cmidrule(r){2-4} \cmidrule(r){5-7} \textbf{ba}seline (19/20/21) &1,048,575 & 2,097,151 & 4,194,303 & 0 & 0 & 0 \\ \textbf{bl}ocks (10/20/30) &1,000,000 & 2,000,000 & 3,000,000 & 1 & 1 & 1\\ \textbf{inc}remental (1/2/3) & 1,000,000 & 2,000,000 & 3,000,000 & 2.0 $\pm$ 1.4 &2.0 $\pm$ 1.4 &2.0 $\pm$ 1.4 \\ \bottomrule \end{tabular} } \end{table*} The first type of data sets does have beneficial properties for consistency checking and fixpoint computation. Further it does not have any properties attached to the proptotypes. Hence, we will call these \emph{baseline} data sets. To generate the data we start with one prototype which derives from $\emptyProtoID$, next we create two prototypes which derive form the one, then we create four prototypes which derive from these two, and so on until we create $2^{n}$ prototypes which derive from $2^{n-1}$ (for $n \in \{19,20,21\}$). For the second type we change the set-up to be less ideal and introduce properties. We create $10$, $20$, and $30$ blocks of $100,000$ prototypes and hence this type will be called \emph{blocks}. All prototypes in each block derive from a randomly chosen prototype in a lower block. Then, each of the prototypes has a property with a value randomly chosen from the block below. In the lowest block, the base is always $\emptyProtoID$ and the value for the property is always the same fixed prototype. The third type of data sets, which we call \emph{incremental}, is more demanding for the fixpoint computation and consistency check. This time we add 1, 2, and 3 million prototypes to the KB, one at a time. Each prototype gets a randomly selected earlier created one as its base. Furthermore, each prototype gets between $0$ and $4$ properties chosen from $10$ distinct ones (with replacement). The value of each property is chosen randomly among the prototypes. \vspace{-1em} \subsubsection{Results:} For each data set we measure how long it takes to perform the consistency check and to compute the fixpoint of all prototypes (i.e., compute the whole interpretation). We also measure the final average number of properties per prototype. These results can be found in \cref{tab:experiments}. \begin{table*}[t] \caption{The outcomes of the benchmark. For each dataset the table shows how long the consistency check took to complete and the time needed for the computation of the fixpoint. The last three columns show the average number (and standard deviation) of properties after computation of the fixpoint.} \label{tab:experiments} \centering \setlength{\tabcolsep}{4pt} \makebox[\textwidth][c]{ % \begin{tabular}{lccccccccc} \toprule Data set & \multicolumn{3}{c}{consistency(ms)} & \multicolumn{3}{c}{fixpoint(ms)} &\multicolumn{3}{c}{prop. per prototype in fp.} \\ \cmidrule(r){2-4} \cmidrule(r){5-7} \cmidrule(r){8-10} \textbf{ba} (19/20/21) & 2,659 & 4,083 & 8,150 & 5,281 & 7,344 & 15,055 & 0 & 0 & 0\\ \textbf{bl} (10/20/30) & 3,517 & 6,195 & 9,278 & 12,740 & 27,367 & 50,003 & 5.5 $\pm$ 2.9 & 10.5 $\pm$ 5.8 & 15.5 $\pm$ 8.6\\ \textbf{inc} (1/2/3) & 4,580 &8,469 & 10,436 & 23,597 &57,151 &94,702 & 26.7 $\pm$ 9.0& 27.3 $\pm$ 9.1& 30.0 $\pm$ 9.6\\ \bottomrule \end{tabular} } \end{table*} As can be seen from the table, the consistency check scales linear with the number of prototypes in the system. In some cases it seems like the behavior is even sub-linear. This is likely caused by just-in-time compilation. For the fixpoint, the baseline dataset provides close to linear performance, which is expected. Again, larger sets seem to compute even faster (or rather, the small set is handled slower because the JIT compiler has not yet optimized the code). The blocks and incremental experiments also show the expected scalability. They do, however, not have a linear scaling because in the larger experiments the numer of properties per prototype in the fixpoints is larger. The results are obtained by running the code on a single `Intel(R) Xeon(R) E5-2670 @ 2.60GHz' core (using \emph{taskset}). To keep results comparable we allowed the JVM to use large amounts of memory. The memory was mainly used to store the fixpoint of the KB, something one would in practice rarely keep in memory. These results show that the prototype system can scale well, even to millions of prototypes. \section{Distributed Knowledge Bases} \label{sec:remote} \label{sec:distributed} Our goal is to create a KB which can be used on the web. Hence, it is not sufficient to show that our implementation can be used locally. In this section we present how we implemented the client-server version of our KB. In our implementation we use the ubiquitous HTTP protocol and well known data formats. We also illustrate that KBs can have explicit links to each other trough the \texttt{Link} header of the HTTP protocol. To show the knowledge sharing scenario in action we perform benchmarks in which we query the KB in a simulated web environment. \subsection{Prototype Serialization and Joining} To communicate the Prototypes between server and client we want to use a language which is platform independent. Furthermore, the serialization should be reasonably easy to parse using different modern programming languages. Despite the simple textual serialization already available in the software and demonstrated in the research paper~\cite{researchPaper}, we choose to implement JSON serialization for the client--server interaction. This would also enable more straightforward integration with Javascript clients (and servers) at a later point. The JSON serialization itself is straightforward. A prototype is converted to the following structure: \begin{lstlisting} {"id":"theID", "base":"baseID", "add":{"propA":["id1", ...], ...}, "rem":{"propB":["id3", ...], ...}, "remAll":{"propC", ...}} \end{lstlisting} \subsubsection{Joining of Prototype Definitions} is needed when retrieving the representation of a prototype from multiple sources. In general the approach to this problem is dependent on the data sources and the amount of trust the client has in them. Imagine that one queries data source A for information about Germany. One of the properties of Germany returned by A is that the country has 80M inhabitants. When service B is asked about the same prototype a population of 20M is claimed. Now, the client has a specification of the schema it wants the countries to fulfill. Concrete, it could be that a country can only have one number for the population count property. Hence, the application would choose for the amount from the most trusted source. Several implementations of joining the changesets of prototypes are provided. \subsection{Deployment On Web Architecture} To serve prototypes on the web we use the HTTP protocol. To get the (serialized form of) the prototype, one needs to send a GET request with the protoype ID as a query parameter with the name $p$. For example, if the server is located at \url{http://example.com/} and one wants to request the prototype \url{isbn:123-4-56-789012-3}, then the request URL will be \url{http://example.com?p=isb We also implemented a way to serve fixpoints of prototypes. Since we are using HTTP, we can also use the existing optimizations and caching startegies available. From the server perspective, we use gzip compression in case the client supports it. Further, the server indicates how long the prototype will remain unchanged using the Cache-Control header \cite{rfc7234}. Besides, the ETag header \cite{rfc7232} is used; if the client wants to use the prototype but the cache time has expired, then it only needs to check with the server whether the ETag has changed to know whether it is up-to-date. The server implementation uses an embedded Jetty server which can be configured as desired. For instance, it is possible to deploy the implemented handler using HTTPS or HTTP/2. The client side (\lstinline|RemoteKB|) supports content compression, implements the caching of prototypes, and uses the ETag information. Besides, HTTP connections are reused for multiple prototypes in order to avoid the TCP handshake. Further, the client (and server) can handle multiple requests concurrently. The \texttt{Link: rel="alternate"} HTTP header~\cite{rfc5988} is used to indicate that other providers might have more information about a given prototype. Concrete, if a client asks for a prototype from service provider A, A might tell that B might have a definition of the given prototype as well. An application can then, in turn, get the definition from service B. We refrain from defining the exact meaning of the link between the two prototypes and rely on the application to choose its interpretation. We could have chosen to use the semantics of \url{owl:sameAs}, but this has lead to confusion and misuse in the past (see also~\cite{halpin2010owl}). The server also supports the retrieval of multiple prototypes at once. However, in that mode ETags and alternates will not work. Furthermore, if any of the prototypes requested is not found, the server will return an error code. Having this functionality is reasonable since the payload of the responses is relatively small and hence the serving time is dominated by the round trip time. \subsection{Benchmarks} In order to get repeatable results in our benchmarks, we work in a simulated web environment. For the simulation of a web environment there are two essential components. First, there are delays and losses on the network and second limited transmission rates. To simulate the delays and losses we use the netem\footnote{\url{http://www.linuxfoundation.org/collaborate/workgroups/networking/netem}} kernel component available in the Linux kernel. We attempt to obtain realistic settings by performing prior experiments in which we measure these factors when connecting from Finland (Europe) to New York city (USA). This way we found an average round trip time of 184.7 ms (min: 184.4 ms, max: 185.1 ms, stdev: 0.3 ms). We did not get any packet loss during out experiments, but will anyway use a package loss of 0.04\% (i.e 4 in 10,000 packets got lost in transmission). We arbitrary capped the bandwidth to 1024 KBit/s using token bucket filtering. This speed is arguably on the lower side, but shows that the KB can also be used when only limited bandwidth is available. In this benchmark we use the blocks(30) data set as presented in \cref{sec:datasets} and perform benchmarks in two environments. In the first and second benchmark, the network speed is virtually unconstrained (both server and client are on the same host). The difference between these two experiments is that first we only allow one concurrent request, while in the second one we allow up to 100. In the third experiment client and server are both deployed on their own virtual machine with the network modifications described above, the client can make up to 100 requests concurrently. In the benchmarks the client selects random prototypes and requests the fixpoint from the server. The timing for different amounts of prototypes can be found from~\cref{tab:remote}. As can be seen, it is beneficial to perform multiple requests simultaneously. Further, the network delay has a major impact on the timings. It becomes clear that requesting a couple of thousand fixpoints from the server does not really put it under stress. This can be seen from the second benchmark where hundred requests are send at the same time: the time to get 10,000 fixpoints is only half a second longer than to get 1,000 fixpoints. For the remote case we further observe a linear relation between the time needed and number of prototypes and note that in our simulated web environment we are able to retrieve roughly 50 prototype fixpoints per second. \begin{table*}[t] \caption{The outcomes of the benchmark. For each setting the time needed to fetch the fixpoints is shown in function of the number of fixpoints requested. \emph{Single} is the benchmark with the server and the client on the same host and one simultaneous request. \emph{Multi} is the same settign with multiple requests. \emph{Web Multi} is the setting where the client and server are on separate host with a simulated web link between them. All timings are in milliseconds.} \label{tab:remote} \centering \setlength{\tabcolsep}{4pt} \makebox[\textwidth][c]{ % \begin{tabular}{rrrrrrrr} \toprule Amount & Single & Multi & Web Multi & Amount & Single & Multi & Web Multi \\ \cmidrule(r){1-4} \cmidrule(r){5-8} 1,000 & 2,653 & 1,567 & 27,303 & 6,000 & 8,780 & 1,919 & 117,295 \\ 2,000 & 2,889 & 1,190 & 37,715 & 7,000 & 10,181 & 2,124 & 131,140 \\ 3,000 & 4,513 & 1,348 & 56,367 & 8,000 & 11,613 & 2,085 & 149,882 \\ 4,000 & 6,042 & 1,381 & 74,976 & 9,000 & 13,493 & 2,227 & 168,415 \\ 5,000 & 6,559 & 1,460 & 93,496 & 10,000 & 14,426 & 2,062 & 187,379 \\ \bottomrule \end{tabular} } \end{table*} \section{Conclusions} In this paper we presented the software we developed to spearhead further work in the Prototype based knowledge representation. We reviewed parts of the implementation of the prototype knowledge base and showed the results of benchmarks for consistency checking, fixpoint computing, and remote knowledge base access. At the current stage we see this software mainly used in research settings. However, the benchmarks show that the software provides scalability needed for use in production environment. \section*{Acknowledgments} Stefan Decker would like to thank Pat Hayes, Eric Neumann, and Hong-Gee Kim for discussions about Prototypes and Knowledge Representation in general. Michael Cochez performed parts of this research at the Industrial Ontologies Group of the University of Jyväskylä (Finland) and at the Insight Centre for Data Analytics in Galway, Ireland. Furthermore, it has to be mentioned that the implementation of the software was greatly simplified by Google's Guava library, Apache HttpComponents, the Jetty webserver, Google GSON, JUnit, Apache Abdera, and the Apache Commons Math™ library. \bibliographystyle{splncs03}
e651e1de3845eb3028ae2a54ff5b77d421981454
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction } \paragraph{}In recent years, many authors have worked on fractional integral inequalities by using different fractional integral operator such as Riemann-Liouville, Hadamard, Saigo and Erdelyi-Kober, see \cite{A,BA,BE,C1,C2,C3,C4,C5,D1,KA,P1,YI}. In \cite{KI2} S. Kilinc and H. Yildirim establish new generalized k-fractional integral inequalities involving Gauss hypergeometric function related to Chebyshev functional. In \cite{C2,D2} authors gave the following fractional integral inequalities, using the Hadamard and Riemann-Liouville fractional integral for extended Chebyshev functional. \begin{theorem} Let $f$ and $g$ be two synchronous function on $[0,\infty[$, and $r,p,q:[0,\infty)\rightarrow [0,\infty)$. Then for all $t>0$, $\alpha>0$, we have \begin{equation} \begin{split} &2_{H}D_{1,t}^{-\alpha}r(t) \left[_{H}D_{1,t}^{-\alpha}p(t) _{H}D_{1,t}^{-\alpha}(qfg)(t)+ _{H}D_{1,t}^{-\alpha}q(t)_{H}D_{1,t}^{-\alpha}(pfg)(t)\right]+\\ &2 _{H}D_{1,t}^{-\alpha}p(t)_{H}D_{1,t}^{-\alpha}q(t)_{H}D_{1,t}^{-\alpha}(rfg)(t)\geq\\ &_{H}D_{1,t}^{-\alpha}r(t) \left[_{H}D_{1,t}^{-\alpha}(pf)(t)_{H}D_{1,t}^{-\alpha}(qg)(t)+_{H}D_{1,t}^{-\alpha}(qf)(t)_{H}D_{1,t}^{-\alpha}(pg)(t)\right]+\\ &_{H}D_{1,t}^{-\alpha}p(t)\left[_{H}D_{1,t}^{-\alpha}(rf)(t)_{H}D_{1,t}^{-\alpha}(qg)(t)+_{H}D_{1,t}^{-\alpha}(qf)(t)_{H}D_{1,t}^{-\alpha}(rg)(t)\right]+\\ &_{H}D_{1,t}^{-\alpha}q(t)\left[_{H}D_{1,t}^{-\alpha}(rf)(t)_{H}D_{1,t}^{-\alpha}(pg)(t)+_{H}D_{1,t}^{-\alpha}(pf)(t)_{H}D_{1,t}^{-\alpha}(rg)(t)\right] \end{split} \end{equation} \end{theorem} \begin{theorem} Let $f$ and $g$ be two synchronous function on $[0,\infty[$, and $r,p,q:[0,\infty)\rightarrow [0,\infty)$. Then for all $t>0$, $\alpha>0$, we have: \begin{equation} \begin{split} &_{H}D_{1,t}^{-\alpha}r(t)\times\\ & \left[_{H}D_{1,t}^{-\alpha}q(t) _{H}D_{1,t}^{-\beta}(pfg)(t)+2 _{H}D_{1,t}^{-\alpha}p(t)_{H}D_{1,t}^{-\beta}(qfg)(t)+_{H}D_{1,t}^{-\beta}q(t)_{H}D_{1,t}^{-\alpha}(pfg)(t)\right]\\ &+\left[_{H}D_{1,t}^{-\alpha}p(t)_{H}D_{1,t}^{-\beta}q(t)+_{H}D_{1,t}^{-\beta}p(t)_{H}D_{1,t}^{-\alpha}q(t)\right]_{H}D_{1,t}^{-\alpha}(rfg)(t)\geq\\ &_{H}D_{1,t}^{-\alpha}r(t) \left[_{H}D_{1,t}^{-\alpha}(pf)(t)_{H}D_{1,t}^{-\beta}(qg)(t)+_{H}D_{1,t}^{-\beta}(qf)(t)_{H}D_{1,t}^{-\alpha}(pg)(t)\right]+\\ &_{H}D_{1,t}^{-\alpha}p(t)\left[_{H}D_{1,t}^{-\alpha}(rf)(t)_{H}D_{1,t}^{-\beta}(qg)(t)+_{H}D_{1,t}^{-\beta}(qf)(t)_{H}D_{1,t}^{-\alpha}(rg)(t)\right]+\\ &_{H}D_{1,t}^{-\alpha}q(t)\left[_{H}D_{1,t}^{-\alpha}(rf)(t)_{H}D_{1,t}^{-\beta}(pg)(t)+_{H}D_{1,t}^{-\beta}(pf)(t)_{H}D_{1,t}^{-\alpha}(rg)(t)\right]. \end{split} \end{equation} \end{theorem} \paragraph{}The main objective of this paper is to establish some Chebyshev type inequalities and some other inequalities using generalized k-fractional integral operator. The paper has been organized as follows. In Section 2, we define basic definitions related to generalized k-fractional integral operator. In section 3, we obtain Chebyshev type inequalities using generalized k-fractional. In Section 4 , we prove some inequalities for positive continuous functions. \section{ Preliminaries} \paragraph{}In this section, we present some definitions which will be used later discussion. \begin{definition} Two function $f$ and $g$ are said to synchronous (asynchronous) on $[a,b],$ if \begin{equation} \left((f(u)-f(v))(g(u)-g(v))\right)\geq (\leq)0, \end{equation} for all $ u, v \in [0,\infty)$. \end{definition} \begin{definition} \cite{KI2,YI} The function $f(x)$, for all $x>0$ is said to be in the $L_{p,k}[0,\infty),$ if \begin{equation} L_{p,k}[0,\infty)=\left\{f: \|f\|_{L_{p,k}[0,\infty)}=\left(\int_{0}^{\infty}|f(x)|^{p}x^{k}dx\right)^{\frac{1}{p}} < \infty \, \, 1 \leq p < \infty \, k \geq 0\right\}, \end{equation} \end{definition} \begin{definition} \cite{KI2,SAO,YI} Let $f \in L_{1,k}[0,\infty),$. The generalized Riemann-Liouville fractional integral $I^{\alpha,k}f(x)$ of order $\alpha, k \geq 0$ is defined by \begin{equation} I^{\alpha,k}f(x)= \frac{(k+1)^{1-\alpha}}{\Gamma (\alpha)}\int_{0}^{x}(x^{k+1}-t^{k+1})^{\alpha-1}t^{k} f(t)dt. \end{equation} \end{definition} \begin{definition} \cite{KI2,YI} Let $k\geq0,\alpha>0 \mu >-1$ and $\beta, \eta \in R $. The generalized k-fractional integral $I^{\alpha,\beta,\eta,\mu}_{t,k}$ (in terms of the Gauss hypergeometric function)of order $\alpha$ for real-valued continuous function $f(t)$ is defined by \begin{equation} \begin{split} I^{\alpha,\beta,\eta,\mu}_{t,k}[f(t)]& = \frac{(k+1)^{\mu+\beta+1}t^{(k+1)(-\alpha-\beta-2\mu)}}{\Gamma (\alpha)}\int_{0}^{t}\tau^{(k+1)\mu}(t^{k+1}-\tau^{k+1})^{\alpha-1} \times \\ & _{2}F_{1} (\alpha+ \beta+\mu, -\eta; \alpha; 1-(\frac{\tau}{t})^{k+1})\tau^{k} f(\tau)d\tau. \end{split} \end{equation} \end{definition} where, the function $_{2}F_{1}(-)$ in the right-hand side of (2.4) is the Gaussian hypergeometric function defined by \begin{equation} _{2}F_{1} (a, b; c; t)=\sum_{n=0}^{\infty}\frac{(a)_{n}(b)_{n}}{(c)_{n}} \frac{t^{n}}{n!}, \end{equation} and $(a)_{n}$ is the Pochhammer symbol\\ $$(a)_{n}=a(a+1)...(a+n-1)=\frac{\Gamma(a+n)}{\Gamma(a)}, \,\,\,(a)_{0}=1.$$ Consider the function \begin{equation} \begin{split} F(t,\tau)&= \frac{(k+1)^{\mu+\beta+1}t^{(k+1)(-\alpha-\beta-2\mu)}}{\Gamma (\alpha)}\tau^{(k+1)\mu}\\ &(t^{k+1}-\tau^{k+1})^{\alpha-1} \times _{2}F_{1} (\alpha+ \beta+\mu, -\eta; \alpha; 1-(\frac{\tau}{t})^{k+1})\\ &=\sum_{n=0}^{\infty}\frac{(\alpha+\beta+\mu)_{n}(-n)_{n}}{\Gamma(\alpha+n)n!}t^{(k+1)(-\alpha-\beta-2\mu-\eta)}\tau^{(k+1)\mu}(t^{k+1}-\tau^{k+1})^{\alpha-1+n}(k+1)^{\mu+\beta+1}\\ &=\frac{\tau^{(k+1)\mu}(t^{k+1}-\tau^{k+1})^{\alpha-1}(k+1)^{\mu+\beta+1}}{t^{k+1}(\alpha+\beta+2\mu)\Gamma(\alpha)}+\\ &\frac{\tau^{(k+1)\mu}(t^{k+1}-\tau^{k+1})^{\alpha}(k+1)^{\mu+\beta+1}(\alpha+\beta+\mu)(-n)}{t^{k+1}(\alpha+\beta+2\mu+1)\Gamma(\alpha+1)}+\\ &\frac{\tau^{(k+1)\mu}(t^{k+1}-\tau^{k+1})^{\alpha+1}(k+1)^{\mu+\beta+1}(\alpha+\beta+\mu)(\alpha+\beta+\mu+1)(-n)(-n+1)}{t^{k+1}(\alpha+\beta+2\mu+1)\Gamma(\alpha+2)2!}+... \end{split} \end{equation} It is clear that $F(t,\tau)$ is positive because for all $\tau \in (0, t)$ , $(t>0)$ since each term of the (2.6) is positive. \section{Fractional Integral Inequalities for Extended Chebyshev Functional} In this section, we establish some Chebyshev type fractional integral inequalities by using the generalized k-fractional integral (in terms of the Gauss hypergeometric function) operator. The following lemma is used for the our main result. \begin{lemma} Let $f$ and $g$ be two synchronous function on $[0,\infty[,$ and $x,y:[0,\infty)\rightarrow$ $[0,\infty)$ be two nonnegative functions. Then for all $k \geq 0,$ $t>0$, $\alpha > max\{0,-\beta-\mu\}$, $\beta < 1,$ $\mu >-1,$ $\beta -1< \eta <0,$ we have, \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}x(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(yfg)(t)+ I^{\alpha,\beta,\eta,\mu}_{t,k}y(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(yg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(yf)(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} \end{lemma} \textbf{Proof}: Since $f$ and $g$ are synchronous on $[0,\infty[$ for all $\tau \geq 0$, $\rho\geq 0$, we have \begin{equation} (f(\tau)-f(\rho)) (g(\tau)-g(\rho))\geq 0. \end{equation} From (3.2), \begin{equation} f(\tau)g(\tau)+f(\rho)g(\rho)\geq f(\tau)g(\rho)+f(\rho)g(\tau). \end{equation} \noindent Now, multiplying both side of (3.3) by $ \tau^{k}x(\tau)F(t,\tau)$, $\tau \in (0,t)$, $t>0$. Then the integrating resulting identity with respect to $\tau$ from $0$ to $t$, we obtain by definition (2.4) \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)+f(\rho)g(\rho) I^{\alpha,\beta,\eta,\mu}_{t,k}(x)(t)\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(yg)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)+f(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} \noindent Now, multiplying both side of (3.4) by $ \rho^{k}y(\rho)F(t,\rho)$, $\rho \in (0,t)$, $t>0$, where $F(t,\rho)$ defined in view of (2.6). Then the integrating resulting identity with respect to $\rho$ from $0$ to $t$, we obtain by definition (2.4) \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}y(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)+ I^{\alpha,\beta,\eta,\mu}_{t,k}(yfg)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(x)(t)\\ &\geq g(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(yf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} This complete the proof of (3.1)\\ \noindent Now, we gave our main result here. \begin{theorem} Let $f$ and $g$ be two synchronous function on $[0,\infty[$, and $r,p,q:[0,\infty)\rightarrow [0,\infty)$. Then for all $k \geq 0,$ $t>0$, $\alpha > max\{0,-\beta-\mu\}$, $\beta < 1,$ $\mu >-1,$ $\beta -1< \eta <0,$ we have, \begin{equation} \begin{split} &2I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) \left[I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qfg)(t)+ I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)\right]+\\ &2 I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\geq\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) \left[I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t)\right]+\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]+\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right] \end{split} \end{equation} \end{theorem} \textbf{Proof}: To prove above theorem, putting $x=p, \ y=q$, and using lemma 3.1, we get \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qfg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t). \end{split} \end{equation} \noindent Now, multiplying both side by (3.7) $I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)$, we have \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}p(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(qfg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)\right]\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t)\right], \end{split} \end{equation} \noindent putting $x=r, y=q$, and using lemma 3.1, we get \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(qfg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t), \end{split} \end{equation} multiplying both side by (3.9) $I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)$, we have \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(qfg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t) \right]\geq\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(qg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]. \end{split} \end{equation} With the same arguments as before, we can write \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\right]\geq\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]. \end{split} \end{equation} Adding the inequalities (3.8), (3.10) and (3.11), we get required inequality (3.6).\\ Here, we give the lemma which is useful to prove our second main result. \begin{lemma} Let $f$ and $g$ be two synchronous function on $[0,\infty[$. and $x,y:[0,\infty[\rightarrow$ $[0,\infty[$. Then for all $k \geq 0,$ $t>0$, $\alpha > max\{0,-\beta-\mu\}$,$\gamma> max\{0,-\delta-\upsilon\}$ $\beta,\delta < 1,$ $\upsilon,\mu >-1,$ $\beta -1< \eta <0,$ $\delta-1<\zeta <0,$ we have, \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}x(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfg)(t)+ I^{\gamma,\delta,\zeta,\upsilon}_{t,k}y(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yf)(t) I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} \end{lemma} \textbf{Proof}: \noindent Now multiplying both side of (3.4) by \begin{equation} \begin{split} &\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}y(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} \end{split} \end{equation} which remains positive in view of the condition stated in (3.12), $\rho \in (0,t)$, $t>0$, we obtain \begin{equation} \begin{split} &\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}y(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)\\ &+\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}y(\rho)f(\rho)g(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} I^{\alpha,\beta,\eta,\mu}_{t,k}x(t)\geq \\ &\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}y(\rho)g(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}y(\rho)f(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t), \end{split} \end{equation} \noindent then integrating (3.14) over (0,t), we obtain \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}y(t)+ I^{\alpha,\beta,\eta,\mu}_{t,k}(x)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfg)(t)\\ &\geq I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}yg(t) +I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}yf(t), \end{split} \end{equation} \noindent this ends the proof of inequality (3.12). \begin{theorem} Let $f$ and $g$ be two synchronous function on $[0,\infty[$, and $r,p,q:[0,\infty)\rightarrow [0,\infty)$. Then for all $t>0$, $\alpha>0$, we have: \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)\times\\ & \left[I^{\alpha,\beta,\eta,\mu}_{t,k}q(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(pfg)(t)+2 I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qfg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)\right]\\ &+\left[I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}q(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}p(t)I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\right]I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\geq\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) \left[I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t)\right]+\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]+\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(pg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]. \end{split} \end{equation} \end{theorem} \textbf{Proof}: To prove above theorem, putting $x=p, \ y=q$, and using lemma 3.3 we get \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qfg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t). \end{split} \end{equation} \noindent Now, multiplying both side by (3.17) $I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)$, we have \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}p(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qfg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pfg)(t)\right]\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(pf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(pg)(t)\right], \end{split} \end{equation} \noindent putting $x=r, \ y=q$, and using lemma 3.3, we get \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qfg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t), \end{split} \end{equation} multiplying both side by (3.19) $I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)$, we have \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}r(t) I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qfg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}q(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\geq \right]\\ &I^{\alpha,\beta,\eta,\mu}_{t,k}p(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(qf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]. \end{split} \end{equation} With the same argument as before, we obtain \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}r(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(pfg)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}p(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rfg)(t)\right]\geq \\ &I^{\alpha,\beta,\eta,\mu}_{t,k}q(t)\left[I^{\alpha,\beta,\eta,\mu}_{t,k}(rf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(pg)(t)+(pf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(rg)(t)\right]. \end{split} \end{equation} Adding the inequalities (3.18), (3.20) and (3.21), we follows the inequality (3.16). \begin{remark} If $ f,g,r,p \ and\ q $ satisfies the following condition, \begin{enumerate} \item The function f and g is asynchronous on $[0,\infty)$. \item The function r,p,q are negative on $[0,\infty)$. \item Two of the function r,p,q are positive and the third is negative on $[0,\infty)$. \end{enumerate} then the inequality 3.6 and 3.16 are reversed. \end{remark} \section{Other fractional integral inequalities} In this section, we proved some fractional integral inequalities for positive and continuous functions which as follows: \begin{theorem} Suppose that $f$, $g$ and $h$ be three positive and continuous functions on $[0,\infty[$, such that \begin{equation} (f(\tau)-f(\rho))(g(\tau)-g(\tau))(h(\tau)+h(\rho))\geq 0; \ \tau, \rho \in(0,t)\ \ t>0, \end{equation} and $x$ be a nonnegative function on $[0,\infty)$. Then for all $k \geq 0,$ $t>0$, $\alpha > max\{0,-\beta-\mu\}$,$\gamma> max\{0,-\delta-\upsilon\}$ $\beta,\delta < 1,$ $\upsilon,\mu >-1,$ $\beta -1< \eta <0,$ $\delta-1<\zeta <0,$ we have, \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(x)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xfgh)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xfg)(t)\\ &+I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xh)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xfgh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(x)(t)\\ & \geq I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xgh)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xfh)(t)\\ &+I^{\alpha,\beta,\eta,\mu}_{t,k}(xgh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xf)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xfh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xg)(t). \end{split} \end{equation} \end{theorem} \textbf{Proof}: Since $f$, $g$ and $h$ be three positive and continuous functions on $[0,\infty[$ by (4.1), we can write \begin{equation} \begin{split} &f(\tau)g(\tau)h(\tau)+f(\rho)g(\rho)h(\rho)+f(\tau)g(\tau)h(\rho)+f(\rho)g(\rho)h(\tau)\\ &\geq f(\tau)g(\rho)h(\tau)+f(\tau)g(\rho)h(\rho)+f(\rho)g(\tau)h(\tau)+f(\rho)g(\tau)h(\rho). \end{split} \end{equation} \noindent Now, multiplying both side of (4.3) by $ \tau^{k}x(\tau)F(t,\tau)$, $\tau \in (0,t)$, $t>0$. Then the integrating resulting identity with respect to $\tau$ from $0$ to $t$, we obtain by definition (2.4) \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(xfgh)(t)+f(\rho)g(\rho)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}x(t)+g(\tau)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+f(\rho)g(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xh)(t)\geq g(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xfh)(t)+g(\rho)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+f(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xgh)(t)+f(\rho)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} \noindent Now multiplying both side of (4.4) by \begin{equation} \begin{split} &\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}x(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} \end{split} \end{equation} which remains positive in view of the condition stated in (4.2), $\rho \in (0,t)$, $t>0$ and integrating resulting identity with respective $\rho $ from $0$ to $t$, we obtain \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(xfgh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}x(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xfgh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}x(t)\\ &+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xgf)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xfg)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xh)(t)\\ &\geq I^{\gamma,\delta,\zeta,\upsilon}_{t,k}xg(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xfh)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xgh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xgh)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(xfh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t).\end{split} \end{equation} which implies the proof inequality 4.2.\\ Here, we give another inequality which is as follows. \begin{theorem}Let $f$, $g$ and $h$ be three positive and continuous functions on $[0,\infty[$, which satisfying the condition (4.1) and $x$ and $y$ be two nonnegative functions on $[0,\infty)$. Then for all $k \geq 0,$ $t>0$, $\alpha > max\{0,-\beta-\mu\}$,$\gamma> max\{0,-\delta-\upsilon\}$ $\beta,\delta < 1,$ $\upsilon,\mu >-1,$ $\beta -1< \eta <0,$ $\delta-1<\zeta <0,$ we have, \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(x)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfgh)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfg)(t)\\ &+I^{\alpha,\beta,\eta,\mu}_{t,k}(xfg)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yh)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xfgh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}y(t)\\ & \geq I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(ygh)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfh)(t)\\ &+I^{\alpha,\beta,\eta,\mu}_{t,k}(xgh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yf)(t)+I^{\alpha,\beta,\eta,\mu}_{t,k}(xfh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yg)(t). \end{split} \end{equation} \end{theorem} \textbf{Proof}: \noindent Multiplying both side of (4.3) by $ \tau^{k}x(\tau)F(t,\tau)$, $\tau \in (0,t)$, $t>0$, where $F(t,\tau)$ defined by (2.6). Then the integrating resulting identity with respect to $\tau$ from $0$ to $t$, we obtain by definition (2.4) \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(xfgh)(t)+f(\rho)g(\rho)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}x(t)+g(\tau)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+f(\rho)g(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xh)(t)\geq g(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xfh)(t)+g(\rho)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+f(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xgh)(t)+f(\rho)h(\rho)I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} \noindent Now multiplying both side of (4.8) by \begin{equation} \begin{split} &\frac{(k+1)^{\upsilon+\delta+1}t^{(k+1)(-\delta-\gamma-2\upsilon)}}{\Gamma (\gamma)}\rho^{(k+1)\upsilon}y(\rho)\\ &(t^{k+1}-\rho^{k+1})^{\gamma-1} \times _{2}F_{1} (\gamma+ \delta+\upsilon, -\zeta; \gamma; 1-(\frac{\rho}{t})^{k+1})\rho^{k} \end{split} \end{equation} which remains positive in view of the condition stated in (4.7), $\rho \in (0,t)$, $t>0$ and integrating resulting identity with respective $\rho $ from $0$ to $t$, we obtain \begin{equation} \begin{split} &I^{\alpha,\beta,\eta,\mu}_{t,k}(xfgh)(t)I^{\gamma,\delta,\zeta,\upsilon}_{t,k}y(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfgh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}x(t)\\ &+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xgf)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfg)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xh)(t)\\ &\geq I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yg)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xfh)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(ygh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xf)(t)\\ &+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yf)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xgh)(t)+I^{\gamma,\delta,\zeta,\upsilon}_{t,k}(yfh)(t)I^{\alpha,\beta,\eta,\mu}_{t,k}(xg)(t). \end{split} \end{equation} which implies the proof inequality 4.7.
c9944f1b2673dde626d6bbc38729eb0f97e35fa6
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Advancement in the current technology of trapping and controlling single atoms cooled down to ultra low temperatures has opened new research directions in quantum optics and quantum communication. Spatial configurations of linear atomic chains or two-dimensional atomic lattices have been engineered and have been widely applied in various experimental setups~\cite{coldgas,ryd1,ryd2}. Recently, the subject of utilizing super-cold atoms as highly reflecting mirrors has gained much attention. In particular, it has been demonstrated that a collection of cold atoms trapped near the surface of a one-dimensional waveguide can form a nearly perfect mirror for the radiation incident on the atoms~\cite{QST,cj12,Chang}. The waveguide represents a photonic channel which enhances the electromagnetic field to which the atoms are coupled thereby leading to a strong collective behavior of the atoms. As a consequence, a large part of the incident light is directed, reflected back, to the medium from which it originated. This striking mirror property of atoms is in contrary to the usual observation where atoms absorb/scatter all or most of the incident energy. Another kind of systems that can exhibit mirror properties or equivalently a highly directive radiative properties are atoms chirally coupled to a waveguide~\cite{pr15,vr16}. Chirality in atom-waveguide coupling is an effect associated with a broken symmetry of emission of photons from the atoms into the right and left propagating modes of the waveguide. As a result, the emitted photons are channeled into one of the two directions of the waveguide. It has been shown that the chiral property of the emission can enhance entanglement between two distant atoms~\cite{gg15}. Directive radiative properties have also been demonstrated for a single atom trapped at front of a distant dielectric mirror~\cite{hs11}. It has been demonstrated both theoretically and experimentally that the atom can behave as an optical mirror effectively forming, together with the dielectric mirror, a Fabry-P\'erot cavity. Related studies have shown that An atom mirror cannot only serve as a single mirror for a one-dimensional cavity, but also could be arranged to behave like a high-finesse cavity~\cite{supercavity}. In the course of previous work on directional emission the underlying atoms independently couple to a one-dimensional field of a waveguide, or nano-cavity or a nano-fibre. Although systems involving independent atoms exhibit interesting directional properties, there can be similar features created by an open system of atoms in which the atoms are coupled to a common three-dimensional field. It was Dicke~\cite{d54} who pointed out that a collection of a large number of atoms coupled to a common EM field can radiate collectively such that the spontaneous radiation can be enhanced in certain directions. Since then there have been many studies of the collective radiative properties of multi-atom system demonstrating the dependence of the emitted radiation on the number of atoms and the geometry of the emitting system~\cite{Lehmberg,r82,bb84,f86,corr1,corr2,corr3,corr4,Carmichael,mf07,Cirac,Wstate,lz14}. In this paper, we investigate radiative properties of an open system of a line of few atoms and demonstrate that the dipole-dipole interaction between atoms may lead to a collective ordering of the emission along some specific directions. To determine directions of the emission, we introduce the directivity function of the emitted radiation field and study the dependence of the function on the distance between the atoms. We analyze the directional properties of the radiation field for two different configurations of the atomic chains, one mimicking an atom in front of a mirror and the other an atom inside a cavity. In the first, we choose the left-hand-side atom of the chain as a probe atom and examine conditions under which the system may radiate only to those modes whose propagation vectors lie within a small solid angle about the inter atomic axis and oriented in one of the two directions of the inter atomic axis, either right (forward) or left (backward) direction. Such a system can be regarded as an atom in front of perfectly reflecting or perfectly transmitting atomic mirror. In the second arrangement, we choose the middle atom of the chain as a probe atom and examine conditions under which the system may radiate only to those modes whose propagation vectors lie within a small solid angle about the inter atomic axis. Such a scheme can be regarded as an atom inside a single-mode cavity. We find that there is a threshold for the inter atomic distances above which a highly directional emission could be achieved. Below the threshold the emission is spherically symmetric. However, we find a population trapping in one of the mirror atoms. For the cavity-like arrangement, the directivity function depends strongly on the number of atoms contained in the chain and the distance between them. For a $3$-atom chain and atomic distances above the threshold, two radiative modes of different spatial directions are available for the emission, one in the direction parallel and the other in the direction normal to the inter-atomic axis. Below the threshold, the system can radiate only to the mode normal to the inter atomic axis. For a $5$-atom chain and distances above the threshold, only one mode is available for emission, either normal or parallel to the atomic line. Thus, there exist ranges of the inter atomic distances under which the atomic chain exhibits features characteristic of a single-mode cavity. The paper is organized as follows. In Sec.~\ref{sec2}, we describe the master equation of the density operator of the system and the mathematical approach used in the evaluation of the density matrix elements. We introduce the definitions of the directivity function, reflection and transmission coefficients of the radiation field emitted by a chain of atoms. In Sec.~\ref{sec3}, we examine the conditions for the mirror-type behavior of short chains composed of $3$ and $5$ atoms. We observe the transient transfer of the population between the atoms and the transient directivity function for an initial condition in which the probe atom is prepared in its excited state. Then, we examine the directivity function of the stationary field when the probe atom is driven by a continuous wave (cw) laser field. Section~\ref{sec4} is devoted to the problem of a cavity formation with atomic mirrors. We are particularly interested in the possibility of the system to concentrate the radiation along the inter atomic axis and thus to behave as a single-mode cavity. Polar diagrams are given to illustrate the mirror and cavity-like features of atomic chains and to show how the features are sensitive to distances between the atoms. The results are summarized in Sec.~\ref{sec5}. The paper concludes with an Appendix6A in which we give details of the derivation of the atomic correlation functions in terms of the populations of the collective states of a three-atom system and the coherences between them and Appendix6B where the calculation of equations of motion has been presented. \section{Radiative properties of a chain of atoms}\label{sec2} We consider a system composed of $N$ identical two-level atoms located at fixed positions $\vec{r}_{i}$ and coupled to the three-dimensional electromagnetic field whose modes are initially in a vacuum state $\ket{\{0\}}$. Each atom has an excited state $\ket{e_{i}}$ and a ground state $\ket{g_{i}}$ separated by energy~$\hbar\omega_{0}$ and connected by a transition dipole moment~$\vec{\mu}$. \begin{figure}[t] \centering \includegraphics[width=0.8\columnwidth]{qrfig1} \caption{(Color online) Two different arrangements of atoms in a line to demonstrate that a chain of closely located and interacting atoms can act as an atomic mirror or cavity. (a) The left-side atom of the chain, specially chosen as a "probe" atom, is located at distance $r_{0}$ from its nearest neighbor with the remaining atoms equally separated from each other by a distance $r_{m}< r_{0}$. (b) The middle atom of the chain, chosen as a probe atom, is separated from its next-neighbors by $r_{0}$, while the remaining atoms are equally separated by a distance~$r_{m}< r_{0}$. } \label{fig1} \end{figure} The atoms are arranged in a line, and we consider two cases shown in~Fig.~\ref{fig1}. In the first case, illustrated in Fig.~\ref{fig1}(a), we assume that the left-side atom, chosen as a ``probe" atom is separated from its next-neighbor by a distance $r_{0}$ which is larger than the separation $r_{m}$ between the remaining $N-1$ atoms, $r_{0}> r_{m}$. If the probe atom is excited into its upper level, it will spontaneously decay into the ground state emitting the radiation field that can be absorbed by the chain of closely located atoms and then re-emitted by the atoms towards the probe atom. Thus, the chain of closely located atoms could act as a mirror, directing the emitted radiation into a cone about the interatomic axis and turned towards the probe atom. In the second case, illustrated in Fig.~\ref{fig1}(b), we assume that the middle atom of the chain is separated from its adjacent neighbors by a distance $r_{0}$, which is much larger that the separation $r_{m}$ between the remaining atoms. This arrangement may model a situation of an atom located inside a cavity whose mirrors are formed by two chains of equally distant atoms. In practice this scheme could be realized by extending the recently demonstrated scheme involving two superconducting qubits coupled to a 1D field~\cite{mlynek, nissen,loo} to the case of three or five qubits coupled to a 2D field. In the experiment of Ref.~\cite{loo}, effective separations of $\lambda$ and $3\lambda/4$ were achieved between the fixed qubits by changing the qubit transition frequencies. \subsection{Master equation} When the system is coupled to a reservoir the state of the total system, the chain of atoms plus the reservoir field, is described by the density operator $\rho_{T}$. The reduced density operator describing the properties of only the chain of atoms is obtained by tracing the total density operator $\rho_{T}$ over the states of the reservoir, $\rho ={\rm Tr}_{R}\rho_{T}$. The master equation describing the time evolution of the reduced density operator has the form~\cite{agarwal,fs,ficek-book} \begin{align} \frac{\partial \rho}{\partial t} &= -\frac{i}{\hbar}\left[H_{0}+H_{L}+H_{dd},\rho\right] \nonumber\\ &-\frac{1}{2}\sum_{i=1}^N\gamma \left( \left[S_{i}^{+},S_{i}^{-}\rho\right] + \textrm{H.c.}\right)\nonumber\\ &-\frac{1}{2}\sum_{i\neq j =1}^N\gamma_{ij} \left( \left[S_{i}^{+},S_{j}^{-}\rho\right] + \textrm{H.c.}\right) ,\label{q1} \end{align} where $\gamma$ is the spontaneous emission damping rate of the individual atoms, equal to the Einstein $A$ coefficient, $S_{i}^{+}=\ket{e_{i}}\bra{g_{i}}$ and $S_{i}^{-}=\ket{g_{i}}\bra{e_{i}}$ are the dipole raising and lowering operators of atom $i$, and $\gamma_{ij}$ is the collective damping rate \begin{align} \gamma_{ij} &= \frac{3}{2}\gamma\left\{\left[1 -\left(\hat{\mu}\cdot\hat{r}_{ij}\right)^{2}\right]\frac{\sin\eta_{ij}}{\eta_{ij}}\right. \nonumber\\ &\left. + \left[1-3\left(\hat{\mu}\cdot\hat{r}_{ij}\right)^{2}\right]\left(\frac{\cos\eta_{ij}}{\eta_{ij}^2}-\frac{\sin\eta_{ij}}{\eta_{ij}^3}\right)\right\} , \label{q2} \end{align} with \begin{align} \eta_{ij} = k\,r_{ij}=2\pi r_{ij}/\lambda ,\quad \vec{r}_{ij} = r_{ij}\hat{r}_{ij}=\vec{r}_{j} -\vec{r}_{i} , \end{align} in which $r_{ij}$ is the distance between atoms $i$ and $j$, $\hat{r}_{ij}$ is the unit vector in the direction $\vec{r}_{ij}$, and $\lambda$ is the resonant wavelength. The master equation (\ref{q1}) describes the atomic dynamics under the Born-Markov and rotating-wave approximations~\cite{scullybook}, $H_{0}$ is the Hamiltonian describing the free energy of the atoms \begin{align} H_{0} = \hbar\sum_{i=1}^{N}\omega_{0}S^{+}_{i}S_{i}^{-} ,\label{q3} \end{align} $H_{L}$ is the Hamiltonian describing the interaction of the probe atom with an external driving field of frequency $\omega_{L}$: \begin{align} H_{L} = \frac{1}{2}\hbar\Omega_{0}\left(S_{1}^{+}e^{-i\omega_{L}t} + S_{1}^{-}e^{i\omega_{L}t}\right) ,\label{q3a} \end{align} where $\Omega_{0}$ is the Rabi frequency of the driving field, and $H_{dd}$ is the Hamiltonian describing the dipole-dipole interaction between the atoms \begin{align} H_{dd} = \hbar\sum_{i\neq j=1}^{N}\Omega_{ij}\left(S_{i}^{+}S_{j}^{-} + S_{j}^{+}S_{i}^{-}\right) ,\label{q4} \end{align} where $\Omega_{ij}$ is the dipole-dipole interaction strength between atoms $i$ and $j$, defined by \begin{align} \Omega_{ij} &= \frac{3}{4}\gamma\left\{\left[1-3\left(\hat{\mu}\cdot\hat{r}_{ij}\right)^{2}\right]\left(\frac{\sin\eta_{ij}}{\eta_{ij}^2}+\frac{\cos\eta_{ij}}{\eta_{ij}^3}\right)\right. \nonumber\\ &\left. -\left[1-\left(\hat{\mu}\cdot\hat{r}_{ij}\right)^{2}\right]\frac{\cos\eta_{ij}}{\eta_{ij}}\right\} .\label{q5} \end{align} The parameters $\gamma_{ij}$ and $\Omega_{ij}$ depend on the separation between the atoms. For large separations, $\eta_{ij}\gg 1$, and then both coupling parameters approach zero. For $\eta_{ij}\ll 1$ the parameter $\gamma_{ij}$ reduces to $\gamma$ while $\Omega_{ij}$ becomes large and strongly dependent on $r_{ij}$. It is well known that $\Omega_{ij}$ plays the important role in the collective behavior of multi-atom systems and we shall see that it has important effect on the distribution of the radiation field emitted by a chain of atoms. The calculation of the equations of motion for atomic populations and coherences for a time-dependent state vector has been outlined briefly in~Appendix6B. \subsection{Directivity function, reflection and transmission coefficients}\label{sec2a} The intensity of the radiation field emitted at time~$t$ in the direction specified by the polar angle $\theta$ between the direction of observation $\vec{R}$ and the direction of the atomic axis $\vec{r}_{ij}$ can be expressed in terms of the correlation functions of the atomic dipole operators as \begin{align} I(\theta,t) = u(\phi)\sum_{i,j=1}^{N}\gamma \langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle e^{ikr_{ij}\cos\theta} ,\label{q6} \end{align} where $u(\phi)=(3/8\pi)\sin^{2}\phi$ is the radiation pattern of a single atomic dipole, with $\phi$ the angle between the observation direction $\vec{R}$ and the direction of the atomic transition dipole moment $\vec{\mu}_{i}$. Since the radiation intensity $I(\theta,t)$ is symmetric about the interatomic axis, it defines a two-dimensional surface called the polar radiation pattern of the emitting system. We may introduce the {\it directivity} function determining of how effective the atoms are in converging the emitted radiation into a small solid angle centered about the direction $\theta$. The directivity function $D(\theta,t)$ in the direction $\theta$ at time $t$ is defined as the ratio of the radiation intensity $I(\theta,t)$ emitted in the direction $\theta$ divided by the total radiation intensity $I(t)$: \begin{align} D(\theta, t) = \frac{u(\phi)}{I(t)}\sum_{i,j=1}^{N}\gamma \langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle e^{ikr_{ij}\cos\theta} ,\label{q9} \end{align} where the total radiation intensity $I(t)$ at time $t$ is obtained by integrating $I(\theta,t)$ over $\theta$: \begin{align} I(t) = \int I(\theta,t)d\theta = \sum_{i,j=1}^{N}\gamma_{ij}\langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle ,\label{q10} \end{align} in which $\gamma_{ii}=\gamma_{jj}=\gamma$ and $\gamma_{ij}\, (i\neq j)$ is given in Eq.~(\ref{q2}). The directivity function is a measure of how effective the system is in concentrating the radiation in a given direction. It is equivalent to the probability density of detecting a fluorescence photon traveling in the direction $\theta$. Since our interest is in situations where a chain of atoms works as an atomic mirror, an important factor is the ability of producing highly directional patterns of the radiation concentrated in one of the two directions along the interatomic axis, either $\theta=0$ or $\theta=\pi$. Following the arrangement illustrated in Fig.~\ref{fig1}(a), $D(\theta=\pi,t)$ describes the radiation field emitted along the atomic axis in the direction towards the probe atom, the "backward" direction. Thus, it would correspond to the reflection coefficient. On the other hand, the directivity $D(\theta=0,t)$ describes the radiation field emitted along the atomic axis in the direction away from the probe atom, the ``forward" direction. Therefore, it would correspond to the transmission coefficient of the atomic mirror. Thus, we may define the {\it reflection} coefficient of the chain of atoms as the ratio of the radiation intensity emitted in the direction $\theta= \pi$ to the total radiation intensity \begin{align} R(t) \equiv D(\theta\!=\!\pi, t) = \frac{u(\phi)}{I(t)}\!\sum_{i,j=1}^{N}\!\gamma \langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle e^{ikr_{ij}} .\label{q12} \end{align} The reflection coefficient is a measure of how effective the atoms are in concentrating the radiation about one side of the interatomic axis, i.e. about the direction $\theta =\pi$. The coefficient $R(t)$ is equivalent to the probability density of detecting a fluorescence photon traveling in the direction~$\theta =\pi$. Similarly, we can define the {\it transmission} coefficient of the chain as the ratio of the radiation intensity emitted in the direction $\theta=0$ to the total intensity of the field emitted \begin{align} T(t) \equiv D(\theta\!=\!0,t) = \frac{u(\phi)}{I(t)}\!\sum_{i,j=1}^{N}\!\gamma \langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle e^{-ikr_{ij}} .\label{q11} \end{align} Obviously, $T(t)=1$ would correspond to complete transmission whereas $R(t)=1$ would correspond to complete reflection of the radiation field emitted along the atomic axis. \subsection{Directional properties of the radiation field}\label{sec2b} The quantity of central interest is the directivity function which can be determined from analyzing the polar radiation pattern of a chain of atoms. We may determine general conditions under which the radiation pattern of $N$ atoms would be highly non spherical and its maximum is concentrated along the inter atomic axis. The conclusions will serve as reference for choosing distances between the atoms and for calculations of the atomic populations and correlations. The expression (\ref{q6}) can be written as a sum of $N$ terms \begin{align} I(\theta,t) = \sum_{i<j=1}^{N}I_{ij}(\theta,t) ,\label{q7} \end{align} where \begin{align} I_{ij}(\theta,t) &= u(\phi)\gamma \left\{\frac{1}{N\!-\!1}\!\left(\langle S_{i}^{+}(t)S_{i}^{-}(t)\rangle\!+\!\langle S_{j}^{+}(t)S_{j}^{-}(t)\rangle\right)\right. \nonumber\\ &\left. +\, 2{\rm Re}\{\langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle\}\cos\left(kr_{ij}\cos\theta\right)\right. \nonumber\\ &\left. -\, 2{\rm Im}\{\langle S_{i}^{+}(t)S_{j}^{-}(t)\rangle\}\sin\left(kr_{ij}\cos\theta\right)\right\} .\label{q8} \end{align} We see that the contribution of the atoms to the intensity occurs in pairs of different combinations of the atoms. Therefore, the system of radiating atoms can be considered as made up of a number of short two-atom elements and the total intensity is obtained by summing up the intensities of the fields produced by all the elements. If we wish a short chain of atoms to work like a mirror with a large convergence and reflectivity of the radiation emitted by a probe atom located at either end of the chain, we should arrange the atoms such that the total field emitted (scattered) could be highly focused along the interatomic axis with a pronounced maximum in the backward direction $\theta = \pi$ and a minimum, preferably zero emission in the forward direction $\theta = 0$ with respect to the line center. In order to find the conditions for concentrating the radiation in the direction $\theta= \pi$, let us examine the intensity (\ref{q8}) in more details. From Eq.~(\ref{q8}) it is seen that there are three terms determining the radiation pattern. The first term in Eq.~(\ref{q8}) is just the sum of the populations of the two atoms involved, the probabilities that the atoms are in their excited states. This term is independent of $\theta$ and therefore contributes uniformly in all directions. The second term depends on $\theta$ and varies as $\cos\left(kr_{ij}\cos\theta\right)$ with an amplitude equal to the real part of atomic correlations. Consequently, this term could contribute to the radiation pattern only if the correlations between the atoms would have nonzero real part, ${\rm Re}\{\langle S_{i}^{+}S_{j}^{-}\rangle\}\neq 0$. However, the cosine term, although dependent on $\theta$, would produce intensity maxima in the $\theta =0, \pi/2, 3\pi/2$ and $\pi$ directions. As such this term is not effective in concentrating the radiation along the interatomic axis. This simple argument suggests that the atoms should be arranged such that the cosine term vanishes. This could be achieved when the correlations between the atoms have zero real part. The third term contributing to the radiation intensity~(\ref{q8}) varies as $\sin\left(kr_{ij}\cos\theta\right)$ and hence can affect the radiation intensity in a decidedly different way than the cosine term. An important difference is that $\sin\left(kr_{ij}\cos\theta\right)$ vanishes for $\theta =\pi/2$ and $3\pi/2$. This means that the sine term does not contribute to the radiation emitted in the direction perpendicular to the interatomic axis. Consequently, an appreciable concentration of the radiation could be achieved along the interatomic axis by choosing proper distances $r_{ij}$ between the atoms at which $\sin\left(kr_{ij}\right)=\pm 1$. Furthermore, since the sine is an odd function, it follows that $\sin\left(kr_{ij}\cos 0^{\circ}\right)= -\sin\left(kr_{ij}\cos 180^{\circ}\right)$, which means that independent of the separation between the atoms a maximum in the backward direction is always accompanied by a minimum in the forward direction. Obviously, the sine term could influence the angular distribution of the radiation intensity only if the atomic correlations would have nonzero imaginary parts, ${\rm Im}\{\langle S_{i}^{+}S_{j}^{-}\rangle\}\neq 0$. In addition, the sign of the imaginary part of the atomic correlations dictates the choice of distances between the atoms, at which the emission would be enhanced in the backward direction (high reflection) and reduced in the forward direction (low transmission). Thus, if ${\rm Im}\{\langle S_{i}^{+}S_{j}^{-}\rangle\} > 0$, the sine term will display a maximum in the backward direction for atomic distances at which $\sin(kr_{ij})=1$. This condition is satisfied when the atomic separations are \begin{align} r_{ij} = \frac{1}{4}(2n+1)\lambda ,\quad n\in\{0,2,4,\ldots \} \end{align} However, if the coefficient of the sine term is negative, ${\rm Im}\{\langle S_{i}^{+}S_{j}^{-}\rangle\} < 0$, a different choice of the distances is required at which $\sin(kr_{ij})=-1$. This condition is satisfied when \begin{align} r_{ij} = \frac{3}{4}(2n+1)\lambda ,\quad n\in\{0,2,4,\ldots \} \end{align} We see that there is a lower bound imposed on the distances between the atoms, either $r_{ij}=\lambda/4$ or $r_{ij}=3\lambda/4$, above which a one-sided emission along the interatomic axis can be achieved, i.e., a maximum of radiation in the direction $\theta = 0$. For distances smaller that the lower bound, the one-sided emission is expected to be significantly reduced. In the following, we limit ourselves to short chains containing only $N=3$ and $N=5$ atoms. Also in the absence of the driving field, $\Omega_{0}=0$, the $N=3$ case can be solved in closed form yielding simple expressions. In the~Appendix6A, we outline the calculation of atomic correlation functions in the collective state basis. \subsection{Mathematical approach} To study the radiative behavior of a chain of interacting atoms, we require the time evolution of the populations of the atoms and the coherence between them. These are given by the diagonal and off-diagonal density matrix elements, respectively. If the space of the atomic system is spanned in the basis of the eigenstates of the free Hamiltonian $H_{0}$, we readily find that the basis is composed of $2^{N}$ state vectors, i.e. for a chain composed of $N=3$ atoms, the basis is composed of $8$ vectors $\ket{i_{1}j_{2}k_{3}}$, whereas for $N=5$ atoms it is composed of $64$ vectors $\ket{i_{1}j_{2}k_{3}l_{4}m_{5}},\, \{i,j,k,l,m\} \in\{ g,e\}$. Hence, in the simplest case of $N=3$ atoms, the master equation (\ref{q1}) provides us with a system of $64$ coupled linear equations to be solved, in principle, a $63\times 63$ matrix to be diagonalized. For $N=5$ we get a system of $4096$ coupled linear equations whose solution requires a $4095\times 4095$ matrix to be diagonalized. Needless to say, it is not easily accomplished. Therefore, we shall use numerical methods. To get solutions for the density matrix elements, the following approach will be taken. From the master equation (\ref{q1}), we find equations of motion for the density matrix elements, which can be written in a matrix form as \begin{equation} \dot{\vec{X}}\left( t\right) = {\cal A} \vec{X}\left( t\right) +\vec{R}, \label{q17} \end{equation} where $\vec{X}\left( t\right) $ is a column vector composed of the density matrix elements, $\vec{R}$ is a column vector composed of the inhomogenous terms, and ${\cal A}$ is a matrix of the coefficients appearing in the equations of motion of the density matrix elements. A direct integration of Eq.~(\ref{q17}) leads to the following formal solution for $\vec{X}\left( t\right)$ \begin{align} \vec{X}\left( t\right) =\vec{X}\left( t_{0}\right) e^{{\cal A}t}-\left(1-e^{{\cal A}t}\right) {\cal A}^{-1}\vec{R}, \label{q18} \end{align} where $t_{0}$ is the initial time. Because the determinant of the matrix ${\cal A}$ is different from zero, there exists a complex invertible matrix ${\cal U}$ which diagonalizes ${\cal A}$, and $w = {\cal U}^{-1}{\cal A}{\cal U}$ is the diagonal matrix of complex eigenvalues. By introducing new vectors $\vec{Y}={\cal U}^{-1}\vec{X}$ and $\vec{T}={\cal U}^{-1}\vec{R}$, we can rewrite (\ref{q18}) as \begin{equation} \vec{Y}\left( t\right) =\vec{Y}\left( t_{0}\right) e^{w t}-\left(1-e^{w t}\right) w^{-1}\vec{T} , \label{q19} \end{equation} or in component form \begin{equation} Y_{n}\left( t\right) =Y_{n}\left( t_{0}\right) e^{w_{n}t}-\sum_{m=1}^{s}\!\left( w^{-1}\right)_{nm}\!\left( 1-e^{w_{m}t}\right) T_{m} , \label{q20} \end{equation} in which $s=63$ for the case of $N=3$ atoms and $s=4095$ for $N=5$ atoms. To obtain solutions for $X_{n}\left( t\right) $ we must determine the eigenvalues $w_{n}$ and eigenvectors $Y_{n}\left( t\right) ,$ which are readily evaluated by a numerical diagonalization of the matrix ${\cal A}$. The steady-state values of the density matrix elements can be found from Eq.~(\ref{q20}) by taking $t\rightarrow \infty $, or more directly by setting the left-hand side of Eq.~(\ref{q17}) equal to zero, and then \begin{equation} \vec{X}\left( \infty \right) =-{\cal A}^{-1}\vec{R}, \label{q21} \end{equation} or in component form \begin{equation} X_{n}\left( \infty \right) =-\sum_{m=1}^{s}\left( {\cal A}^{-1}\right)_{nm}R_{m} . \label{q22} \end{equation} In what follows, we shall use the solutions (\ref{q20}) and (\ref{q22}) to illustrate the radiative properties of the atomic chains. In particular, we calculate the transient populations of the atoms and the steady-state directivity function. \section{Radiating atom at front of an atomic mirror}\label{sec3} Consider first the configuration illustrated in Fig.~\ref{fig1}(a), a single atom distance $r_{12}$ from a finite-size chain of closely located atoms. This arrangement could constitute a radiating atom at front of an atomic mirror. In order to find the mirror-like characteristics of the chain that could be inferred from the radiative properties of the systems, we examine the directivity function of the radiation field emitted by the system. The directivity function, which is determined by the angular distribution of the radiation intensity, Eq.~(\ref{q7}), depends on temporal and spatial factors, the time of the evolution of the system and the separation between the atoms. It is clear from Eq.~(\ref{q7}) that in general the temporal and spatial factors cannot be separated. The angular distribution of the radiation field at a given time can be different for different separations between the atoms. Moreover, the transient radiation intensity is a sensitive function of the initial atomic conditions. Even at the initial time $t=0$ the angular distribution of the radiation intensity can depend on the separation $r_{ij}$ if the system was prepared in a state with nonzero interatomic correlations, either ${\rm Re}\{\langle S_{i}^{+}S_{j}^{-}\rangle\}$ or ${\rm Im}\{\langle S_{i}^{+}S_{j}^{-}\rangle\}$ different from zero. In order to study the angular distribution of the radiation field, it is important to understand the radiative behavior of individual atoms in the chain. Therefore, we first consider the evolution of the populations of the atoms. Following this discussion, we display the variation of the directivity function with time for different distances between the atoms. We assume that the probe atom is initially excited and separated from the front atom of the ``mirror", its nearest neighbor, at a distance much larger than the separation between the ``mirror" atoms. The method of preparing the probe atom in the excited state will not concern us but it might be done with a short laser pulse, for example. \subsection{Transient transfer of the population}\label{sec3a} Our interest is in the time evolution of the directivity function of the emitted field for an initially excited probe atom, the case illustrated in Fig.~\ref{fig1}(a). In order to study this evolution, we first look at the time evolution of the populations of the atoms for chains composed of $N=3$ and $N=5$ atoms. The aim is to first determine and then optimize the conditions which not only allow the radiation emitted by the probe atom to be emitted back (reflected) to it but also to achieve a strong convergence of the radiated field along the interatomic axis. We evaluate the transient populations of the atoms assuming there is no driving field $(\Omega_{0}=0)$ and that initially the probe atom was in its excited state. The corresponding results for the time evolution of the atomic populations for different distances between the atoms are presented in~Figs.~\ref{fig2} and~\ref{fig3}. \begin{figure}[h] \includegraphics[width=4.55cm]{fig2a} \hspace{-0.4cm} \includegraphics[width=4.1cm]{fig2b} \caption{(Color online) Time evolution of the atomic populations $\rho_{11}(t)$ (dashed blue line), $\rho_{22}(t)$ (solid red line), $\rho_{33}(t)$ (dotted green line) plotted for the case of an initially excited probe atom $1$ located in front of a line of two atoms, $2$ and $3$, and several different sets of distances between the atoms: (a) $(r_{12}, r_{23}) =(\lambda/2,\lambda/4)$, (b) $(r_{12}, r_{23}) =(\lambda/4,\lambda/6)$, (c) $(r_{12}, r_{23}) =(\lambda/3,\lambda/5)$, and (d) $(r_{12}, r_{23}) =(\lambda/8,\lambda/9)$.} \label{fig2} \end{figure} Figure~\ref{fig2} shows the transient transfer of the population between $N=3$ atoms; the initially excited probe atom and a group of two atoms forming a mirror. The populations are computed for several different sets of distances $(r_{12}, r_{23})$ between the atoms. For distances $r_{ij}>\lambda/5$ the initial population of the probe atom decays exponentially in time whereas the populations of the mirror atoms $2$ and $3$ build up with small oscillations. However, there is no tendency of the mirror atoms to transfer their populations back to the probe atom. A small oscillatory behavior of the population $\rho_{11}(t)$ can be seen for short times, $t<5/\gamma$, but it is not periodic. A periodic oscillatory behavior is observed for the populations of atoms $2$ and $3$ that the mirror atoms periodically exchange their populations without transferring it to the probe atom. The reason is that the distance $r_{23}$ is much smaller than $r_{12}$, resulting in a stronger dipole-dipole interaction between the mirror atoms, $\Omega_{23}\gg \Omega_{12}$. For smaller distances, the population of the probe atom begins to show periodic oscillations, with the population actually oscillating between the probe atom and only the rear atom of the mirror. Except for very short times, the population of the middle atom $2$ remains almost constant during the evolution of the system. Note that the transfer of the population between the atoms is not complete since the populations remain nonzero, $\rho_{ii}(t)\neq 0$ for all $t > 0$. It is interesting that atom $2$, the front atom of the mirror, appears as an mediator in the population exchange between the probe atom and the rear atom of the mirror. In other words, the front atom of the mirror dictates the direction of the emission. Despite of the possibility to emit in any spatial direction, the probe atom ``prefers" to radiate towards the atomic mirror and vice versa, the mirror atoms prefer to radiate towards the probe atom. \begin{figure}[h] \includegraphics[width=4.38cm]{fig3a} \includegraphics[width=4.1cm]{fig3b} \caption{(Color online) Time evolution of the atomic populations, $\rho_{11}$ (cyan solid line), $\rho_{22}$ (black dashed-dotted line), $\rho_{33}$ (red dotted line), $\rho_{44}$ (green solid line), and $\rho_{55}$ (blue dashed line) plotted for the case of an initially excited probe atom $1$ located in front of a line of four atoms and several different sets of distances between the atoms: (a) $(r_{12}, r_{i,i+1}) = (\lambda/2,\lambda/4)$, (b) $(r_{12}, r_{i,i+1}) =(\lambda/4,\lambda/6)$, (c) $(r_{12}, r_{i,i+1}) =(\lambda/3,\lambda/5)$, and (d) $(r_{12}, r_{i,i+1}) = (\lambda/8,\lambda/9)$.} \label{fig3} \end{figure} We now consider a chain containing $N=5$ atoms with the probe atom initially excited and located at a distance $r_{12}$ from the front of a line of four equally separated atoms. The transient transfer of the population between the atoms for several different sets of distances, $(r_{12}, r_{i,i+1}), \, i \in \{2,3,4 \}$ is shown in Fig.~\ref{fig3}. For large distances the situation is similar to that of three atoms, the population rapidly escapes from the system leaving the atoms unpopulated over a short evolution time. For small distances, we observe periodic oscillations of the populations with the periodicity, as before for three atoms, determined by the dipole-dipole interaction strength. However, unlike the three atoms case, the population can be completely transferred from the mirror atoms back to the probe atom such that the mirror atoms are unpopulated at times when the population of the probe atom is maximal. This is a substantial difference compared to the case of $N=3$ atoms where a part of the population was trapped by one of the mirror atoms. Clearly, longer chains are more effective in the complete transfer of an excitation from the probe atom to the mirror atoms and vice versa. \subsection{Transient directivity function}\label{sec3b} The periodic exchange of the populations between the atoms, in particular, between the atoms forming an atomic mirror, can lead to the emission of the atoms into certain preferred directions. In order to demonstrate this behavior, we examine the time variation of the directivity function $D(\theta,t)$ of the field radiated by the group of mirror atoms only. Of course it might be argued that it could be hard for experiments to detect the field radiated from a fraction of atoms, but we would like to see under which conditions the field radiated by the group of mirror atoms could be concentrated in the preferred direction along the interatomic axis, either $\theta=0$ or $\theta =\pi$. With a high concentration of the radiated field into a small solid angle centered around $\theta=0$, the atoms could be regarded as a highly reflecting mirror scattering the field back towards the probe atom. \begin{figure}[h] \includegraphics[width=4.27cm]{fig4a} \includegraphics[width=4.27cm]{fig4b} \includegraphics[width=4.27cm]{fig4c} \includegraphics[width=4.27cm]{fig4d} \caption{(Color online) Polar diagram of the directivity function $D(\theta,t)$ describing the concentration of the field emitted by two atoms forming a mirror when an initially excited probe atom is at distance $r_{12}=\lambda/2$ from the front of the mirror atoms. Distances between the mirror atoms are: in (a) $r_{23}=\lambda/3$, (b) $r_{23}=\lambda/4$, (c) $r_{23}=\lambda/5$, and in (d) $r_{23}=\lambda/8$. Green (black) curve corresponds to a minimum (maximum) in $\rho_{11}(t)$. The time instants for the green and black curves are: (a) $t\in\{3\gamma^{-1}, 4\gamma^{-1} \}$, (b) $t\in\{5\gamma^{-1}, 3.3\gamma^{-1} \}$, (c) $t\in\{9\gamma^{-1}, 3.6\gamma^{-1} \}$, (d) $t\in\{9.4\gamma^{-1}, 6\gamma^{-1} \}$, respectively. } \label{fig4} \end{figure} The directionality function of the field emitted by two (mirror) atoms and detected on a sphere around the line of three atoms is shown in Fig.~\ref{fig4}. We keep the distance of the probe atom from the front of the mirror atoms at $r_{12}=\lambda/2$ and consider the variation of $D(\theta,t)$ with the distance between the mirror atoms at two different times. It is seen that the emission is concentrated along the interatomic axis. For a distance between the mirror atoms $r_{23}=\lambda/4$, the directivity function is significantly enhanced in one direction, that the emission is almost one-sided either in $\theta=0$ or $\theta=\pi$. The directivity function oscillates in time in such a manner that it develops a maximum in the direction $\theta =0$ at later times in which it was a minimum at earlier times and vice versa. The oscillation of the directivity results from the oscillation of the population between the mirror atoms, as we have see above in Fig.~\ref{fig2}. The beam width or equivalently the solid angle inside which the emitted field is concentrated remains almost constant. \begin{figure}[h] \includegraphics[width=4.28cm]{fig5a} \includegraphics[width=4.28cm]{fig5b} \includegraphics[width=4.28cm]{fig5c} \includegraphics[width=4.2cm]{fig5d} \caption{(Color online) Polar diagram of the directivity function $D(\theta,t)$ describing the concentration of the field emitted by four atoms forming a mirror when an initially excited probe atom is at distance $r_{12}=\lambda/2$ from the front of the mirror atoms. Distances between the mirror atoms are: in (a) $r_{i,i+1}=\lambda/3$, (b) $r_{i,i+1}=\lambda/4$, (c) $r_{i,i+1}=\lambda/5$, and in (d) $r_{i,i+1}=\lambda/8$. Green (black) curve corresponds to a minimum (maximum) in $\rho_{11}(t)$. The time instants for the green and black curves are: (a) $t\in\{7.76\gamma^{-1}, 3.7\gamma^{-1} \}$, (b) $t\in\{6.8\gamma^{-1}, 3.4\gamma^{-1} \}$, (c) $t\in\{7\gamma^{-1}, 3.715\gamma^{-1} \}$, (d) $t\in\{7.6\gamma^{-1}, 9.1\gamma^{-1} \}$, respectively. } \label{fig5} \end{figure} One can also notice from Fig.~\ref{fig4} that the ability of the coupled atoms to concentrate the radiation in one direction decreases with a decreasing distance between the atoms. It is clearly seen that for distances $r_{23}\leq \lambda/8$ the directivity function is almost symmetrical along the interatomic axis. This result is consistent with the general property of the radiation pattern discussed in Sec.~\ref{sec2a} that there is a minimal value of the distance between the atoms $(r_{ij}=\lambda/4)$ above which the sine term can be maximal in the $\theta=0$ direction. For distances smaller than the minimal value the contribution of the sine term is necessarily smaller resulting in the reduction of the ability of the system to concentrate the emission in one direction. The one-sided emission and then the reflectivity coefficient can be enhanced by increasing the number of atoms forming the atomic mirror. This is demonstrated in Fig.~\ref{fig5} which shows the polar diagram of the directivity function $D(\theta,t)$ for the case of four atoms forming the atomic mirror. As above for three atoms, we keep the probe atom at a fixed distance from the front of the atomic mirror, $r_{12}=\lambda/2$, and consider the variation of $D(\theta,t)$ with the distance between the mirror atoms. It can be seen that similar to the case of two atoms the emission is concentrated mainly along the interatomic axis. The one-sided emission is significantly enhanced and persists even for small distances. Small peaks can be seen at about $2\pi/3$ and $4\pi/3$ degrees. It is easy to verify that the function $\sin\left(kr_{ij}\cos\theta\right)$ when evaluated for $r_{25}=3\lambda/8$, corresponding to the distance between the front and rear atoms of the mirror, attains its maximal value of $\sin\left(kr_{ij}\cos\theta\right)= 1$ for $\cos\theta= 2/3$, which corresponds to directions $\theta \approx 130^{\circ}$. \subsection{Directivity function of the stationary field}\label{sec3c} In practice a photo detector located at some position $\vec{R}$ in the far field zone of the radiation field emitted by the atoms would detect the field emitted by the entire set of atoms rather than a fraction of selected atoms only. It is a consequence of the fact that the fields from the probe atom and the mirror atoms are unresolved at the detector. Therefore, we now consider the directivity function of the field radiated by the entire system of atoms including the field radiated from the probe atom. Moreover, we assume that initially all atoms were in their ground states and then the probe atom was exposed to the incident weak laser light. We shall look into the radiation pattern of the emitted field in the steady-state limit, $t\rightarrow \infty$, and compute the directivity function $D(\theta) \equiv \lim_{t\rightarrow\infty}D(\theta, t)$ for equally distant as well as for non-equally distant $N=3$ and $N=5$ atoms. Since the directivity function involves the contribution from the populations of the atoms, that a significant directionality can be obtained when the populations are small, we shall assume that the driving field is weak so that the Rabi frequency $\Omega$ much smaller than $\gamma$. \begin{center} \begin{figure}[h] \includegraphics[width=4.27cm]{fig6a} \includegraphics[width=4.27cm]{fig6b} \includegraphics[width=4.27cm]{fig6c} \includegraphics[width=4.27cm]{fig6d} \caption{(Color online) Polar diagram of the directivity function $D(\theta)$ of the stationary field emitted by three atoms in a line. The left-side atom of the line, which constitutes a probe atom is driven by a cw laser field of the Rabi frequency $\Omega_{0}=0.01\gamma$. In frame (a) $r_{12}=\lambda/2, r_{23}=\lambda/4$. In frame (b) $r_{12}=\lambda/4, r_{23}=\lambda/8$. In frame (c) $r_{12}= r_{23} = \lambda/4$, and in frame (d) $r_{12}=\lambda/4, r_{23}=\lambda/6$.\\ } \label{fig6} \end{figure} \end{center} The directivity function for the radiation field emitted by $N=3$ atoms is shown in Fig.~\ref{fig6}, where frame (a) is for equally distant atoms, while frame (b) is for unequally distant atoms. Unfortunately, as one can see from the figure, in both cases that the stationary field radiated by the system is concentrated in directions $\theta=\pi/3$ and $\theta = 5\pi/3$, which significantly departure from the direction of the interatomic axis. Therefore, the system of three equally or unequally separated atoms with the continuously driven probe atom cannot be regarded as suitable for the mirror-type behavior of the maximal directivity along the interatomic axis. \begin{figure}[h] \includegraphics[width=4.27cm]{fig7a} \includegraphics[width=4.27cm]{fig7b} \includegraphics[width=4.27cm]{fig7c} \includegraphics[width=4.27cm]{fig7d} \caption{(Color online) Polar diagram of the directivity function $D(\theta)$ of the stationary field radiated by a line of $5$ atoms. The left-side atom of the line, which constitutes a probe atom is driven by a cw laser field of the Rabi frequency $\Omega_{0}=0.01\gamma$. In frame (a), the atoms are unequally separated with $r_{12}=\lambda/2$, and $r_{23}=r_{34}=r_{45}=\lambda/4$. In frame (b), the atoms are unequally separated with $r_{12}=\lambda/4$, and $r_{23}=r_{34}=r_{45}=\lambda/8$. In frame (c), the atoms are equally separated with $r_{12}= r_{23} = r_{34}=r_{45}= \lambda/4$, and in frame (d), the atoms are unequally separated with $r_{12}=\lambda/4$, and $r_{23}=r_{34}=r_{45}=\lambda/6$.} \label{fig7} \end{figure} Figure~\ref{fig7} shows the corresponding situation for the case of $N=5$ atoms. It is seen that in the case of four atoms forming the mirror the directivity function depends crucially on the separation between the atoms. In particular when the distance of the probe atom from the front of the mirror atoms is $r_{12}=\lambda/4$, the radiated field modes available for the emission are only those contained inside a cone centered about the direction $\theta =\pi$. Thus, at that particular distance the chain of four undriven atoms acts as a perfectly reflecting mirror by ``pushing" the radiation in the backward direction. It can be seen that the optimal conditions for the one-sided emission with a maximum in the direction $\theta =\pi$ are achieved when the probe atom is separated from the front of the mirror atoms by $r_{12}=\lambda/4$ and the mirror atoms are themselves separated by $r_{ij}=\lambda/6$. In this case, the directivity function is nonzero only in directions contained inside a cone limited by $\theta\leq \pm\pi/3$. Thus, the emission is entirely one-sided with no radiation in the $\theta=0$ direction. This means that the transmission coefficient $T=0$, therefore this type of behavior can be regarded as a mirror type with perfect reflectivity. Of course, the reflectivity is accompanied by losses in the sense that the radiation is emitted into a cone of a finite solid angle $2\theta$ determining the beamwidth of the radiated field. The solid angle subtended by the cone seen in Fig.~\ref{fig7}(c) is $2\theta = 120^{\circ}$. It was pointed out in~\cite{Carmichael} that a strong directivity of the emitted radiation along the interatomic axis can be obtained in a chain of independent atoms, i.e. in the absence of the dipole-dipole interaction. However, the one sided emission seen in frames (c) and (d) of Fig.~\ref{fig7} requires a non-zero dipole-dipole interaction. This is illustrated in Fig.~\ref{fig7a} which shows the directivity function for the same parameters as in Fig.~\ref{fig7}(d) but for independent atoms. Clearly, in the absence of the dipole-dipole interaction between the atoms, the emission is strongly directional, but is symmetrical in the $\theta=0$ and $\pi$ directions. \begin{figure}[h] \includegraphics[width=4.27cm]{fig7dn} \caption{(Color online) Polar diagram of the directivity function $D(\theta)$ of the stationary field radiated by a line of $5$ independent atoms $(\Omega_{ij}=0)$. The atoms are unequally separated with $r_{12}=\lambda/4$, and $r_{23}=r_{34}=r_{45}=\lambda/6$.} \label{fig7a} \end{figure} In summary of this section, we have seen that the directivity of the emission and the mirror like behavior of a line of atoms depends on the distance between the atoms and the nature of the excitation. At particular distances between the atoms, the system can radiate along the interatomic axis in only one direction $\theta=\pi$, which can be interpreted as a perfect reflection of the radiation towards the probe atom. \section{Cavity formation with atomic mirrors}\label{sec4} We now wish to create a cavity with atomic mirrors. For this purpose, we place a probe atom between a pair of finite-size chains of closely located atoms, as illustrated in Fig.~\ref{fig1}(b). These would constitute a radiating atom located inside a one-dimensional cavity. We determine the parameter ranges in which the two chains of $(N-1)/2$ atoms could act as mirrors to the field emitted by the probe atom. We examine two systems containing different numbers of atoms. First, we consider the simplest system composed of $N=3$ equidistant atoms in a line. Then, we extend our discussion to a larger system composed of $N=5$ atoms. We illustrate our considerations by examining the time evolution of the atomic populations. For the initial conditions we choose the middle (probe) atom to be in its excited state and the other atoms to be in their ground states. We also analyze the directivity function for features indicative of cavity-type modifications of the radiation pattern of the radiation field. \subsection{Transient regime}\label{sec4a} The simplest system which could exhibit features characterizing a cavity formed by atomic mirrors is a chain of $N=3$ equidistant atoms. Let us first examine the process of population transfer between the middle (probe) atom of the chain and the side (mirror) atoms. The time evolution of the populations of the atoms is illustrated in~Fig.~\ref{fig8}, where we present the dependence of the population transfer on the distance between the atoms. It is seen that the population is periodically transferred between the probe atom and the mirror atoms. The transfer occurs with frequency determined by the dipole-dipole coupling $\Omega_{12}\, (=\Omega_{23})$ between the probe and mirror atoms. The oscillations show an interesting behavior that the population transfer is not affected (modulated) by the presence of the dipole-dipole coupling $\Omega_{13}$ between the mirror atoms. \begin{figure}[t] \includegraphics[width=4.4cm]{fig8a} \includegraphics[width=4.1cm]{fig8b} \caption{(Color online) Time evolution of the atomic populations, $\rho_{22}$ (red solid line), $\rho_{11}$ (blue dashed line), $\rho_{33}$ (green dotted line), plotted for the cavity-type configuration of $3$ atoms and for different distances between the atoms: (a) $r_{12}=r_{23}=\lambda/4$, (b) $r_{12}=r_{23}=\lambda/6$, (c) $r_{12}=r_{23}=\lambda/8$, and (d) $r_{12}=r_{23}=\lambda/16$. Initially, the middle (probe) atom $2$ was excited, $\ket{\psi(0)} =\ket 2$. } \label{fig8} \end{figure} The oscillations are accompanied by a steady decay of the populations. However, depending on the distance between the atoms, the populations may not decay to zero but rather to long-lived non-zero values. It is clearly seen from the figure that for atomic distances $r_{ij}<\lambda/4$, i.e. the spacing between the mirror atoms $r_{13}<\lambda/2$, a significant part of the initial population remains trapped in the probe atom from which it decays very slowly. We have thus a situation similar to that one encounters in a Fabry-P\'erot cavity composed of a pair of parallel plate mirrors~\cite{dowling}. In the cavity, the number of modes for interaction with an atom decreases with an decreasing separation $L$ between the mirrors. For $L<\lambda/2$ the number of modes is suppressed resulting in the suppression of the radiation from the atom. \begin{figure}[t] \includegraphics[width=4.4cm]{fig9a} \includegraphics[width=4.15cm]{fig9b} \caption{(Color online) Time evolution of the atomic populations plotted for the cavity-type configuration of a chain of $5$ atoms and different distances between the atoms: (a) $r_{12}=r_{45}=\lambda/2$, $r_{23}=r_{34}=\lambda$, (b) $r_{12}=r_{45}=\lambda/4$, $r_{23}=r_{34}=\lambda/2$, (c) $r_{12}=r_{45}=\lambda/6$, $r_{23}=r_{34}=\lambda/3$, and (d) $r_{12}=r_{45}=\lambda/16$, $r_{23}=r_{34}=\lambda/8$. Initially, the middle (probe) atom $3$ was excited, $\ket{\psi(0)} =\ket 3$. Color and style specifications as in Fig.~\ref{fig3}. } \label{fig9} \end{figure} Figure~\ref{fig9} illustrates the cavity-like situation involving $5$ atoms. Here, each of the cavity mirrors is formed with two atoms. The results are similar to those we encountered for the case of $3$ atoms, Fig.~\ref{fig8}, however, one can see some interesting differences. Again, trapping of the population for small distances between the atoms is evident. It is worth noting that the system has a tendency to trap the population in the mirror atoms rather than in the probe atom. It can also be noted that for large distances the probe atom exchanges the population with the next neighbors, the front atoms rather than with the rear atoms forming the mirrors. However, for small distances, the probe atom exchanges the population with the rear atoms leaving the populations of the front atoms almost constant in time. Similar effects appeared to the population transfer in the system composed of a probe atom located at front of an atomic mirror, see Sec.~\ref{sec3a}. \subsection{Stationary regime}\label{sec4b} We now turn to the problem of determining the conditions under which the two chains of atoms could act like cavity mirrors to select modes centered about the atomic axis as the only modes available for the emission. We assume that the probe atom is continuously driven by a coherent laser field and consider the field in the steady state limit. \begin{figure}[t] \includegraphics[width=4.26cm]{fig10a} \includegraphics[width=4.26cm]{fig10b} \includegraphics[width=4.26cm]{fig10c} \includegraphics[width=4.26cm]{fig10d} \caption{(Color online) Polar diagram of the directivity function $D(\theta)$ of the stationary field in the cavity-like situation involving $3$ atoms. The middle atom, which constitutes a probe atom is driven by a cw laser field of the Rabi frequency $\Omega_{0}=0.01\gamma$. The atoms are separated by (a) $r_{ij}=\lambda/2$, (b) $r_{ij}=\lambda/4$, (c) $r_{ij}=\lambda/5$, and (d) $r_{ij}=\lambda/6$.} \label{fig10} \end{figure} Figure~\ref{fig10} illustrates the effect of decreasing distance between atoms on the directivity function $D(\theta)$ of the stationary field for the cavity-like situation involving $3$ atoms. It is seen that the directivity function is very sensitive to the distances between the atoms. For large distances, the system radiates along the cavity axis as well as in the direction normal to the cavity axis. When the atomic separations are reduced below $\lambda/4$, which corresponds to the mirror atoms spacing less than $\lambda/2$, we see a cancellation of the radiation along the interatomic axis. Thus, the system turns off the radiation along the cavity axis. It radiates only in the directions normal to the cavity axis, the property which clearly is not characteristic of a cavity. \begin{figure}[th] \includegraphics[width=4.22cm]{fig11a} \includegraphics[width=4.3cm]{fig11b} \includegraphics[width=4.24cm]{fig11c} \includegraphics[width=4.24cm]{fig11d} \caption{(Color online) Polar diagram of the directivity function $D(\theta)$ of the stationary field in the cavity-like situation involving $5$ atoms. The middle atom, which constitutes a probe atom is driven by a cw laser field of the Rabi frequency $\Omega_{0}=0.01\gamma$. In frame (a) the probe atom is separated from its next neighbors by $r_{23}=r_{34}=\lambda/2$, while the atoms which constitute the cavity mirrors are separated by $r_{12} = r_{45}=\lambda/4$. In frame (b) $r_{23} = r_{34}=\lambda/4$ and $r_{12} = r_{45}=\lambda/8$. In frame (c) $r_{23} = r_{34}=\lambda/4$ and $r_{12} = r_{45}=\lambda/10$. In frame (d) $r_{23} = r_{34}=\lambda/8$ and $r_{12} = r_{45}=\lambda/8$.} \label{fig11} \end{figure} Turning next to the case of $5$ atoms, we plot in Fig.~\ref{fig11} the directivity function for several different distances between the atoms. We see that the directivity function differs significantly from what we observed for the case of three atoms. Thus, the effect of additional atoms in forming the cavity mirrors is clearly more pronounced on the directivity function of the stationary field than on the transient population of the atoms. First we note from the figure that there are distances between the atoms at which the directivity function is nonzero only along the interatomic axis [ Figs.~\ref{fig11}(b) and~\ref{fig11}(c) ]. It can also be noticed that there is an optimal distance between the mirror atoms at which the solid angle subtended by the cone is minimal, or equivalently the width of the cavity axial mode is very narrow. As seen from Fig.~\ref{fig11}(c), a reduction of the separation may result in an increase of the beamwidth. We also see quite clearly that larger separations between the mirror atoms, additional directions of emission appear [ Fig.~\ref{fig11}(d) ]. \section{Summary}\label{sec5} We have studied directional properties of the radiation field emitted by a chain of closely located and dipole-dipole interacting two-level atoms. Two geometrical configurations of atoms in the chain have been studied: A probe atom in front of a finite-size chain of closely located atoms, and a probe atom between a pair of chains of closely located atoms. We call the earlier case as an atom in front of an atomic mirror, and the former as an atom inside a cavity whose the mirrors are formed by the two chains of atoms. We have found that it is possible to account for certain mirror and cavity-like features, such as We have examined the conditions for one-sided emission centered about the interatomic axis, and have found a lower bound for the distance between the atoms above which a one-sided emission along the interatomic axis can be achieved. The one-sided emission focused into a cone about the interatomic axis and oriented in the backward direction can be regarded as a mirror type with perfect reflectivity. For atomic distances smaller than the lower bound, a part of the population can be trapped in the probe atom indicating that at these distances there are no radiative modes available for the emission. This is a situation similar to that one encounters in a Fabry-P\'erot cavity composed of a pair of parallel plate mirrors~\cite{dowling}. Polar diagrams have been presented showing the variation of the directivity function of the time dependent as well as the stationary fields with the separation between the atoms. We have found that the directional properties of the radiation field are generally more manifested in the stationary field when the probe atom is continuously driven by a coherent laser field. The control of directionality of photon emission could have applications in quantum information processing in realizing a directional quantum network. For example, the mirror-like behavior of the chain could be used to realize a chiral quantum network, where atoms interact via one-way emission~\cite{pr15,gg15,vr16}. It could also be used as an optical reflector or an optical mirror to create, together with a distant dielectric mirror, a Fabry-P\'erot cavity~\cite{hs11}. \acknowledgments Q. G. acknowledges support from 6th research grant (proposal number 3775) from Jazan University, Saudi Arabia.
244f55be183dadbe0832ae9dd9b2e53e3fc4e5ae
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Let $K$ be a convex body in ${\mathbb R}^n$, with barycenter at the origin (we call these bodies centered). We denote by ${\rm as}(K)$ the average volume of the central hyperplane sections of $K$: \begin{equation}\label{eq:intro-1}{\rm as}(K)=\int_{S^{n-1}}|K\cap \xi^{\perp }|\,d\sigma (\xi ),\end{equation} where $|\cdot |$ denotes volume in the appropriate dimension, $\xi^{\perp }$ is the subspace perpendicular to $\xi $, and $\sigma $ is the rotationally invariant probability measure on $S^{n-1}$. More generally, for any $1\leqslant r\leqslant n-1$ we define \begin{equation}\label{eq:intro-2}{\rm as}_r(K)=\int_{{\rm Gr}_{n-r}}|K\cap E|\,d\nu_{n-r}(E),\end{equation} where $\nu_{n-r}$ is the Haar probability measure on the Grassmannian ${\rm Gr}_{n-r}$ of $(n-r)$-dimensional subspaces of ${\mathbb R}^n$. Thus, ${\rm as}_r(K)$ is the average volume of $r$-codimensional central sections of $K$; note that ${\rm as}(K)={\rm as}_1(K)$. The fourth named author proved in \cite{Koldobsky-2013} that if $K$ is an intersection body in ${\mathbb R}^n$ (see Section \ref{notation} for definitions and background information) then \begin{equation}\label{eq:intro-3}{\rm as}(K)\leqslant b_{n,1}\,|K|^{\frac{1}{n}}\max_{\xi\in S^{n-1}}{\rm as}(K\cap\xi^{\perp }),\end{equation} where $$b_{n,1}:=\frac{\omega_{n-1}}{\omega_{n-2}\omega_n^{\frac{1}{n}}}\simeq 1$$ (and $\omega_m$ denotes the volume of the Euclidean unit ball $B_2^m$ in ${\mathbb R}^m$). Whenever we write $a\lesssim b$ we mean that there exists an absolute constant $c>0$ such that $a\leqslant cb$, and whenever we write $a\simeq b$, we mean that $a\lesssim b$ and $b\lesssim a$. Note that \eqref{eq:intro-3} is sharp: it becomes equality if $K=B_2^n$. The purpose of this article is to discuss similar inequalities for the average volume of hyperplane sections of an arbitrary centered convex body $K$ in ${\mathbb R}^n$. More precisely, we study the following question. \begin{question}\label{question-intro}Let $1\leqslant k<n$ and define $\gamma_{n,k}$ as the smallest constant $\gamma >0$ for which the following holds true: for every centered convex body $K$ in ${\mathbb R}^n$ we have \begin{equation}\label{eq:intro-4}{\rm as}(K)\leqslant \gamma^k|K|^{\frac{k}{n}}\,\max_{E\in {\rm Gr}_{n-k}}{\rm as}(K\cap E).\end{equation} Is it true that $\sup_{n,k}\gamma_{n,k}<\infty $? \end{question} In Section \ref{ovr} we generalize \eqref{eq:intro-3} using as a parameter the outer volume ratio distance $d_{\rm ovr}(K,{\cal{BP}}_k^n)$ from an origin-symmetric convex body $K$ to the class ${\cal{BP}}_k^n$ of generalized $k$-intersection bodies. Our estimates are based on the next more general theorem which is valid for the larger class of origin-symmetric star bodies in ${\mathbb R}^n$ and for any even continuous density on ${\mathbb R}^n$. \begin{theorem}\label{th:intro-2}Let $1\leqslant k\leqslant n-1$, let $K$ be an origin-symmetric star body in ${\mathbb R}^n$, and let $f$ be a non-negative even continuous function on ${\mathbb R}^n.$ Then \begin{equation}\label{eq:intro-5} \int_{S^{n-1}} \rho_K^{n-1}(\theta) f(\rho_K(\theta)\theta)\ d\theta \leqslant c_{n,k}^k \ d_{\rm ovr}^k (K,{\cal{BP}}_k^n)\ |K|^{\frac kn} \max_{E\in {\rm Gr}_{n-k}} \int_{S^{n-1}\cap E} \rho_K^{n-k-1}(\theta)f(\rho_K(\theta)\theta)\ d\theta. \end{equation} \end{theorem} In the statement above, $\rho_K$ is the radial function of a star body $K$ and we use the notation $d\theta$ for the non-normalized measure on the sphere with density $1$. The constant $c_{n,k}$ is given by $$c_{n,k}^k=\frac{n\omega_n^{\frac{n-k}n}}{(n-k)\omega_{n-k}},$$ and one can check that $c_{n,k}\simeq 1.$ Theorem \ref{th:intro-2} provides a first estimate on the constants $\gamma_{n,k}$ of Question \ref{question-intro}. Choosing $f\equiv 1$ we see that \eqref{eq:intro-5} implies the following. \begin{theorem}\label{th:intro-3}Let $1\leqslant k\leqslant n-1$, and let $K$ be an origin-symmetric star body in ${\mathbb R}^n$. Then, \begin{equation}\label{eq:intro-6}{\rm as}(K)\leqslant b_{n,k}^kd_{\rm ovr}^k(K,{\cal{BP}}_k^n)\, |K|^{\frac{k}{n}}\, \max_{E\in Gr_{n-k}} {\rm as}(K\cap E).\end{equation} In other words, $\gamma_{n,k}\leqslant b_{n,k}d_{\rm ovr}(K,{\cal{BP}}_k^n)$.\end{theorem} The constant $b_{n,k}$ in Theorem \ref{th:intro-3} is given by $$b_{n,k}^k =\frac{\omega_{n-1}}{\omega_{n-k-1} \omega_n^{\frac{k}{n}}},$$ and one can also check that $b_{n,k}\simeq 1$. \smallskip In the case where the body $K$ is convex, the distance $d_{\rm ovr}(K,{\cal{BP}}_k^n)$ was estimated in \cite{KPZ}. In particular, the available bounds for $d_{\rm ovr}(K,{\cal{BP}}_k^n)$ show that $\gamma_{n,k}$ is bounded by a function of $n/k$, and hence it remains bounded as long as $k$ is proportional to $n$. More generally, we have: \begin{theorem} \label{th:intro-4}For every origin-symmetric convex body $K$, for every $1\leqslant k \leqslant n-1$ and every even non-negative continuous function $f$ on ${\mathbb R}^n$, \begin{equation}\label{eq:intro-7}\int_{S^{n-1}} \rho_K^{n-1}(\theta) f(\rho_K(\theta)\theta)\ d\theta \leqslant \big ( c_1h(n/k)\big )^k \ |K|^{\frac{k}{n}} \max_{E\in {\rm Gr}_{n-k}} \int_{S^{n-1}\cap E} \rho_K^{n-k-1}(\theta)f(\rho_K(\theta)\theta)\ d\theta, \end{equation} where $c_1>0$ is an absolute constant and $h(t)=\sqrt{t}\cdot (\log (et))^{\frac{3}{2}}$, $t\geqslant 1$. In particular, \begin{equation}\label{eq:intro-8}\gamma_{n,k}\leqslant c_1\sqrt{n/k}\,[\log (en/k)]^{\frac{3}{2}}.\end{equation} \end{theorem} It is also known that for many classes of convex bodies the distance $d_{\rm ovr}(K,{\cal{BP}}_k^n)$ is bounded by an absolute constant. This includes unconditional bodies, unit balls of subspaces of $L_p,$ and others. Therefore, the restriction of Question \ref{question-intro} to all these classes has an affirmative answer. \smallskip Theorem \ref{th:intro-2} also allows us to prove an analogue of Theorem \ref{th:intro-3} for the quantities ${\rm as}_r(K).$ \begin{theorem}\label{th:intro-5}Let $1\leqslant k < n-2$ and $1\leqslant r < n-k$. For any origin-symmetric star body $K$ in ${\mathbb R}^n$ we have that \begin{equation}\label{eq:intro-9}{\rm as}_r(K)\leqslant \phi_{n,k,r}^k d_{\rm ovr}^k(K,{\cal{BP}}_k^n)\, |K|^{\frac{k}{n}}\, \max_{E\in Gr_{n-k}} {\rm as}_r(K\cap E).\end{equation} \end{theorem} Here $$\phi_{n,k,r}^k =\frac{\omega_{n-r}}{\omega_{n-k-r}\omega_n^{\frac kn}},$$ and one can check that $\phi_{n,k,r}\simeq \sqrt{\frac{n}{n-r}}$. \smallskip In Section \ref{iso-bounds} we show that an analogue of \eqref{eq:intro-3} holds true in full generality, up to the value of the isotropic constant of $K$. In order to state our main result we recall the definition of the isotropic position. A centered convex body $K$ of volume $1$ in ${\mathbb R}^n$ is called isotropic if there exists a constant $L_K>0$ such that \begin{equation}\label{eq:intro-10}\int_K\langle x,\xi\rangle^2dx =L_K^2\end{equation} for every $\xi\in S^{n-1}$. Every centered convex body $K$ has an {\sl isotropic position} $T(K)$, $T\in GL(n)$, which is uniquely defined modulo orthogonal transformations, and hence the {\sl isotropic constant} $L_K$ is an invariant of the linear class of $K$. A well-known question in asymptotic convex geometry asks if there exists an absolute constant $C>0$ such that $L_K\leqslant C$ for every $n$ and every centered convex body $K$ in ${\mathbb R}^n$. The best known upper bound \begin{equation}\label{eq:intro-11}L_n:=\sup\{L_K:K\;\hbox{isotropic in}\;{\mathbb R}^n\}\leqslant c\sqrt[4]{n}\end{equation} is due to Klartag \cite{Klartag-2006} (see also \cite{BGVV-book} for the history of the problem and recent developments in this area). On the other hand, one always has $L_K\geqslant L_{B_2^n}\geqslant c$, where $c>0$ is an absolute constant. In other words, the question is if $L_K\simeq 1$ for all centered convex bodies. \begin{theorem}\label{th:intro-6}Let $K$ be a centered convex body in ${\mathbb R}^n$. Then, for every $1\leqslant k\leqslant n-1$, \begin{equation}\label{eq:intro-12}{\rm as}(K)\leqslant (c_2L_K)^k\,|K|^{\frac{k}{n}}\,\max_{E\in {\rm Gr}_{n-k}}{\rm as}(K\cap E),\end{equation} where $c_2>0$ is an absolute constant and $L_K$ is the isotropic constant of $K$. \end{theorem} It is known that for many classes of convex bodies the isotropic constant $L_K$ is bounded by an absolute constant (see \cite[Chapter 4]{BGVV-book}). Theorem \ref{th:intro-6} provides an affirmative answer to Question \ref{question-intro} for all these classes. On the other hand, it is interesting to note that \eqref{eq:intro-12} is essentially the best bound we can hope for. We show in Proposition \ref{prop:isotropic} that if $K$ is an isotropic convex body in ${\mathbb R}^n$ then \begin{equation}\label{eq:intro-13}{\rm as}(K)\simeq L_K\,\max_{\xi\in S^{n-1}}{\rm as}(K\cap\xi^{\perp })\,|K|^{\frac{1}{n}}.\end{equation} This shows that the estimate of Theorem \ref{th:intro-6} is asymptotically sharp: if $\gamma >0$ is a constant such that \eqref{eq:intro-4} holds for $k=1$ and all $K$ then we must have $\gamma \geqslant cL_K$. Combining this fact with Theorem \ref{th:intro-6} we actually conclude that \begin{equation}\label{eq:intro-14}\gamma_{n,k}\lesssim \gamma_{n,1}\simeq L_n\end{equation} for all $1\leqslant k\leqslant n-1$ (see Proposition \ref{prop:worse-k}). One of the tools that are used in the proof of Theorem \ref{th:intro-6} is a variant of Meyer's dual Loomis-Whitney inequality \cite{Meyer-1988} that was recently obtained in \cite{Brazitikos-Giannopoulos-Liakopoulos-2016}; see \eqref{eq:local-2}. The second tool is a lower bound for the dual affine quermassintegrals \begin{equation}\label{eq:intro-15}\tilde{\Phi}_k(K):=\frac{\omega_n}{\omega_{n-k}}\left (\int_{{\rm Gr}_{n-k}}|K\cap E|^n\,d\nu_{n-k}(E)\right )^{\frac{1}{n}} \end{equation}of a convex body $K$ in ${\mathbb R}^n$ in terms of the isotropic constant of $K$ (see \cite{Dafnis-Paouris-2012}). In fact, one can check that the problem to obtain asymptotically sharp lower bounds for $\tilde{\Phi }_k(K)$ is equivalent to the question whether $\gamma_{n,1}\simeq L_n\simeq 1$ (see Remark \ref{rem:dual-affine}). When the codimension $k$ is proportional to $n$ the available lower bounds are independent from the isotropic constant of $K$ (see \cite{Dafnis-Paouris-2012} and \cite[Section~6.4]{BGVV-book}). Thus, we get a variant of Theorem \ref{th:intro-4}. \begin{theorem}\label{th:intro-7} Let $1\leqslant k \leqslant n-1$ and let $K$ be a centered convex body in ${\mathbb R}^n.$ Then, \begin{equation}\label{eq:intro-16} {\rm as}(K)\leqslant \big (c_3h(n/k)\big )^k \,|K|^{\frac{k}{n}}\,\max_{E\in Gr_{n-k}} {\rm as}(K\cap E),\end{equation} where $c_2>0$ is an absolute constant and $h(t)=\sqrt{t}\cdot (\log (et))^{\frac{3}{2}}$, $t\geqslant 1$. \end{theorem} The methods that are used for the proof of Theorem \ref{th:intro-6} and Theorem \ref{th:intro-3} are independent. Note that the first method allows us to work with (not necessarily symmetric) centered convex bodies while the second method allows us to work with origin-symmetric (not necessarily convex) star bodies and to consider even continuous densities in place of volume. Therefore, the two results complement each other. A link between the two bounds is given by the inequality \begin{equation}\label{eq:link}L_K\leqslant cL_k\cdot d_{\rm ovr}(K,{\cal{BP}}_k^n)\end{equation} which is due to E.~Milman (see \cite[Corollary~5.4]{Milman-2006}). However, since we only know that $L_k=O(\sqrt[4]{k})$, the estimates of Theorem \ref{th:intro-6} and Theorem \ref{th:intro-3} are incomparable for $k\gg 1$. \smallskip In Section \ref{positions} we discuss the mean value of the average section functional ${\rm as}(K\cap E)$ over all $E\in {\rm Gr}_{n-k}$, $1\leqslant k\leqslant n-1$. We obtain the next general upper and lower bounds. \begin{theorem}\label{th:intro-8}Let $K$ be a centered convex body in ${\mathbb R}^n$ and define $p(K):=R(K)/|K|^{\frac{1}{n}}$, where $R(K)$ is the circumradius of $K$. Then, for every $1\leqslant k\leqslant n-1$ we have that \begin{equation}\label{eq:intro-17}\left (\frac{c_4\sqrt{n}}{p(K)}\right )^k\,{\rm as}(K)\leqslant |K|^{\frac{k}{n}}\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E)\leqslant \left (\frac{c_5p(K)}{\sqrt{n}}\right )^{\frac{k}{n-1}}\,{\rm as}(K),\end{equation} where $c_4, c_5>0$ are absolute constants. \end{theorem} Since $R(K)$ is polynomial in $n$ for all the classical positions of a convex body $K$ (isotropic position, minimal surface position, minimal mean width position, John's and L\"{o}wner's position) the right hand side inequality of \eqref{eq:intro-17} implies the following. \begin{theorem}\label{th:intro-9}Let $K$ be a centered convex body in ${\mathbb R}^n$. If $K$ is in one of the classical positions then \begin{equation}\label{eq:intro-18}|K|^{\frac{k}{n}}\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E)\leqslant c_6^k\,{\rm as}(K)\end{equation} for every $1\leqslant k\leqslant n-1$, where $c_6>0$ is an absolute constant. \end{theorem} Closing this introductory section we would like to note that the results of this article are dual to the ones in \cite{Giannopoulos-Koldobsky-Valettas-2016}. In that work, the main question was to compare the surface area $S(K)$ of a convex body $K$ in ${\mathbb R}^n$ to the minimal, average or maximal surface area of its hyperplane or lower dimensional projections. One of the main results in \cite{Giannopoulos-Koldobsky-Valettas-2016} states that there exists an absolute constant $c_1>0$ such that, for every convex body $K$ in ${\mathbb R}^n$, \begin{equation}\label{eq:intro-19}|K|^{\frac{1}{n}}\,\min_{\xi\in S^{n-1}}S(P_{\xi^{\perp }}(K))\leqslant\frac{c_7\partial_K}{\sqrt{n}}\,S(K),\end{equation} where $c_7>0$ is an absolute constant and \begin{equation}\label{eq:intro-20}\partial_K:=\min\Big\{ S(T(K))/|T(K)|^{\frac{n-1}{n}}:T\in GL(n)\Big\}\end{equation} is the {\sl minimal surface area parameter} of $K$. Another result from \cite{Giannopoulos-Koldobsky-Valettas-2016} asserts that if $K$ is in some of the classical positions mentioned above, then \begin{equation}\label{eq:intro-21}|K|^{\frac{1}{n}}\,\int_{S^{n-1}}S(P_{\xi^{\perp }}(K))\,d\sigma (\xi )\geqslant c_8S(K),\end{equation} where $c_9>0$ is an absolute constant. The analogy with Theorem \ref{th:intro-6} and Theorem \ref{th:intro-9} is clear; the role of the average section functional ${\rm as}(K)$ is played by the surface area $S(K)$, and the role of the isotropic constant is played by the minimal surface area parameter. \section{Notation and background}\label{notation} We work in ${\mathbb R}^n$, which is equipped with a Euclidean structure $\langle\cdot ,\cdot\rangle $. We denote by $\|\cdot \|_2$ the corresponding Euclidean norm, and write $B_2^n$ for the Euclidean unit ball and $S^{n-1}$ for the unit sphere. Volume is denoted by $|\cdot |$. We write $\omega_n$ for the volume of $B_2^n$ and $\sigma $ for the rotationally invariant probability measure on $S^{n-1}$. We use the notation $d\theta$ for the non-normalized measure on the sphere with density $1$. If $\xi\in S^{n-1}$, then $\xi^{\perp }=\{x\in {\mathbb R}^n:\langle x,\xi\rangle =0\}$. The Grassmann manifold ${\rm Gr}_m$ of $m$-dimensional subspaces of ${\mathbb R}^n$ is equipped with the Haar probability measure $\nu_{m}$. For every $1\leqslant m\leqslant n-1$ and $E\in {\rm Gr}_m$ we write $P_E$ for the orthogonal projection from $\mathbb R^{n}$ onto $E$, and we set $B_E=B_2^n\cap F$ and $S_E=S^{n-1}\cap E$. The letters $c,c^{\prime }, c_1, c_2$ etc. denote absolute positive constants which may change from line to line. We refer to the books \cite{Gardner-book} and \cite{Schneider-book} for basic facts from the Brunn-Minkowski theory and to the book \cite{AGA-book} for basic facts from asymptotic convex geometry. \medskip \noindent {\bf 2.1. Star bodies and convex bodies}. A convex body in ${\mathbb R}^n$ is a compact convex subset $K$ of ${\mathbb R}^n$ with non-empty interior. We say that $K$ is symmetric if $x\in K$ implies that $-x\in K$, and that $K$ is centered if its barycenter $\frac{1}{|K|}\int_Kx\,dx $ is at the origin. A compact set $K$ in ${\mathbb R}^n$ will be called star-shaped at $0$ if it contains the origin in its interior and every line through $0$ meets $K$ in a line segment. For such a set, the radial function $\rho_K $ is defined on $S^{n-1}$ by \begin{equation}\label{eq:notation-1}\rho_K(\theta )=\max \{\lambda >0:\lambda\theta\in K\},\qquad \theta\in S^{n-1}.\end{equation} If $\rho_K $ is continuous, then we say that $K$ is a star body. Then, the volume of $K$ in polar coordinates is given by \begin{equation}\label{eq:notation-2}|K|=\omega_n\int_{S^{n-1}}\rho_K^n(\theta )\,d\sigma (\theta ).\end{equation} The radial sum $K\tilde{+} D$ of two star bodies $K$ and $D$ is defined by \begin{equation}\label{eq:notation-3}\rho_{K\tilde{+} D}= \rho_K + \rho_D.\end{equation} We equip the class ${\cal S}_n$ of star bodies with the radial metric \begin{equation}\label{eq:notation-4}d_r(K,D):=\sup_{\xi\in S^{n-1}} |\rho_K(\xi)-\rho_D(\xi)|.\end{equation} The support function of a convex body $K$ is defined by $h_K(y)=\max \{\langle x,y\rangle :x\in K\}$, and the mean width of $K$ is \begin{equation}\label{eq:notation-5}w(K)=\int_{S^{n-1}}h_K(\theta )\,d\sigma (\theta ). \end{equation} The circumradius of $K$ is the smallest $R>0$ for which $K\subseteq RB_2^n$. If $0\in {\rm int}(K)$ then we write $r(K)$ for the inradius of $K$ (the largest $r>0$ for which $rB_2^n\subseteq K$) and we define the polar body $K^{\circ }$ of $K$ by \begin{equation}\label{eq:notation-6}K^{\circ }:=\{ y\in {\mathbb R}^n: \langle x,y\rangle \leqslant 1 \;\hbox{for all}\; x\in K\}. \end{equation} The section of a star body $K$ with $\xi^{\perp }$ is denoted by $K\cap\xi^{\perp }$, and we write $P_{\xi^{\perp } }(K)$ for the orthogonal projection of $K$ onto $\xi^{\perp }$. The volume radius of $K$ is the quantity ${\rm vrad}(K)=\left (|K|/|B_2^n|\right )^{1/n}$. We also define $\|\theta \|_K=\min\{ t>0:\theta \in tK\}$ and \begin{equation}\label{eq:notation-7}M(K)=\int_{S^{n-1}}\rho_K^{-1}(\theta )\,d\sigma (\theta )=\int_{S^{n-1}}\|\theta\|_K\,d\sigma (\theta ).\end{equation} \medskip \noindent {\bf 2.2. Dual mixed volumes}. Lutwak introduced dual mixed volumes in \cite{Lutwak-1975}; he first considered convex bodies, but then extended his definition to the class ${\mathcal S}_n$ of star bodies. Given $K_1,\ldots,K_n\in {\mathcal S}_n$, their dual mixed volume is the integral \begin{equation}\label{eq:notation-8}\tilde{V}(K_1,\ldots,K_n)=\omega_n \int_{S^{n-1}} \rho_{K_1}(\theta )\cdots \rho_{K_n}(\theta )d\sigma (\theta ).\end{equation} The observation is that such integrals have properties analogous to those of mixed volumes if one replaces Minkowski addition by radial addition. The function $\tilde{V}$ is clearly non-negative, symmetric and monotone with respect to its arguments, positive linear with respect to $\tilde{+}$ in each of its arguments, and has volume as its diagonal. A simple calculation shows that if $K_1,\ldots,K_m\in {\mathcal S}_n$ and $\lambda_1,\ldots,\lambda_m>0$, then \begin{equation}\label{eq:notation-9}|\lambda_1 K_1 \tilde{+}\cdots\tilde{+}\lambda_mK_m|= \sum_{i_1,\ldots,i_n=1}^m \tilde{V}(K_{i_1},\ldots,K_{i_n})\lambda_{i_1}\ldots\lambda_{i_n}.\end{equation} In particular, if $K,D\in {\mathcal S}_n$ and $t>0$ then \begin{equation}\label{eq:notation-10}|K\tilde{+}tD|=\sum_{j=0}^n {n\choose j}\tilde{V}_j(K,D)\;t^j,\end{equation} where $\tilde{V}_j(K,D):=\omega_n\int_{S^{n-1}}\rho_K^{n-j}(\theta )\rho_D^{j}(\theta )d\sigma (\theta )$, is the $j$-th dual mixed volume of $K$ and $D$. An inequality which further illustrates the analogy with the mixed volumes is the dual Minkowski's inequality: for every $K,D\in {\mathcal S}_n$ an application of H\"older's inequality gives \begin{equation}\label{eq:notation-11}\tilde{V}_1(K,D) \leqslant \left( \omega_n\int_{S^{n-1}}\rho_K^n(\theta ) d\sigma (\theta )\right)^{\frac{n-1}{n}}\left(\omega_n\int_{S^{n-1}}\rho_D^n(\theta )d\sigma (\theta )\right)^{\frac{1}{n}} \leqslant |K|^{\frac{n-1}{n}}|D|^{\frac{1}{n}}.\end{equation} \medskip \noindent {\bf 2.3. Intersection bodies}. The class of intersection bodies was introduced by Lutwak \cite{Lutwak-1988}. The intersection body of a star body $K$ in ${\mathbb R}^n$ with $\rho_K\in C(S^{n-1})$ is the star body $IK$ with radial function \begin{equation}\label{eq:notation-12}\rho_{IK}(\xi )=|K\cap \xi^{\perp }|=\omega_{n-1}\int_{S(\xi^{\perp })}\rho_K^{n-1}(\theta )d\sigma_{\xi }(\theta ),\end{equation} where $S(\xi^{\perp })=S^{n-1}\cap\xi^{\perp }$ is the Euclidean unit sphere of $\xi^{\perp }$ and $\sigma_{\xi }$ denotes the rotationally invariant probability measure on $S(\xi^{\perp })$. If $K$ is a centered convex body then $IK$ is a symmetric convex body. It is known that \begin{equation}\label{eq:notation-13}I(TK)=|\det T|\,(T^{-1})^{\ast }(IK)\end{equation} for every $T\in GL(n)$. In particular, if $T\in SL(n)$ we see that $|I(TK)|=|IK|$. The class of intersection bodies ${\cal{I}}_n$ is defined as the closure in the radial metric of intersection bodies of star bodies. Zhang introduced more general classes of bodies in \cite{Zhang-1996}. For $1\leqslant k \leqslant n-1,$ the {\it $(n-k)$-dimensional spherical Radon transform} $R_{n-k}:C(S^{n-1})\to C({\rm Gr}_{n-k})$ is a linear operator defined by \begin{equation}\label{eq:notation-14}R_{n-k}g (E)=\int_{S^{n-1}\cap E} g(\theta)\ d\theta,\qquad E\in {\rm Gr}_{n-k}\end{equation} for every function $g\in C(S^{n-1}).$ We say that an origin-symmetric star body $D$ in ${\mathbb R}^n$ is a {\it generalized $k$-intersection body}, and write $D\in {\cal{BP}}_k^n$, if there exists a finite non-negative Borel measure $\mu_D$ on ${\rm Gr}_{n-k}$ so that for every $g\in C(S^{n-1})$ \begin{equation}\label{eq:notation-16} \int_{S^{n-1}} \rho_D^{k}(\theta) g(\theta)\ d\theta=\int_{{\rm Gr}_{n-k}} R_{n-k}g(H)\ d\mu_D(H). \end{equation} For a star body $K$ in ${\mathbb R}^n$ and $1\leqslant k \leqslant n-1$, we denote by $$d_{\rm {ovr}}(K,{\cal{BP}}_k^n) = \inf \left\{ \left( \frac {|D|}{|K|}\right)^{1/n}:\ K\subset D,\ D\in {\cal{BP}}_k^n \right\}$$ the outer volume ratio distance from $K$ to the class ${\cal{BP}}_k^n$. The reader will find more information on the Radon transform and intersection bodies in the book \cite{Koldobsky-book}. \section{Bounds in terms of the outer volume ratio distance to the class of generalized $k$-intersection bodies} \label{ovr} The main result of this section is Theorem \ref{th:intro-2} which is valid for the larger class of origin-symmetric star bodies in ${\mathbb R}^n$ and for any even continuous density on ${\mathbb R}^n$. By an appropriate choice of the density $f$ we obtain Theorem \ref{th:intro-3} and its generalization in Theorem \ref{th:intro-5}. \medskip \noindent{\bf Proof of Theorem \ref{th:intro-2}.} Let $\varepsilon>0.$ For every $E\in {\rm Gr}_{n-k},$ we have \begin{equation}\label{eq:ovr-1}\int_{(K\tilde{+} \varepsilon B_2^n)\cap E} f(x) dx - \int_{K\cap E} f(x) dx\leqslant \max_{F\in {\rm Gr}_{n-k}} \left (\int_{(K\tilde{+} \varepsilon B_2^n)\cap F} f(x) dx - \int_{K\cap F} f(x) dx\right ).\end{equation} Note that $\rho_{K\tilde{+}\varepsilon B_2^n}=\rho_K+\varepsilon$. Expressing the integrals in polar coordinates we get \begin{equation}\label{eq:ovr-2}R_{n-k}\left(\int_{\rho_K(\cdot)}^{\rho_{K}(\cdot)+\varepsilon} r^{n-k-1}f(r\cdot) dr \right)(E) \leqslant \max_{F\in {\rm Gr}_{n-k}} \left(\int_{S^{n-1}\cap F} \int_{\rho_K(\theta)}^{\rho_{K}(\theta)+\varepsilon} r^{n-k-1}f(r\theta) dr d\theta \right). \end{equation} Let $D\in {\cal{BP}}_k^n$ such that $K\subset D.$ Integrating the latter inequality by $E$ over ${\rm Gr}_{n-k}$ with the measure $\mu_D$ corresponding to $D$ by \eqref{eq:notation-14}, we get \begin{align}\label{eq:ovr-3}&\int_{S^{n-1}}\rho_D^k(\theta)\int_{\rho_K(\theta)}^{\rho_{K}(\theta)+\varepsilon} r^{n-k-1}f(r\theta) dr d\theta\\ \nonumber &\hspace*{1.5cm} \leqslant \mu_D({\rm Gr}_{n-k})\cdot \max_{F\in {\rm Gr}_{n-k}} \left(\int_{S^{n-1}\cap F} \int_{\rho_K(\theta)}^{\rho_{K}(\theta)+\varepsilon} r^{n-k-1}f(r\theta) dr d\theta \right). \end{align} We divide both sides by $\varepsilon$ and send $\varepsilon$ to zero. Note that we can interchange the limit with the maximum, because the convergence is uniform with respect to $F$. Thus, we get \begin{equation}\label{eq:ovr-4}\int_{S^{n-1}}\rho_D^k(\theta)\rho_K^{n-k-1}(\theta) f(\rho_K(\theta)\theta)d\theta\leqslant \mu_D({\rm Gr}_{n-k})\cdot \max_{F\in {\rm Gr}_{n-k}} \left(\int_{S^{n-1}\cap F} \rho_K^{n-k-1}(\theta) f(\rho_K(\theta)\theta) d\theta\right).\end{equation} The integral in the left hand side can be estimated from below by $\int_{S^{n-1}} \rho_K^{n-1}(\theta) f(\rho_K(\theta)\theta)\ d\theta,$ because $K\subset D.$ To estimate $\mu_D({\rm Gr}_{n-k})$ from above, we combine the fact that $1= R_{n-k}{\bf 1}(E)/|S^{n-k-1}|$ for every $E\in {\rm Gr}_{n-k}$ with Definition \eqref{eq:notation-14} and H\"older's inequality to write \begin{align}\label{eq:ovr-5} \mu_D({\rm Gr}_{n-k}) &= \frac 1{\left|S^{n-k-1}\right|} \int_{{\rm Gr}_{n-k}} R_{n-k}{\bf 1}(E) d\mu_D(E)\\ \nonumber &=\frac 1{\left| S^{n-k-1} \right| } \int_{S^{n-1}} \|\theta\|_D^{-k}\ d\theta \\ \nonumber &\leqslant \frac 1{\left|S^{n-k-1}\right|} \left|S^{n-1}\right|^{\frac{n-k}n} \left(\int_{S^{n-1}} \|\theta\|_D^{-n}\ d\theta\right)^{\frac kn}\\ \nonumber &= \frac{1}{\left|S^{n-k-1}\right|} \left|S^{n-1}\right|^{\frac{n-k}n} n^{\frac{k}{n}}|D|^{\frac{k}{n}}. \end{align} These estimates show that \begin{align}\label{eq:ovr-6}&\int_{S^{n-1}} \rho_K^{n-1}(\theta)f(\rho_K(\theta)\theta) d\theta \\ \nonumber &\hspace*{1.5cm} \leqslant \frac{1}{\left|S^{n-k-1}\right|} \left|S^{n-1}\right|^{\frac{n-k}n} n^{\frac{k}{n}}|D|^{\frac{k}{n}} \max_{F\in {\rm Gr}_{n-k}} \left(\int_{S^{n-1}\cap F} \rho_K^{n-k-1}(\theta)f(\rho_K(\theta)\theta) d\theta\right). \end{align} Finally, we choose $D$ so that $|D|^{1/n}\leqslant (1+\delta)d_{\rm ovr}(K,{\cal{BP}}_k^n)|K|^{1/n},$ and then send $\delta$ to zero. \qed \medskip All the other results of this section are consequences of Theorem \ref{th:intro-2}. \medskip \noindent {\bf Proof of Theorem \ref{th:intro-3}.} First, we express the average section functionals ${\rm as}(K)$ and ${\rm as}(K\cap E)$ in terms of the radial function of $K$. Using \eqref{eq:notation-2} we write \begin{align}\label{eq:ovr-7}{\rm as}(K) &=\int_{S^{n-1}}|K\cap\xi^{\perp }|\,d\sigma (\xi )=\omega_{n-1}\int_{S^{n-1}} \int_{S(\xi^{\perp })}\rho_K^{n-1}(\theta )\,d\sigma_{\xi }(\theta )\,d\sigma (\xi )\\ \nonumber &=\omega_{n-1}\int_{S^{n-1}}\rho_K^{n-1}(\theta )\,d\sigma (\theta ).\end{align} Similarly, for every $1\leqslant k\leqslant n-1$ and any $E\in {\rm Gr}_{n-k}$, we have \begin{equation}\label{eq:ovr-8}{\rm as}(K\cap E)=\omega_{n-k-1}\int_{S_E}\rho_K^{n-k-1}(\theta )\,d\sigma_E(\theta ),\end{equation} where $\sigma_E$ is the rotationally invariant probability measure on $S_E=S^{m-1}\cap E$. Applying Theorem \ref{th:intro-2} for the density $f\equiv {\bf 1}$ we get \begin{equation}\label{eq:ovr-9} \int_{S^{n-1}} \rho_K^{n-1}(\theta) \ d\theta \leqslant c_{n,k}^k \ d_{\rm ovr}^k (K,{\cal{BP}}_k^n)\ |K|^{\frac kn} \max_{E\in {\rm Gr}_{n-k}} \int_{S^{n-1}\cap E} \rho_K^{n-k-1}(\theta)\ d\theta , \end{equation} and Theorem \ref{th:intro-3} follows from \eqref{eq:ovr-7} and \eqref{eq:ovr-8} and an adjustment of the constants. \qed \begin{remark}\label{rem:ovr}\rm For certain classes of origin-symmetric convex bodies the distance $d_{\rm ovr}(K,{\cal{BP}}_k^n)$ is bounded by an absolute constant. These classes include unconditional convex bodies and duals of bodies with bounded volume ratio (see \cite{Koldobsky-2015}) and the unit balls of normed spaces that embed in $L_p,\ -n<p<\infty$ (see \cite{Koldobsky-2016}, \cite{Milman-2006} and \cite{Koldobsky-Pajor}). If we restrict Question \ref{question-intro} to any of these classes then Theorem \ref{th:intro-3} provides an affirmative answer. \end{remark} \noindent {\bf Proof of Theorem \ref{th:intro-4}.} We combine Theorem \ref{th:intro-2} with the following result from \cite{KPZ}: For every origin-symmetric convex body $K$ in ${\mathbb R}^n$, \begin{equation}\label{eq:ovr-10}d_{\rm ovr}(K,{\cal{BP}}_k^n)\leqslant c\sqrt{n/k}\,[\log (en/k)]^{\frac{3}{2}},\end{equation} where $c>0$ is an absolute constant. \qed \medskip \noindent {\bf Proof of Theorem \ref{th:intro-5}.} We choose $f(x)=\|x\|_2^{-r+1}$ in Theorem \ref{th:intro-2} to get \begin{equation}\label{eq:ovr-11} \int_{S^{n-1}} \rho_K^{n-r}(\theta)\ d\theta \leqslant c_{n,k}^k \ d_{\rm ovr}^k (K,{\cal{BP}}_k^n)\ |K|^{\frac kn} \max_{E\in {\rm Gr}_{n-k}} \int_{S^{n-1}\cap E} \rho_K^{n-k-r}(\theta)\ d\theta. \end{equation} Then, we apply the formula \begin{equation}\label{eq:ovr-12}{\rm as}_r(K)= \omega_{n-r} \int_{S^{n-1}} \rho_K^{n-r}(\theta) d\sigma(\theta)\end{equation} which generalizes \eqref{eq:ovr-7} and is easily verified in the same way. \qed \section{Bounds in terms of the isotropic constant}\label{iso-bounds} Let $K$ be a centered convex body in ${\mathbb R}^n$. In this section we compare ${\rm as}(K)$ with the corresponding average section functional ${\rm as}(K\cap E)$ for any $k$-codimensional subspace $E$ of ${\mathbb R}^n$. Our main tool will be a recent result from \cite{Brazitikos-Giannopoulos-Liakopoulos-2016} which is a restricted version of Meyer's dual Loomis-Whitney inequality \begin{equation}\label{eq:local-1}|K|^{n-1}\geqslant \frac{n!}{n^n}\prod_{i=1}^n|K\cap e_i^{\perp }|\end{equation} where $\{e_1,\ldots ,e_n\}$ is any orthonormal basis of ${\mathbb R}^n$ (see \cite{Meyer-1988}) and in a sense dualizes the uniform cover inequality of Bollob\'{a}s and Thomason (see \cite{Bollobas-Thomason-1995}). In order to give the precise statement, we introduce some notation. For every non-empty $\tau\subset [n]:=\{1,\ldots ,n\}$ we set $F_{\tau }={\rm span}\{e_j:j\in\tau \}$ and $E_{\tau }=F_{\tau }^{\perp }$. Given $s\geqslant 1$ and $\sigma\subseteq [n]$, following the terminology of \cite{Bollobas-Thomason-1995} we say that the (not necessarily distinct) sets $\sigma_1,\ldots ,\sigma_t\subseteq \sigma $ form an $s$-uniform cover of $\sigma $ if every $j\in \sigma $ belongs to exactly $s$ of the sets $\sigma_i$. Then, \cite[Theorem~1.3]{Brazitikos-Giannopoulos-Liakopoulos-2016} states that for any centered convex body $K$ in ${\mathbb R}^n$, for any $t\geqslant 1$ and any $s$-uniform cover $(\sigma_1,\ldots ,\sigma_t)$ of a subset $\sigma $ of $[n]$ we have \begin{equation}\label{eq:local-2}\prod_{i=1}^t|K\cap E_{\sigma_i}|\leqslant \left (\frac{c_0t}{s}\right )^{ds}|K\cap E_{\sigma }|^s|K|^{t-s},\end{equation} where $d=|\sigma |$. We will need a special case of this inequality. We consider $1\leqslant k\leqslant n-1$ and a $(k+1)$-tuple of orthonormal vectors $e_1,\ldots ,e_k,e_{k+1}:=\xi $ in ${\mathbb R}^n$. Note that the sets $\sigma_1=[k]$ and $\sigma_2=\{k+1\}$ form a $1$-uniform cover of the set $\sigma =[k+1]$. Applying \eqref{eq:local-2} with $t=2$, $s=1$ and $d=k+1$ we obtain the next lemma. \begin{lemma}\label{lem:local-1}Let $K$ be a centered convex body in ${\mathbb R}^n$. For every $1\leqslant k\leqslant n-1$, for any $E\in {\rm Gr}_{n-k}$ and any $\xi\in S^{n-1}\cap E$ we have \begin{equation}\label{eq:local-3}|K\cap E|\cdot |K\cap\xi^{\perp }|\leqslant c_0^{k+1}|K\cap E\cap\xi^{\perp }|\cdot |K|,\end{equation} where $c_0>0$ is an absolute constant. \end{lemma} Using Lemma \ref{lem:local-1} we can compare ${\rm as}(K)$ to ${\rm as}(K\cap E)$ for every $E\in {\rm Gr}_{n-k}$. We need the next well-known properties of the parameter $M$ which was defined by \eqref{eq:notation-7}. If $D$ is a symmetric convex body in ${\mathbb R}^m$ then for every $1\leqslant s\leqslant m-1$ and $F\in {\rm Gr}_s({\mathbb R}^m)$ we have that \begin{equation}\label{eq:M-1}M(D\cap F)=\int_{S_F}\|\xi \|_D\,d\sigma_F (\xi )\leqslant c_1\sqrt{m/s}\,\int_{S^{m-1}}\|\xi \|_D\,d\sigma (\xi ) =c_1\sqrt{m/s}\,M(D),\end{equation} where $c_1>0$ is an absolute constant. It is also known that \begin{equation}\label{eq:M-2}\int_{S^{m-1}}\rho_D(\theta )\,d\sigma (\theta )=\int_{S^{m-1}}\|\theta \|_D^{-1}\,d\sigma (\theta )\simeq \frac{1}{M(D)}.\end{equation} For a proof of \eqref{eq:M-1} and \eqref{eq:M-2} see \cite[Section~5.2.1]{AGA-book} and \cite[Theorem~5.8.7]{AGA-book} respectively. \begin{theorem}\label{th:BGL-1}Let $K$ be a centered convex body in ${\mathbb R}^n$. For every $1\leqslant k\leqslant n-1$ and $E\in {\rm Gr}_{n-k}$ we have \begin{equation}\label{eq:BGL-1}|K\cap E|\cdot {\rm as}(K)\leqslant c_2^k\,{\rm as}(K\cap E)\cdot |K|,\end{equation} where $c_2>0$ is an absolute constant. \end{theorem} \begin{proof}We consider an orhonormal basis $\{e_1,\ldots ,e_k\}$ of $E^{\perp }$ and any unit vector $\xi\in E$. From Lemma \ref{lem:local-1} we have \begin{equation}\label{eq:BGL-2}|K\cap E|\cdot |K\cap\xi^{\perp }|\leqslant c_0^{k+1}|K\cap E\cap\xi^{\perp }|\cdot |K|.\end{equation} Integrating \eqref{eq:BGL-2} with respect to $\xi\in S_E$ we see that \begin{align}\label{eq:BGL-3}|K\cap E|\cdot \int_{S_E}|K\cap \xi^{\perp }|d\sigma_E(\xi ) &\leqslant c_0^{k+1}\int_{S_E}|(K\cap E)\cap\xi^{\perp }|d\sigma_E(\xi )\cdot |K|\\ \nonumber &= c_0^{k+1}\,{\rm as}(K\cap E)\cdot |K|.\end{align} Applying \eqref{eq:M-2} for the symmetric convex body $IK\cap E$ we see that \begin{equation}\label{eq:BGL-4}\int_{S_E}|K\cap\xi^{\perp }|d\sigma_E (\xi ) =\int_{S_E}\rho_{IK}(\xi )\,d\sigma_E (\xi )\simeq \frac{1}{M(IK\cap E)},\end{equation} and hence, using \eqref{eq:M-1} with $m=n$ and $s=n-k$ and then applying \eqref{eq:M-2} for the body $IK$ this time, we obtain \begin{align}\int_{S_E}|K\cap\xi^{\perp }|d\sigma_E (\xi )&\geqslant \frac{c\sqrt{n-k}}{\sqrt{n}}\frac{1}{M(IK)}\simeq \frac{\sqrt{n-k}}{\sqrt{n}}\int_{S^{n-1}}\rho_{IK}(\xi )\,d\sigma (\xi )\\ \nonumber &= \frac{\sqrt{n-k}}{\sqrt{n}}\int_{S^{n-1}}|K\cap\xi^{\perp }|\,d\sigma (\xi )=\frac{\sqrt{n-k}}{\sqrt{n}}{\rm as}(K). \end{align} Therefore, \begin{equation}\label{eq:BGL-5}|K\cap E|\,{\rm as}(K)\leqslant \frac{c_1\sqrt{n}}{\sqrt{n-k}}c_0^{k+1}\,{\rm as}(K\cap E)\cdot |K|\leqslant c_2^k\,{\rm as}(K\cap E)\cdot |K|\end{equation} for every $E\in {\rm Gr}_{n-k}$. \end{proof} \medskip For the proof of Theorem \ref{th:intro-6} we use Theorem \ref{th:BGL-1} and estimates for the dual affine quermassintegrals of a centered convex body $K$: these are defined, for any $1\leqslant k\leqslant n-1$, as follows: \begin{equation}\label{eq:main-3}\tilde{R}_k(K):=\frac{1}{|K|^{n-k}}\int_{{\rm Gr}_{n-k}}|K\cap E|^n\,d\nu_{n-k}(E).\end{equation} The quantities $\tilde{R}_k(K)$ were introduced by Lutwak in \cite{Lutwak-1984} and \cite{Lutwak-1988}. More precisely, he considered the quantities $\tilde{\Phi}_k(K)$ that were introduced in \eqref{eq:intro-15}, which clearly satisfy the identity \begin{equation}\tilde{\Phi }_k(K)=\frac{\omega_n}{\omega_{n-k}}|K|^{\frac{n-k}{n}}\big [\tilde{R}_k(K)\big ]^{\frac{1}{n}}.\end{equation} Grinberg proved in \cite{Grinberg-1990} that the quantity $\tilde{R}_k(K)$ is invariant under $T\in GL(n)$: one has \begin{equation}\label{eq:main-5}\tilde{R}_k(T(K))=\tilde{R}_k(K)\end{equation} for every $T\in GL(n)$. He also proved that \begin{equation}\label{eq:main-4}\tilde{R}_k(K)\leqslant \tilde{R}_k(B_2^n):=\frac{\omega_{n-k}^n}{\omega_n^{n-k}}\leqslant e^{\frac{kn}{2}}.\end{equation} On the other hand, it was observed by Dafnis and Paouris in \cite{Dafnis-Paouris-2012} that \begin{equation}\label{eq:main-6}\tilde{R}_k(K)\geqslant \left (\frac{c_4}{L_K}\right )^{kn},\end{equation} where $c_4>0$ is an absolute constant. We will use this lower bound, which is an immediate consequence of \eqref{eq:main-5} and of the fact that if $K$ is isotropic then $|K\cap E|^{\frac{1}{k}}\geqslant\frac{c_4}{L_K}$ for every $E\in {\rm Gr}_{n-k}$ (see \cite[Proposition~5.1.15]{BGVV-book} for a proof). \medskip \noindent {\bf Proof of Theorem \ref{th:intro-6}.} Let $K$ be a centered convex body in ${\mathbb R}^n$ and fix $1\leqslant k\leqslant n-1$. From Theorem \ref{th:BGL-1} we know that for every $E\in {\rm Gr}_{n-k}$ we have \begin{equation}\label{eq:BGL-15}|K\cap E|\cdot {\rm as}(K)\leqslant c_2^k\,{\rm as}(K\cap E)\cdot |K|,\end{equation} where $c_2>0$ is an absolute constant. Therefore, \begin{equation}\label{eq:BGL-6}\max_{E\in {\rm Gr}_{n-k}}\,|K\cap E|\cdot {\rm as}(K)\leqslant c_2^k\,\max_{E\in {\rm Gr}_{n-k}}{\rm as}(K\cap E)\cdot |K|.\end{equation} Next, from \eqref{eq:main-5} we see that \begin{equation}\label{eq:BGL-8}\max_{E\in {\rm Gr}_{n-k}}|K\cap E|\geqslant \left (\int_{{\rm Gr}_{n-k}}|K\cap E|^n\,d\nu_{n-k}(E)\right )^{\frac{1}{n}}\geqslant \left (\frac{c_4}{L_K}\right )^k\,|K|^{\frac{n-k}{n}}.\end{equation} Going back to \eqref{eq:BGL-6} we see that \begin{equation}\label{eq:BGL-9}\left (\frac{c_4}{L_K}\right )^k|K|^{\frac{n-k}{n}}{\rm as}(K)\leqslant c_2^k|K|\,\max_{E\in {\rm Gr}_{n-k}}{\rm as}(K\cap E),\end{equation} and this proves Theorem \ref{th:intro-6}. $\quad \hfill \Box$ \medskip The next proposition shows that if $K$ is isotropic and if we consider the hyperplane case (where $k=1$) then the estimate of Theorem \ref{th:intro-6} is sharp: we have an asymptotic formula. \begin{proposition}\label{prop:isotropic}Let $K$ be an isotropic convex body in ${\mathbb R}^n$. Then, ${\rm as}(K)\simeq L_K^{-1}$ and ${\rm as}(K\cap\xi^{\perp })\simeq L_K^{-2}$ for all $\xi\in S^{n-1}$. In particular, \begin{equation}\label{eq:isotropic-1}{\rm as}(K)\simeq L_K\,|K|^{\frac{1}{n}}\max_{\xi\in S^{n-1}}{\rm as}(K\cap\xi^{\perp }).\end{equation} \end{proposition} \begin{proof}It is a general fact (following from \cite[Proposition~5.1.15]{BGVV-book}) that if $K$ is an isotropic convex body then, for every $E\in {\rm Gr}_{n-k}$ we have \begin{equation}\label{eq:isotropic-2}\frac{c_1}{L_K}\leqslant |K\cap E|^{\frac{1}{k}}\leqslant \frac{cL_k}{L_K}\leqslant \frac{c_2(k)}{L_K},\end{equation} where $c_1>0$ is an absolute constant and $c_2(k)$ is a positive constant depending only on $k$ (in fact, $c_2(k)\leqslant c\sqrt[4]{k}$ by Klartag's estimate on $L_k$). Applying \eqref{eq:isotropic-2} with $k=1$ we see that all hyperplane sections $K\cap\xi^{\perp }$ of $K$ have volume equal (up to an absolute constant) to $L_K^{-1}$. In particular, \begin{equation}\label{eq:isotropic-3}{\rm as}(K)=\int_{S^{n-1}}|K\cap \xi^{\perp }|\,d\sigma (\xi )\simeq L_K^{-1}.\end{equation} Applying \eqref{eq:isotropic-2} with $k=2$ we see that all $2$-codimensional sections $K\cap E$ of $K$ have volume equal (up to an absolute constant) to $L_K^{-2}$. In particular, for every $\xi\in S^{n-1}$ we get \begin{equation}\label{eq:isotropic-4}{\rm as}(K\cap \xi^{\perp })=\int_{S(\xi^{\perp })}|K\cap E_{\xi ,\theta }|\,d\sigma_{\xi }(\theta )\simeq L_K^{-2},\end{equation} where $E_{\xi ,\theta }=[{\rm span}\{\xi ,\theta\}]^{\perp }$. This shows that \begin{equation}\label{eq:isotropic-5}{\rm as}(K)\simeq L_K\,{\rm as}(K\cap\xi^{\perp })=L_K\,{\rm as}(K\cap\xi^{\perp })\,|K|^{\frac{1}{n}}\end{equation} In particular, \eqref{eq:isotropic-5} implies \eqref{eq:isotropic-1}. \end{proof} \begin{remark}\rm Proposition \ref{prop:isotropic} and the definition of $\gamma_{n,1}$ show that \begin{equation}\label{eq:worse-k-1}L_K^{-1}\simeq {\rm as}(K)\leqslant \gamma_{n,1}\,\max_{\xi\in S^{n-1}}{\rm as}(K\cap \xi^{\perp })\simeq \gamma_{n,1}L_K^{-2}\end{equation} for every isotropic convex body $K$ in ${\mathbb R}^n$. Therefore, $L_K\leqslant c\gamma_{n,1}$ for some absolute constant $c>0$, which implies that \begin{equation}\label{eq:worse-k-2}L_n\leqslant c\gamma_{n,1}.\end{equation} Note that by Theorem \ref{th:intro-6} we can then conclude that $\gamma_{n,k}\leqslant cL_n\leqslant c\gamma_{n,1}$. Finally, Theorem \ref{th:intro-6} shows that $\gamma_{n,1}\leqslant c^{\prime }L_n$. We summarize in the next proposition. \end{remark} \begin{proposition}\label{prop:worse-k}For any $1\leqslant k\leqslant n-1$ we have \begin{equation}\label{eq:worse-k-3}\gamma_{n,k}\lesssim \gamma_{n,1}\simeq L_n,\end{equation} where $c>0$ is an absolute constant. \end{proposition} Proposition \ref{prop:worse-k} shows that a positive answer to Question \ref{question-intro} (actually, in the case $k=1$) is equivalent to the uniform boundedness of the isotropic constants of all convex bodies in all dimensions (this is exactly the hyperplane conjecture). It also shows that the question becomes ``easier" when the codimension $k$ increases, in the sense that $\gamma_{n,k}\lesssim \gamma_{n,1}$. In fact, we can show that if $k$ is proportional to $n$ then $\gamma_{n,k}$ is bounded (this is precisely the content of Theorem \ref{th:intro-7}): \medskip \begin{theorem}\label{th:gamma-DP}For any $1\leqslant k\leqslant n-1$ we have \begin{equation}\label{eq:worse-k-4}\gamma_{n,k}\leqslant c\sqrt{n/k}\,[\log (en/k)]^{\frac{3}{2}},\end{equation} where $c>0$ is an absolute constant. \end{theorem} \begin{proof}We repeat the proof of Theorem \ref{th:intro-6} using the estimate (see \cite[Theorem~1.3]{Dafnis-Paouris-2012}) \begin{equation}\label{eq:main-7}\tilde{R}_k(K)\geqslant \left (\frac{c_5}{\sqrt{n/k}\,[\log (en/k)]^{\frac{3}{2}}}\right )^{kn}\end{equation} instead of \eqref{eq:main-6}. \end{proof} \begin{remark}\label{rem:dual-affine}\rm Let $\alpha_{n,k}$ be the largest constant $\alpha >0$ with the property that $\tilde{R}_k(K)\geqslant \alpha^{kn}$ for every centered convex body $K$ in ${\mathbb R}^n$. Repeating the proof of Theorem \ref{th:intro-6} or Theorem \ref{th:gamma-DP} we see that \begin{equation}\label{eq:worse-k-5}\gamma_{n,k}\leqslant \frac{c_1}{\alpha_{n,k}},\end{equation} where $c_1>0$ is an absolute constant. In particular, \begin{equation}\gamma_{n,k}\lesssim \gamma_{n,1}\simeq L_n\lesssim \alpha_{n,1}^{-1}.\end{equation} On the other hand, \eqref{eq:main-6} shows that $\alpha_{n,k}\geqslant c/L_n$ for all $1\leqslant k\leqslant n-1$, and hence $\alpha_{n,1}^{-1}\lesssim L_n$. Therefore, \begin{equation}\gamma_{n,1}\simeq L_n\simeq \alpha_{n,1}^{-1}.\end{equation} In other words, the question whether \begin{equation}\label{eq:worse-k-6}\tilde{R}_k(K)\geqslant c^{kn}\end{equation} for all $1\leqslant k\leqslant n-1$ which is studied in \cite{Dafnis-Paouris-2012} (see also \cite[Section~9.4]{Gardner-book}) is equivalent to the hyperplane conjecture and to Question \ref{question-intro}. \end{remark} \section{Reverse inequalities in the classical positions}\label{positions} Next, we pass to estimates for the mean value of the average section functional of hyperplane sections of $K$. We start by expressing ${\rm as}(K)$ in terms of dual mixed volumes. Note that by \eqref{eq:ovr-7} we have \begin{equation}\label{eq:DMX-3}{\rm as}(K)=\omega_{n-1}\int_{S^{n-1}}\rho_K^{n-1}(\theta )\,d\sigma (\theta )=\frac{\omega_{n-1}}{\omega_n}\tilde{V}(K,\ldots ,K,B_2^n), \end{equation} and using \eqref{eq:ovr-8} we see that \begin{equation}\label{eq:DMX-4}\int_{G_{n,,n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E)=\omega_{n-k-1}\int_{S^{n-1}}\rho_K^{n-k-1}(\theta )\,d\sigma (\theta ) =\frac{\omega_{n-k-1}}{\omega_n}\tilde{V}(K\,[n-k-1],B_2^n\,[k+1]),\end{equation} where $A\,[s]$ means $A,\ldots ,A$ repeated $s$-times. \begin{theorem}\label{th:BGL-2}Let $K$ be a centered convex body in ${\mathbb R}^n$. Then, \begin{equation}\label{eq:BGL-10}{\rm as}(K)^{k+1}\leqslant c^k\,|K|^k\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k} (E)\end{equation} and \begin{equation}\label{eq:BGL-10b}\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k} (E)\leqslant c^k\,{\rm as}(K)^{\frac{n-k-1}{n-1}}, \end{equation} where $c>0$ is an absolute constant. \end{theorem} \begin{proof}From H\"{o}lder's inequality we see that \begin{equation}\left (\int_{S^{n-1}}\rho_K^{n-1}(\theta )\,d\sigma (\theta )\right )^{k+1}\leqslant \left (\int_{S^{n-1}}\rho_K^n(\theta )\,d\sigma (\theta )\right )^k\left (\int_{S^{n-1}}\rho_K^{n-k-1}(\theta )\,d\sigma (\theta )\right ),\end{equation} which can be equivalently written as \begin{equation}\label{eq:BGL-11}\tilde{V}(K,\ldots ,K,B_2^n)^{k+1}\leqslant |K|^k\,\tilde{V}(K\,[n-k-1],B_2^n\,[k+1]).\end{equation} Taking into account \eqref{eq:DMX-3} and \eqref{eq:DMX-4} we rewrite \eqref{eq:BGL-11} in the form \begin{equation}\label{eq:BGL-12}{\rm as}(K)^{k+1}\leqslant |K|^k\,\frac{\omega_{n-1}^{k+1}}{\omega_n^k\omega_{n-k-1}}\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k} (E).\end{equation} A simple computation shows that $\frac{\omega_{n-1}^{k+1}}{\omega_n^k\omega_{n-k-1}}<c^k$ for an absolute constant $c>0$, and \eqref{eq:BGL-10} follows. On the other hand, by H\"{o}lder's inequality, \begin{align}\label{eq:BGL-13}&\frac{1}{\omega_{n-k-1}}\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k} (E) =\int_{S^{n-1}}\rho_K^{n-k-1}(\theta )\,d\sigma (\theta )\\ \nonumber &\hspace*{1.5cm}\leqslant \left (\int_{S^{n-1}}\rho_K^{n-1}(\theta )\,d\sigma (\theta )\right )^{\frac{n-k-1}{n-1}}= \left (\frac{{\rm as}(K)}{\omega_{n-1}}\right )^{\frac{n-k-1}{n-1}}, \end{align} therefore \begin{equation}\label{eq:BGL-14}\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k} (E)\leqslant \varrho_{n,k} \,{\rm as}(K)^{\frac{n-k-1}{n-1}},\end{equation} where $\varrho_{n,k}=\omega_{n-k-1}\cdot \omega_{n-1}^{-\frac{n-k-1}{n-1}}\leqslant c^k$ for an absolute constant $c>0$, which gives \eqref{eq:BGL-10b}. \end{proof} \medskip Let $K$ be a convex body in ${\mathbb R}^n$ with $0\in {\rm int}(K)$. Recall that the radius $R(K)$ of $K$ is the smallest $R>0$ for which $K\subseteq RB_2^n$. Using the monotonicity and homogeneity of dual mixed volumes and \eqref{eq:DMX-3} we may write \begin{align}\label{eq:as-radius-0}{\rm as}(K) &= \frac{\omega_{n-1}}{\omega_n}\tilde{V}(K,\ldots ,K,B_2^n)\geqslant \frac{\omega_{n-1}}{\omega_n R(K)}\tilde{V}(K,\ldots ,K,R(K)B_2^n)\\ \nonumber &\geqslant \frac{\omega_{n-1}}{\omega_n R(K)}\tilde{V}(K,\ldots ,K,K)=\frac{\omega_{n-1}}{\omega_n R(K)}\,|K|. \end{align} In this way we obtain the following general lower bound for ${\rm as}(K)$. \begin{lemma}\label{lem:as-radius-2}Let $K$ be a centered convex body in ${\mathbb R}^n$. If we define $p(K)=R(K)/|K|^{\frac{1}{n}}$ then \begin{equation}\label{eq:as-radius-3}\frac{c\sqrt{n}}{p(K)}\leqslant \frac{{\rm as}(K)}{|K|^{\frac{n-1}{n}}},\end{equation} where $c>0$ is an absolute constant. \end{lemma} \begin{proof}From \eqref{eq:as-radius-0} we see that \begin{equation}\label{eq:as-radius-5}\frac{{\rm as}(K)}{|K|^{\frac{n-1}{n}}}\geqslant \frac{\omega_{n-1}}{\omega_n R(K)}|K|^{1/n}\geqslant \frac{c\sqrt{n}}{R(K)}|K|^{1/n},\end{equation} and the lemma follows by the definition of $p(K)$. \end{proof} \medskip Going back to Theorem \ref{th:BGL-2} we immediately get the following. \begin{theorem}\label{th:as-radius-3}Let $K$ be a centered convex body in ${\mathbb R}^n$. Then, for every $1\leqslant k\leqslant n-1$ we have that \begin{equation}\label{eq:as-radius-6}\left (\frac{c_1\sqrt{n}}{p(K)}\right )^k\,{\rm as}(K)\leqslant |K|^{\frac{k}{n}}\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E)\leqslant \left (\frac{c_2p(K)}{\sqrt{n}}\right )^{\frac{k}{n-1}}\,{\rm as}(K),\end{equation} where $c_1, c_2>0$ are absolute constants. \end{theorem} \begin{proof}The left hand side inequality follows from Lemma \ref{lem:as-radius-2} and \eqref{eq:BGL-10}. We have $$|K|^{\frac{k(n-1)}{n}}\left (\frac{c_1\sqrt{n}}{p(K)}\right )^k{\rm as}(K)\leqslant {\rm as}(K)^{k+1}\leqslant c^k|K|^k\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E),$$ which implies that $$\left (\frac{c_1\sqrt{n}}{cp(K)}\right )^k\,{\rm as}(K)\leqslant |K|^{\frac{k}{n}}\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E).$$ Next, we observe that \begin{equation}\label{eq:as-radius-7}|K|^{\frac{k}{n}}=\left (|K|^{\frac{n-1}{n}}\right )^{\frac{k}{n-1}}\leqslant \left (\frac{p(K){\rm as}(K)}{c\sqrt{n}}\right )^{\frac{k}{n-1}},\end{equation} which implies that \begin{equation}\label{eq:as-radius-8}|K|^{\frac{k}{n}}\,{\rm as}(K)^{\frac{n-k-1}{n-1}}\leqslant \left (\frac{c_2p(K)}{\sqrt{n}}\right )^{\frac{k}{n-1}}{\rm as}(K).\end{equation} Then, the right hand side inequality of \eqref{eq:as-radius-6} follows from \eqref{eq:BGL-10b} in Theorem \ref{th:BGL-2}. \end{proof} \begin{remark}\label{rem:classical}\rm We will discuss the estimates that one can get from Theorem \ref{th:as-radius-3} if the centered convex body $K$ in ${\mathbb R}^n$ is assumed to be in some of the classical positions; we introduce these below. For a detailed presentation and references see \cite{AGA-book}. \begin{enumerate} \item[(i)] We say that $K$ is in minimal mean width position if $w(K)\leqslant w(T(K))$ for every $T\in SL(n)$. It was proved by V.~Milman and the second named author that $K$ has minimal mean width if and only if \begin{equation}\label{eq:notation-17}w(K)=n\int_{S^{n-1}}h_K(\theta )\langle \xi ,\theta\rangle^2d\sigma (\theta )\end{equation} for every $\xi\in S^{n-1}$. From results of Figiel-Tomczak, Lewis and Pisier (see \cite[Chapter 6]{AGA-book}) we know that if a convex body $K$ in ${\mathbb R}^n$ has minimal mean width then $w(K)\leqslant c|K|^{\frac{1}{n}}\sqrt{n}\log n$. From the general fact that $R(K)\leqslant c\sqrt{n}w(K)$ for every centered convex body, we conclude that $R(K)\leqslant c|K|^{\frac{1}{n}}n\log n$ in the minimal mean width position. \item[(ii)] We say that $K$ is in John's position if the ellipsoid of maximal volume inscribed in $K$ is a multiple of the Euclidean unit ball $B_2^n$ and that $K$ is in L\"{o}wner's position if the ellipsoid of minimal volume containing $K$ is a multiple of the Euclidean unit ball $B_2^n$. One can check that this holds true if and only if $K^{\circ }$ is in John's position. The volume ratio of a centered convex body $K$ in ${\mathbb R}^n$ is the quantity \begin{equation}\label{eq:notation-18}{\rm vr}(K)=\inf\left\{ \left(\frac{|K|}{|{\cal E}|}\right )^{\frac{1}{n}}:{\cal E}\;\hbox{is an ellipsoid and}\; {\cal E}\subseteq K\right\}.\end{equation} The outer volume ratio of a centered convex body $K$ in ${\mathbb R}^n$ is the quantity ${\rm ovr}(K)={\rm vr}(K^{\circ })$. K.~Ball proved in \cite{Ball-reverse} that ${\rm vr}(K)\leqslant {\rm vr}(C_n)\simeq \sqrt{n}$ in the symmetric case and ${\rm vr}(K)\leqslant {\rm vr}(\Delta_n)\simeq \sqrt{n}$ in the not necessarily symmetric case, where $C_n=[-1,1]^n$ and $\Delta_n$ is a regular simplex in ${\mathbb R}^n$. Assume that $K$ is in John's position. Then, from a theorem of Barthe \cite{Barthe-width} we know that if $\Delta_n$ is the regular simplex whose maximal volume ellipsoid is $B_2^n$ and $rB_2^n$ is the maximal volume ellipsoid of $K$ we have $w(r^{-1}K)\leqslant w(\Delta_n)\leqslant c\sqrt{\log n}$. Since $|K|^{1/n}\geqslant r|B_2^n|^{1/n}\geqslant cr/\sqrt{n}$, we get \begin{equation}R(K)\leqslant c\sqrt{n}w(K)=cr\sqrt{n}w(r^{-1}K)\leqslant cr\sqrt{n\log n}\leqslant c^{\prime }|K|^{\frac{1}{n}}n\sqrt{\log n}.\end{equation} Next, assume that $K$ is in L\"{o}wner's position; we know that $R(K)B_2^n$ is the minimal volume ellipsoid of $K$, and hence \begin{equation}R(K)|B_2^n|^{1/n}=|K|^{\frac{1}{n}}{\rm ovr}(K)=|K|^{\frac{1}{n}}{\rm vr}(K^{\circ })\leqslant c\sqrt{n}|K|^{\frac{1}{n}},\end{equation} which implies that $R(K)\leqslant cn|K|^{\frac{1}{n}}$. \item[(iii)] We say that $K$ has minimal surface area if $S(K)\leqslant S(T(K))$ for every $T\in SL(n)$. Recall that the area measure $\sigma_K$ of $K$ is the Borel measure on $S^{n-1}$ defined by \begin{equation}\sigma_K(A)=\lambda (\{x\in {\rm bd}(K):\;{\rm the}\;{\rm outer} \;{\rm normal}\;{\rm to}\;K\;{\rm at}\;x\;{\rm belongs}\; {\rm to}\;A\}),\end{equation} where $\lambda $ is the usual surface measure on $K$. Petty proved in \cite{Petty-1961} that $K$ has minimal surface area if and only if $\sigma_K$ satisfies the isotropic condition \begin{equation}\label{eq:notation-19}S(K)=n\int_{S^{n-1}}\langle \xi ,\theta\rangle^2d\sigma_K (\theta )\end{equation} for every $\xi\in S^{n-1}$. It is known that if $K$ has minimal surface area then $w(K)\leqslant cn|K|^{\frac{1}{n}}$ (this was observed by Markessinis, Paouris and Saroglou in \cite{Markessinis-Paouris-Saroglou-2012}). Therefore, $R(K)\leqslant cn^{\frac{3}{2}}|K|^{\frac{1}{n}}$. \item[(iv)] Finally, if $K$ is in the isotropic position then we know that $R(K)\leqslant |K|^{\frac{1}{n}}(n+1)L_K$. This estimate is due to Kannan, Lov\'{a}sz and Simonovits (the asymptotically sharp bound $R(K)\leqslant cnL_K\,|K|^{\frac{1}{n}}$ can be obtained with an elementary argument). \end{enumerate} \end{remark} Since $R(K)$ is polynomial in $n$ for all the classical positions of a convex body $K$, from the right hand side inequality \eqref{eq:as-radius-6} of Theorem \ref{th:as-radius-3} we obtain the next result. \begin{theorem}\label{th:as-radius-4}Let $K$ be a centered convex body in ${\mathbb R}^n$. If $K$ is in any of the classical positions that we discussed in Remark $\ref{rem:classical}$, then \begin{equation}\label{eq:as-radius-9}|K|^{\frac{k}{n}}\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E)\leqslant C^k\,{\rm as}(K)\end{equation} for all $1\leqslant k\leqslant n-1$, where $C>0$ is an absolute constant. \end{theorem} \begin{remark}\rm Similarly, from the left hand side inequality \eqref{eq:as-radius-6} of Theorem \ref{th:as-radius-3} we see that if $K$ is in some of the classical positions that we discussed in Remark \ref{rem:classical} then \begin{equation}\label{eq:as-radius-10}c_n^{-k}\,{\rm as}(K)\leqslant |K|^{\frac{k}{n}}\,\int_{{\rm Gr}_{n-k}}{\rm as}(K\cap E)\,d\nu_{n-k}(E) \end{equation} for every $1\leqslant k\leqslant n-1$, where $c_n\simeq\sqrt{n}$ if $K$ is in L\"{o}wner's position, $c_n\simeq\sqrt{n\log n}$ if $K$ is in John's position, $c_n\simeq\sqrt{n}(\log n)$ if $K$ is in the minimal mean width position, $c_n\simeq n$ if $K$ is in the minimal surface area position, and $c_n\simeq\sqrt{n}L_K$ if $K$ is in the isotropic position. \end{remark} \bigbreak \bigskip \bigskip \noindent {\bf Acknowledgements.} The first named author acknowledges support from Onassis Foundation. The fourth named author was partially supported by the US National Science Foundation grant DMS-1265155. \bigskip \bigskip \footnotesize \bibliographystyle{amsplain}
67d997f5df34a2ef9e1f47687b2c07486d90d0af
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Hyperspectral imaging techniques have been widely used for a variety of applications such as mineral mapping, surveillance, and environmental monitoring. With its rich spectral information, hyperspectral data provide unique spectral patterns of the materials present in the scene. This capability makes hyperspectral data especially suitable for ground cover classification problems. Although hyperspectral data forms a high dimensional feature space, the information is often effectively represented in a much lower dimensional subspace --- hence, finding an appropriate subspace is crucial for effective hyperspectral data classification. \textcolor{blue}{}. Several linear and non-linear dimensionality reduction methods have been proposed in literature to address this issue \cite{HN2004,roweis2000nonlinear,tenenbaum2000global,belkin2001laplacian,Sug2007,gao2015discriminative,LSX2010,raducanu2012supervised,zhang2012graph,cui2012novel,hou2010multiple,song2008unified}. The objective function of such dimensionality reduction methods are typically utilize Euclidean distance information. Different from such Euclidean-based dimensionality reduction methods, angle or correlation-based dimensionality reduction methods have been proposed in literature to exploit specific benefits of angular information --- something that is particularly relevant for hyperspectral data. In \cite{ma2007discriminant}, correlation discriminant analysis (CDA) is proposed to transform the data so that the ratio of between-class and within-class sample correlation is maximized. In \cite{cui15angular}, the authors propose several supervised spectral angle based dimensionality reduction methods for robust hyperspectral data classification. In typical remote sensing data classification applications, the process of collecting labeled training data is often very time consuming and expensive. Unlabeled data on the other hand is easily available and hence unsupervised dimensionality reduction methods that effectively learn the most appropriate subspace can hence be readily utilized. With that in mind, we propose an unsupervised counterpart of our recently proposed supervised subspace algorithm, the angular discriminant Analysis (ADA). This unsupervised counterpart is referred to as local similarity preserving projection (LSPP) in this paper. Unlike ADA which requires labeled training samples, LSPP does not require labeled training samples to learn the projection matrix. Additionally, it is well-known that utilizing spatial information in hyperspectral data can dramatically improve the classification accuracies because any such method accounts for the spatial variability of spectral content in local spatial neighborhoods. This follows from the observation that spatially neighboring pixels are highly likely to belong to the same class and have similar spectral characteristics. To incorporate such spatial information of hyperspectral data into our unsupervised projection, we develop a spatial information driven variant of LSPP (SLSPP) which effectively uses the spatial contextual information around each pixel in hyperspectral images to learn the \emph{optimal projection}. Recently, sparse representation-based classification (SRC) has been proposed for face recognition \cite{WYG2009} and achieved a great amount of attention in various other applications such as digit recognition \cite{labusch2008simple}, speech recognition \cite{gemmeke2011exemplar}, gesture recognition \cite{zhou2013kernel}, vehicle classification \cite{mei2011robust} as well as remote sensing data classification \cite{MStgars2013,CNT2011,srinivas2011exploiting}. The central idea of SRC is to represent a test sample as a sparse linear combination of atoms in a dictionary which is formed by all of the available training samples. The class label of the test sample is assigned to a class whose reconstruction error is the lowest. In \cite{CNT2011}, the authors propose a join sparsity model to incorporate the contextual information of test samples to improve the classification performance of SRC. However, the contextual information of training samples have not been used. Besides the proposed dimensionality reduction methods, we also propose a sparse representation based classifier which takes into account the spatial information for both training and test samples in this work. Experimental results through two real-world hyperspectral datasets demonstrate that our proposed methods outperform other existing approaches. The rest of the paper is organized as follows. In Sec.~\ref{sec:dim_tra}, we briefly introduce the recently developed supervised dimensionality reduction methods (ADA and LADA), and Sec.~\ref{sec:dim_pro1} and Sec.~\ref{sec:dim_pro2} describe the proposed LSPP and SLSPP dimensionality reduction methods respectively. Sec.~\ref{sec:classifier} describes the proposed classification method. In Sec.~\ref{sec:data}, the two practical hyperspectral benchmarking datasets employed to validate the proposed method are described, and the experimental results illustrating the benefits of the proposed work is described in Sec.~\ref{sec:exp}. \section{Spectral Angle-Based Dimensionality Reduction} \subsection{ADA and LADA} \label{sec:dim_tra} Recently, supervised angular discriminant analysis (ADA) was proposed in \cite{cui15angular}. ADA finds a subspace where the angular separation of between-class samples is maximized while the angular separation of within-class samples is simultaneously minimized. It is mainly used to improve the classification performance of NN with cosine angle distance and SRC with OMP as the recovery method. Let us define $\{\boldsymbol{x}_{i} \in \mathbb{R}^{d}, y_i \in \{1,2, \ldots, c\}, i=1,2, \ldots, n\}$ to be the $d$-dimensional $i$-th training sample with an associated class label $y_i$, where $c$ is the number of classes and $n$ is the total number of training samples. $n = \sum_{l=1}^{c}{n_{l}}$ where $n_{l}$ denotes the number of training samples from class $l$. Let $\boldsymbol{\mathit{T}} \in \mathbb{R}^{d \times r}$ be the projection matrix, where $r$ denotes the reduced dimensionality. The projection $\mathit{T}_{\textit{ADA}}$ in ADA can obtained by solving \begin{align} \boldsymbol{\mathit{T}}_{\textit{ADA}} &\approx \operatornamewithlimits{argmin}_{\boldsymbol{\mathit{T}}\in \mathbb{R}^{d \times r}}\left[\operatorname{tr}\big((\boldsymbol{\mathit{T}}^{t}\boldsymbol{\mathit{O}}^{(\text{w})}\boldsymbol{\mathit{T}})^{-1}\boldsymbol{\mathit{T}}^{t}\boldsymbol{\mathit{O}}^{(\text{b})}\boldsymbol{\mathit{T}}\big)\right]. \label{eq:ada} \end{align} where $\boldsymbol{\mathit{O}}^{(\text{w})}$ and $\boldsymbol{\mathit{O}}^{(\text{b})}$ are defined as \begin{align} \boldsymbol{\mathit{O}}^{(\text{w})} &= \sum^{c}_{l=1}\sum^{}_{i:y_i=l} \boldsymbol{\mu}_{l}\boldsymbol{x}_{i}^{t}, \label{eq:Ow} \\ \boldsymbol{\mathit{O}}^{(\text{b})} &= \sum^{c}_{l=1}n_l\boldsymbol{\mu}\boldsymbol{\mu}_{l}^{t}. \label{eq:Ob} \end{align} The problem in \eqref{eq:ada} can be converted to a generalized eigenvalue problem. The authors in \cite{cui15angular} also extend ADA into its localized variant named local angular discriminant analysis (LADA) by incorporating a locality preserving property through an affinity matrix. LADA was shown to handle more complex, potentially multimodal data distributions. The projection matrix $\mathit{T}_{\textit{LADA}}$ of LADA can be obtained by solving \begin{align} \boldsymbol{\mathit{T}}_{\textit{LADA}} &= \operatornamewithlimits{argmin}_{\boldsymbol{\mathit{T}}\in \mathbb{R}^{d \times r}}\left[\operatorname{tr}\big((\boldsymbol{\mathit{T}}^{t}\boldsymbol{\mathit{O}}^{(\text{lw})}\boldsymbol{\mathit{T}})^{-1}\boldsymbol{\mathit{T}}^{t}\boldsymbol{\mathit{O}}^{(\text{lb})}\boldsymbol{\mathit{T}}\big)\right]. \label{eq:lspdr} \end{align} The within-class and between-class outer product matrices of LADA are defined as \begin{align} \boldsymbol{\mathit{O}}^{(\text{lw})} &= \sum_{i,j=1}^{n}\mathit{W}_{ij}^{(\text{lw})}\boldsymbol{x}_{i}\boldsymbol{x}_{j}^{t}, \label{eq:lada_ow} \\ \boldsymbol{\mathit{O}}^{(\text{lb})} &= \sum_{i,j=1}^{n}\mathit{W}_{ij}^{(\text{lb})}\boldsymbol{x}_{i}\boldsymbol{x}_{j}^{t}, \label{eq:lada_ob} \end{align} where the weight matrices are defined based on heat kernel $\mathit{A}_{ij}$ as \begin{align} \mathit{W}^{(\text{lw})}_{ij}&= \begin{cases} \mathit{A}_{ij}/n_l, & \text{if $y_i,y_j=l$}, \\ 0, & \text{if $y_i\ne y_j$}, \end{cases} \label{eq:lada_w} \\ \mathit{W}^{(\text{lb})}_{ij}&= \begin{cases} \mathit{A}_{ij}(1/n-1/n_l), & \text{if $y_i,y_j=l$}, \\ 1/n, & \text{if $y_i\ne y_j$}. \end{cases} \label{eq:lada_b} \end{align} \subsection{LSPP} \label{sec:dim_pro1} In this paper, we seek to make two related contributions within the context of angular discriminant analysis --- (1) developing an unsupervised approach to spectral angle based subspace learning, where local spectral angles are preserved following this unsupervised project, and (2) developing a projection that incorporates spatial information when learning such an \emph{optimal} projection. We first form a unsupervised version of ADA which we refer to as local similarity preserving projection (LSPP). It seeks a lower dimensional space where the correlation or angular relationship {between samples that are neighbors in the feature space} are preserved. We can also think of it as an angular equivalent of the commonly employed locality preserving projection (LPP) \cite{HN2004}. Let $\textit{x}_i$ be the $i$-th training sample and $\textit{P}$ be the $d \times r$ projection matrix, where $r$ is the reduced dimensionality. The objective function of LSPP can be simplified as \begin{align} \textit{I} &= \sum_{ij}\textit{W}_{ij}(\textit{P}^t\textit{x}_i)^t(\textit{P}^t\textit{x}_j) \nonumber \\ &= \sum_{ij}\operatorname{\emph{tr}}\big[\textit{W}_{ij}(\textit{P}^t\textit{x}_i)^t(\textit{P}^t\textit{x}_j)\big] \nonumber \\ &= \sum_{ij}\operatorname{\emph{tr}}\big[\textit{W}_{ij}\textit{P}^t\textit{x}_j(\textit{P}^t\textit{x}_i)^t \big] \nonumber \\ &= \sum_{ij}\operatorname{\emph{tr}}\big[\textit{P}^t\textit{W}_{ij}\textit{x}_i\textit{x}_j^t\textit{P}\big] \nonumber \\ &= \operatorname{\emph{tr}}\big[\textit{P}^t\textit{X}\textit{W}\textit{X}^t\textit{P}\big] \end{align} The heat kernel $\textit{W}_{ij}\in [0,1]$ between $\textit{x}_i$ and $\textit{x}_j$ is defined as \begin{equation} \textit{W}_{ij}=\exp\left(-\frac{\Vert \textit{x}_i-\textit{x}_j\Vert^2}{\sigma}\right) , \label{eq:aff1} \end{equation} where $\sigma$ is the parameter in the heat kernel. We impose a constraint ($\textit{P}^t\textit{X}\textit{D}\textit{X}^t\textit{P}=1$ where $D_{ii}=\sum_j\textit{W}_{ij}$) {to avoid biases caused by different samples}. {The larger the value $D_{ii}$ which is corresponding to $i$-th training sample, the more important $i$-th training sample is.} The final objective function is defined as \begin{equation} \operatornamewithlimits{argmax}_{\textit{P}}{\ \operatorname{\emph{tr}}\big[\textit{P}^t\textit{X}\textit{W}\textit{X}^t\textit{P}}\big] \qquad \operatorname{s. t.} \quad \textit{P}^t\textit{X}\textit{D}\textit{X}^t\textit{P}=I \label{eq:obj1} \end{equation} The problem in \ref{eq:obj1} can be solved as a generalized eigenvalue problem as \begin{equation} \textit{X}\textit{W}\textit{X}^t\textit{P} = \lambda\textit{X}\textit{D}\textit{X}^t\textit{P} \label{eq:eig} \end{equation} The projection matrix $\textit{P}$ are the eigenvectors corresponding to the $r$ largest eigenvalues. \subsection{SLSPP} \label{sec:dim_pro2} It is well understood from many recent works \cite{TBC2009,CNT2011,JBP2012,cui2013locality}, that by aking into account the spatial neighborhood information, hyperspectral image classification accuracy can be significantly increased. This is based on the observation that spatially neighboring samples in hyperspectral data often consist of similar materials,and hence they are spectrally correlated. In order to utilize spatially neighboring samples in a lower-dimensional subspace, the spatial neighborhood relationship of hyperspectral data should be preserved. To address this problem, we propose a spatial variant of LSPP (SLSPP) in this work to further improve the classification accuracies. Let $\{\textit{z}_{k}, k \in \Omega_i\}$ be the spatial neighborhood samples around a training sample $\textit{x}_{i}$, then the objective function of SLSPP can be reduced to \begin{align} \textit{I} &= \sum_{i}\sum_{k \in \Omega_i} \textit{W}_{ik}(\textit{P}^t\textit{x}_i)^t(\textit{P}^t\textit{z}_{k}) \nonumber \\ &= \sum_{i}\sum_{k \in \Omega_i}\operatorname{\emph{tr}}\big[\textit{W}_{ik}(\textit{P}^t\textit{x}_i)^t(\textit{P}^t\textit{z}_{k})\big] \nonumber \\ &= \sum_{i}\sum_{k \in \Omega_i}\operatorname{\emph{tr}}\big[\textit{W}_{ik}\textit{P}^t\textit{z}_{k}(\textit{P}^t\textit{x}_i)^t\big] \nonumber \\ &= \sum_{i}\sum_{k \in \Omega_i}\operatorname{\emph{tr}}\big[\textit{P}^t\textit{W}_{ik}\textit{z}_{k}\textit{x}_i^t\textit{P}\big] \nonumber \\ &= \operatorname{\emph{tr}}\big[\textit{P}^tM\textit{P}\big] \end{align} where $M = \sum_{i}\sum_{k \in \Omega_i}\textit{W}_{ik}\textit{z}_{k}\textit{x}_i^t$. The final objective function is defined as \begin{equation} \operatornamewithlimits{argmax}_{\textit{P}}{\ \operatorname{\emph{tr}}\big[\textit{P}^tM\textit{P}}\big] \qquad \operatorname{s. t.} \quad \textit{P}^t\textit{P}=I \label{eq:obj2} \end{equation} Similar to LSPP, the projection matrix $\textit{P}$ are the eigenvectors corresponding to the $r$ largest eigenvalues. Note that both LSPP and SLSPP are unsupervised projections in the sense that they do not require labels when learning the projections. We also note that a projection such as SLSPP is particularly beneficial when the backend classifier utilizes the spatial structure in the spatially neighboring pixels for each test pixel during classification. Towards that end, we next propose a modified sparse representation based classifier that utilizes sparse representations of the entire spatial neighborhood of a test pixel simultaneously when making a decision. \section{Sparse Representation-Based Classification via SBOMP} \label{sec:classifier} \subsection{SBOMP} The simultaneous orthogonal matching pursuit (SOMP) \cite{tropp2006algorithms} and block orthogonal matching pursuit (BOMP) \cite{eldar2010block} are all variants of orthogonal matching pursuit (OMP) that explore the \emph{block structure} of test samples and training samples respectively. In this work, we effectively combine these two recovery methods and propose SBOMP to explore the block structure of both training and test samples simultaneously. SBOMP is illustrated in Algorithm~\ref{sbomp}. Let $x_t$ be a test sample. Assume $\textit{A}_{i}$ contains spatial neighborhood samples around $x_i$ (inclusive of $x_i$), $S$ contains the spatial neighborhood samples of $x_t$ (inclusive of $x_t$) and $K$ is the sparsity level. {SBOMP estimates the coefficient $\hat{\textit{C}}$ based on the $K$ mostly correlated spatial training samples in $\textit{A}$}. \begin{algorithm}[htbp] \caption{SBOMP} \label{sbomp} \begin{algorithmic}[1] \STATE \textbf{Input:} A spectral-spatial training data $\textit{A} = \{\textit{A}_{i}\}_{i=1}^{n}$, test data $\textit{S}$ and row sparsity level $K$. \\\hrulefill \STATE Initialize $\textit{R}^{0} = \textit{S}$, $\Lambda^{0} = \emptyset$, and the iteration counter $m = 1$. \WHILE{$m \le K$} \STATE Update the support set $\Lambda^{m} = \Lambda^{m-1} \cup \lambda$ by solving \begin{equation} \lambda = \operatornamewithlimits{argmax}_{i=1,2,\ldots, n}\|\textit{A}_i^{t}\textit{R}^{m-1}\|_{2,1}. \nonumber \end{equation} \STATE Derive the coefficient matrix $C^{m}$ based on \begin{equation*} \textit{C}^{m} = \left(\textit{A}^{t}_{\Lambda^{m}}\textit{A}_{\Lambda^{m}}\right)^{-1}\textit{A}^{t}_{\Lambda^{m}}S \end{equation*} \STATE Update the residual matrix $\textit{R}^{m}$ \begin{equation*} \textit{R}^{m} = \textit{S}-\textit{A}_{\Lambda^{m}}\textit{C}^{m} \end{equation*} \STATE $m \leftarrow m + 1$ \ENDWHILE \\\hrulefill \STATE \textbf{Output:} Coefficient matrix $\hat{\textit{C}} = \textit{C}^{m-1}$. \end{algorithmic} \end{algorithm} \subsection{Classification via SBOMP} The classification method employed after SLSPP is SBOMP-based Classification (SBOMP-C), as described in Algorithm~\ref{sbompc}. Since SLSPP can preserve the spatial neighboring samples for both training and test samples in a lower-dimensional subspace, by using SBOMP-C, we can exploit the block structure relationship effectively in the spatial domain between training and test samples. The block diagram of proposed methods are illustrated in Fig.~\ref{fig:diag}. \begin{algorithm}[htbp] \caption{SBOMP-C} \label{sbompc} \begin{algorithmic}[1] \STATE \textbf{Input:} A spectral-spatial training data $\textit{A} = \{\textit{A}_{i}\}_{i=1}^{n}$, test data $\textit{S}$ and row sparsity level $K$. \\\hrulefill \STATE Calculate row-sparsity coefficient $\hat{\textit{C}}$ based on \begin{equation*} \hat{\textit{C}} = \text{SBOMP} \ (\textit{A}, S, K) \end{equation*} \STATE Calculate residuals for each class \begin{equation*} \textit{r}_{k}(\textit{S}) = \|\textit{S}-\textit{A}\delta_{k}(\hat{\textit{C}})\|_2, \quad k = 1,2, \ldots, c \end{equation*} \STATE Determine the class label of $\textit{S}$ based on \begin{equation*} \omega = \operatornamewithlimits{argmin}_{k = 1,2,\ldots, c}(\textit{r}_{k}(\textit{S})). \nonumber \end{equation*} \\\hrulefill \STATE \textbf{Output:} A class label $\omega$. \end{algorithmic} \end{algorithm} \begin{figure*}[htbp] \centering \begin{tabular}{c} \includegraphics[width=13cm]{figure1_eps-eps-converted-to.pdf} \\ \end{tabular} \caption{Block diagram of the proposed work.} \label{fig:diag} \end{figure*} \section{Experimental Hyperspectral Data} \label{sec:data} We validate the proposed approaches with two hyperspectral imagery datasets --- a standard benchmarking dataset from the University of Pavia (an urban ground cover classification problem), and a hyperspectral imagery data collected by our lab to study wetland species composition. \subsection{University of Pavia Data} The first experimental hyperspectral dataset employed was collected using the Reflective Optics System Imaging Spectrometer (ROSIS) sensor \cite{Gam2004}. This image, covering the University of Pavia, Italy, has 103 spectral bands with a spatial coverage of $610 \times 340$ pixels, and 9 classes of interests are considered in this dataset. A three-band true color image and its ground-truth are shown in Figure \ref{fig:data_u}. This dataset represents a standardized benchmark as it is a well understood and widely studied dataset that is used to quantify performance of machine learning algorithms with hyperspectral data. \begin{figure}[htbp] \centering \begin{tabular}{cc} \includegraphics[height=5cm]{figure2_eps-eps-converted-to.pdf} & \includegraphics[height=5cm]{figure3_eps-eps-converted-to.pdf} \\ (a) & (b) \\ \end{tabular} \vspace*{-0.1in} \begin{center} \includegraphics[height=2cm]{figure4_eps-eps-converted-to.pdf} \\ \end{center} \vspace*{-0.1in} \caption{ (a) True color image and (b) ground-truth of the University of Pavia hyperspectral data.} \label{fig:data_u} \end{figure} \subsection{Wetland Data} The second hyperspectral data used in this work was acquired by us in Galveston, Texas in {October, 2014} which includes two different wetland scenes captured at ground-level (side-looking views) over wetlands in Galveston. This is an unprecedented dataset in that it represents ``street view'' type hyperspectral images that have very different characteristics and are very useful for on-ground sensing. The two image cubes are referred to as area 1, and area 2, representing different regions of the wetlands that were imaged --- In addition to common wetland classes, area 2 has Black Mangrove (\emph{Avicennia germinans}) trees in the scene, a species which is of particular interest in ecological studies of wetlands, in addition to Spartina. The images were annotated with species information by an expert in Marine biology, who identified the various species in these hyperspectral scenes. This data was acquired using a Headwall Photonics hyperspectral imager which provides measurements in 325 spectral bands with a spatial size of $1004 \times 5130$. The hyperspectral data uniformly spanned the visible and near-infrared spectrum from $400 nm - 1000 nm$. The objects of interests are primarily vegetation species common in such wetlands. Six different classes were identified in area-1 including soil, \emph{symphyotrichum}, \emph{schoenoplectus}, \emph{spartina patens}, \emph{borrichia} and \emph{rayjacksonia}. The second area includes \emph{Avicennia germinans}, \emph{batis}, \emph{schoenoplectus}, \emph{spartina alterniflora}, \emph{soil}, \emph{water} and \emph{bridge}. Since soil and \emph{schoenoplectus} are included in both areas, the total number of classes in the combined library are eleven. This dataset represents a ``real-world'' application of our method for hyperspectral imaging in the field on arbitrary problems of interest. \begin{figure*}[htbp] \centering \begin{tabular}{c} \includegraphics[height=4cm]{figure5_eps.png} \vspace{-0.2in} \\ Wetland data, area - 1 \\ \includegraphics[height=4cm]{figure6_eps.png} \vspace{-0.2in} \\ Wetland data, area - 2 \\ \end{tabular} \vspace{-0.1in} \begin{center} \includegraphics[width=7cm]{figure7_jpg.jpg} \end{center} \vspace{-0.2in} \caption{True color images of the wetland dataset inset with ground truth.} \end{figure*} \section{Experimental Settings and Results} \label{sec:exp} The efficacy of the proposed LSPP, SLSPP and SBOMP are evaluated as a function of training samples per class using the two practical hyperspectral datasets mentioned previously. SLSPP--SBOMP-C indicates the data is projected based on SLSPP, and SBOMP-C is employed as the backend classifier. A nearest-neighbor (NN) classifier with cosine angle distance is used after LSPP and LADA projection, since they do not take spatial information into account when deriving the projection matrix. Similar to SLSPP--SOMP-C, LSPP--NN and LADA--NN. Each experiment is repeated 10 times using a repeated random subsampling validation technique, and the average accuracy is reported. The number of test samples per class is fixed to 100 for every random subsampling. Fig.~\ref{fig:result_u} shows the classification accuracies as a function of training samples for the University of Pavia hyperspectral data. The parameter for each algorithm is determined via searching through a wide range of the parameter space and the accuracies reported in this plot is based on the optimal parameter values. As can be seen from this figure, our proposed SLSPP followed by SBOMP-C gives the highest classification accuracies consistently over a wide range of the training sample size. LSPP--NN also gives better classification result compared with LADA--NN. We perform a similar analysis using the wetland hyperspectral data. Fig.~\ref{fig:result_g4} and Fig.~\ref{fig:result_g5} plot the classification accuracies with respect to the training sample size per class for wetland area - 1 and wetland area - 2 respectively. It can be seen from the two plots that the proposed methods generally outperform other baseline methods for the vegetation type of hyperspectral data which demonstrate the diverse applicability of the proposed methods. \begin{figure}[h] \centering \includegraphics[height=6cm]{figure8_eps-eps-converted-to.pdf} \\ \caption{Overall classification accuracy (\%) versus number of training samples for the University of Pavia data.} \label{fig:result_u} \end{figure} \begin{center} \begin{figure}[htbp] \centering \includegraphics[height=6cm]{figure9_eps-eps-converted-to.pdf} \\ \caption{Overall classification accuracy (\%) versus number of training samples for the wetland data, area - 1.} \label{fig:result_g4} \end{figure} \end{center} \begin{figure}[htbp] \centering \includegraphics[height=6cm]{figure10_eps-eps-converted-to.pdf} \\ \caption{Overall classification accuracy (\%) versus number of training samples for the wetland data, area - 2.} \label{fig:result_g5} \end{figure} The class specific accuracies for different datasets are shown in Table~\ref{tab:table1}, Table~\ref{tab:table2} and Table~\ref{tab:table3} respectively. In this experiment, the training sample size per class is fixed to 10 and the test sample size is 100 per class. Each experiment is repeated 10 times and the average accuracy is reported. As can be seen from the class-specific tables for the wetland data, the spatial information plays a crucial role, especially for the species with complex textures and shapes such as \emph{Spartina-Patens}, \emph{Sedge} and \emph{Symphyotrichum}. \begin{table*}[htbp] \centering \caption{Class-specific accuracies (\%) for the University of Pavia Data.} \resizebox{13cm}{!}{ \begin{tabular}{ccccc} \toprule \textit{\textbf{Class Name / Algorithm}} & \textit{\textbf{LSPP--SBOMP-C}} & \textit{\textbf{LSPP--SOMP-C}} & \textit{\textbf{LSPP--NN}} & \textit{\textbf{LADA--NN}} \\ \midrule \textit{\textbf{Asphalt}} & 77.2 & 60.8 & 31.4 & 30.9 \\ \textit{\textbf{Meadows}} & 64.4 & 62.5 & 60.2 & 59.6 \\ \textit{\textbf{Gravel}} & 79.2 & 65.5 & 61.1 & 59 \\ \textit{\textbf{Trees}} & 88.2 & 79 & 94.4 & 94.1 \\ \textit{\textbf{Metal Sheets}} & 98.9 & 98.2 & 99.9 & 99.9 \\ \textit{\textbf{Soil}} & 67.6 & 59.1 & 57.1 & 53.4 \\ \textit{\textbf{Bitumen}} & 84.9 & 83 & 84.8 & 84.4 \\ \textit{\textbf{Bricks}} & 60.1 & 64.8 & 66.2 & 62.9 \\ \textit{\textbf{Shadows}} & 99.2 & 96.9 & 96.9 & 95.9 \\ \textit{\textbf{Overall Accuracy}} & 80.0 & 74.4 & 72.4 & 71.1 \\ \bottomrule \end{tabular}% } \label{tab:table1}% \end{table*}% \begin{table*}[htbp] \centering \caption{Class-specific accuracies (\%) for the Wetland, Area-1 data.} \resizebox{13cm}{!}{ \begin{tabular}{ccccc} \toprule \textit{\textbf{Class Name / Algorithm}} & \textit{\textbf{LSPP--SBOMP-C}} & \textit{\textbf{LSPP--SOMP-C}} & \textit{\textbf{LSPP--NN}} & \textit{\textbf{LADA--NN}} \\ \midrule \textit{\textbf{Soil}} & 99.8 & 100 & 99.8 & 99.9 \\ \textit{\textbf{Symphyotrichum}} & 86.8 & 85.5 & 78.1 & 76.1 \\ \textit{\textbf{Sedge}} & 93.4 & 93.3 & 91.6 & 92.1 \\ \textit{\textbf{Spartina-Paten}} & 68.4 & 59.1 & 64.2 & 56 \\ \textit{\textbf{Borrichia}} & 91.6 & 91.2 & 91.7 & 93.7 \\ \textit{\textbf{Rayjacksonia}} & 91.6 & 89 & 84.7 & 92.2 \\ \textit{\textbf{Overall Accuracy}} & 88.6 & 86.4 & 85.0 & 85.0 \\ \bottomrule \end{tabular}% } \label{tab:table2}% \end{table*}% \begin{table*}[htbp] \centering \caption{Class-specific accuracies (\%) for the Wetland, Area-2 data.} \resizebox{13cm}{!}{ \begin{tabular}{ccccc} \toprule \textit{\textbf{Class Name / Algorithm}} & \textit{\textbf{LSPP--SBOMP-C}} & \textit{\textbf{LSPP--SOMP-C}} & \textit{\textbf{LSPP--NN}} & \textit{\textbf{LADA--NN}} \\ \midrule \textit{\textbf{Soil}} & 96.1 & 96.8 & 87.7 & 86.7 \\ \textit{\textbf{Mangrove Tree}} & 97.1 & 95 & 97.1 & 98.5 \\ \textit{\textbf{Batis}} & 99.2 & 99.1 & 97.9 & 97.2 \\ \textit{\textbf{Sedge}} & 97 & 94.3 & 92.3 & 89.8 \\ \textit{\textbf{Spartina-Alterniflora}} & 89.7 & 90.2 & 75.3 & 48 \\ \textit{\textbf{Water}} & 99 & 99 & 98.5 & 98.7 \\ \textit{\textbf{Bridge}} & 99.9 & 99.4 & 94.3 & 85.7 \\ \textit{\textbf{Overall Accuracy}} & 96.9 & 96.3 & 91.9 & 86.4 \\ \bottomrule \end{tabular}% } \label{tab:table3}% \end{table*}% Next, we analyze the effect of the window size (that defines the spatial neighborhood) for the SLSPP method. Fig.~\ref{fig:win_u} depicts the classification accuracies as a function of different window size using the University of Pavia dataset. From this figure, we can see that the optimal window size is 5. \begin{figure}[htbp] \centering \includegraphics[height=6cm]{figure14_eps-eps-converted-to.pdf} \\ \caption{Overall classification accuracy (\%) versus different window size for the University of Pavia and Wetland hyperspectral data.} \label{fig:win_u} \end{figure} In this work, we also visualize the data distributions after projection for the proposed angle-based SLSPP and the Euclidean-based LPP methods. Figure~\ref{fig:vis} (a) shows the subset image for University of Pavia and Figure~\ref{fig:vis} (b) plots all the training samples used in this experiment on an $\ell_2$ normalized sphere. Figure~\ref{fig:vis} (c) and (d) show the same samples after an SLSPP and LPP projection on an $\ell_2$ normalized sphere respectively. $U_1$, $U_2$ and $U_3$ are the three projections found by SLSPP and LPP corresponding to the largest eigenvalues. As can be seen from this figure, SLSPP is much more effective at preserving the inter-sample relationships in terms of spectral angle in the lower-dimensional subspace compared to LPP. \begin{figure}[htbp] \centering \begin{tabular}{cc} \includegraphics[height=3cm]{figure15_eps-eps-converted-to.pdf} & \includegraphics[height=3cm]{figure16_eps-eps-converted-to.pdf} \\ (a) & (b) \\ \includegraphics[height=3cm]{figure17_eps-eps-converted-to.pdf} & \includegraphics[height=3cm]{figure18_eps-eps-converted-to.pdf} \\ (c) & (d) \\ \end{tabular} \caption{Illustrating the (a) subset image of the University of Pavia, (b) original samples, (c) SLSPP projected samples and (d) LPP projected samples on the sphere.} \label{fig:vis} \end{figure} \section{Conclusion} In this work, we have presented an unsupervised variant (LSPP) of the recently developed supervised dimensionality reduction method --- ADA, as well as its spatial variant, SLSPP, that utilize spatial information around the samples when learning the projections. By incorporating spatial information in the dimensionality reduction projection, we are able to learn much more effective subspaces that not only preserves angular information among the training pixels in the feature space, but also their spatial neighbors. We also propose a related sparse representation classifier (SBOMP-C) that is optimized for feature spaces generated by SLSPP. Through experimental results based on two practical real-world hyperspectral datasets, we note that the proposed methods significantly outperform the baseline methods. \bibliographystyle{IEEEtran}
472d17396e1e46c550b7197c35d6ddbbcf62d6c7
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:introduction} Beyond-CMOS technologies using alternate state variables, such as electron spin, magnetic domains, spin waves, phase change, and photons have garnered significant interest lately for low-power and robust computing applications \cite{galatsis2009alternate, allwood2005magnetic, slonczewski1999excitation, qureshi2009scalable}. Spin-based devices, implemented with nanomagnets, have become an extrememly popular choice for magnetic memory, storage, and logic owing to their non-volatility, low power, and low area footprints. Spintronics has established itself as a front-runner in this race of beyond-CMOS technologies, with devices such as the spin-transfer torque (STT) MRAM (that offers high density and high-performance non-volatile storage coupled with low power and cost) predicted to replace conventions DRAMs in the near future \cite{lin200945nm, huai2008spin, kultursay2013evaluating}. The dynamics and performance of the majority of spin-based devices are modeled using the stochastic Landau-Lifshitz-Gilbert-Slonczewski (s-LLGS) equation. This equation describes the temporal evolution of the magnetization vector, $\boldsymbol{M}$, of a monodomain nanomagnet under the effects of magnetic fields, STT and thermal noise \cite{Slonczewski1996, Ralph, Stiles, Liz, stiles2006}. Mathematically, the s-LLGS equation is given as \begin{equation} \frac{d\boldsymbol{M}}{dt} = -\gamma\mu_0(\boldsymbol{M}\times \boldsymbol{H}_{eff}) + \frac{\alpha}{M_s} \Big(\boldsymbol{M}\times\frac{d\boldsymbol{M}}{dt}\Big) - \frac{\boldsymbol{M} \times (\boldsymbol{M} \times \boldsymbol{I_s})}{qN_sM_s}, \label{basic_sLLGS} \end{equation} \noindent where $\gamma$ is the gyromagnetic ratio of electron, $\boldsymbol{H}_{eff}$ is the effective magnetic field experienced by the nanomagnet, $\alpha$ is the dimensionless Gilbert damping constant, $M_s$ is the saturation magnetization of the nanomagnet, $\boldsymbol{I}_s$ is the applied spin current, $q$ is the elementary charge, $N_s$ is the number of spins given as $N_s = \frac{2M_s V}{\gamma\hbar}$, and $V$ is the volume of the nanomagnet.\\ \noindent The nanomagnets used in spintronic devices and memories today are sub-$100$ nm \cite{chun2013scaling, cascales2013magnetization}; for such small dimensions the macrospin approximation \cite{Tannous2006} validates the use of the monodomain model, wherein the nanomagnet is assumed to possess a single coherent magnetization ($\boldsymbol{M}$) over the entire domain and any spatial variation in $\boldsymbol{M}$ is not considered. This approximation is only valid for nanomagnets of dimensions smaller than the critical Stoner radius, which is of the order of a few $100$s of nm for typical magnetic materials \cite{Tannous2006}. The first term on the right hand side (RHS) in \eqref{basic_sLLGS} is the conservative precessional torque that governs the precession of the magnetization vector around the effective field acting on the nanomagnet. This effective field comprises the magnetocrystalline anisotropy field, the shape anisotropy field, and the external applied field. A Langevin field $\boldsymbol{h}_r = h_x \boldsymbol{\hat{x}} + h_y \boldsymbol{\hat{y}} + h_z \boldsymbol{\hat{z}}$, representing Gaussian white noise, is added into the effective field in the s-LLGS equation to model thermal noise. The second term on the RHS in (\ref{basic_sLLGS}) is the Gilbert damping torque, which is responsible for damping the precessions of the magnetization vector and eventually relaxing it to one of its stable states \cite{Aquino2004}. The final term on the RHS in \eqref{basic_sLLGS} is the Slonczewski spin torque arising from the deposition of spin angular momentum by the itinerant electrons of the spin-polarized current. The field-like torque \cite{xiao2005}, (FLT) which results from the non-equilibrium spin accumulation in the nanomagnet, where the transverse component of spin current may persist with a characteristic length of few angstroms or nanometers, is negligible compared to the STT and is not considered in this study. For analysis in this paper, we transform \eqref{basic_sLLGS} into its dimensionless form expressed as\footnote{The spherical coordinates representation of the s-LLGS equation is discussed in section 5.} \begin{equation} \label{eq_llg_implicit1} \frac{d\boldsymbol{m}}{dt} = - \boldsymbol{m} \times \boldsymbol{h}_{eff} + \alpha \left( \boldsymbol{m} \times \frac{d\boldsymbol{m}}{dt} \right) - \boldsymbol{m} \times ( \boldsymbol{m} \times \boldsymbol{i}_s) ,\\ \end{equation} \noindent where we have the normalized quantities $\boldsymbol{m} = \frac{\boldsymbol{M}}{M_s}$, $\boldsymbol{h}_{eff} = \frac{\boldsymbol{H}_{eff}}{M_s}$, and $\boldsymbol{i_s} = \frac{\boldsymbol{I}_s}{I}$. Here, the scaling factor, $I$, for spin current is defined as $I = q\gamma \mu_0M_s N_s$. The time scale is normalized using the factor $(\gamma \mu_0 M_s)^{-1}$. The advantages of the normalized equation \eqref{eq_llg_implicit1} over \eqref{basic_sLLGS} are: (a) it is easier to deal with normalized quantities in terms of numerical complexity, and (b) normalized entities are mathematically well behaved under the application of a numerical scheme. The explicit form of \eqref{eq_llg_implicit1} obtained by decoupling $d\boldsymbol{m}/dt$ is given as (see Appendix A for the detailed derivations) \begin{equation} \label{eq_llg_explicit1} \frac{d\boldsymbol{m}}{dt} = - \frac{1}{1+\alpha^2} \left[ \boldsymbol{m} \times \boldsymbol{h}_{eff} + (\boldsymbol{m} \times (\boldsymbol{m} \times \boldsymbol{i}_s)) + \alpha \left( \boldsymbol{m} \times( \boldsymbol{m} \times \boldsymbol{h}_{eff})) - \alpha (\boldsymbol{m} \times \boldsymbol{i}_s \right)\right]. \end{equation} \noindent Prior works \cite{llg1,llg2,llg3} have reviewed numerical techniques to solve the LLG equation, but a clear and comprehensive treatment of the interplay of thermal noise and deterministic dynamics is lacking. Reference \cite{Aquino2006} applies the implicit midpoint integration technique to the deterministic LLG equation (without thermal noise) and details the intricacies of the accuracy, stability, and time step used in the simulation. But it falls short because all the physical systems in existence experience finite temperature effects; hence, there is a need to include the thermal energy term in the effective field acting on the nanomagnet. The addition of the thermal noise transforms the deterministic LLG equation into a stochastic differential equation (SDE), which warrants an appropriate and careful choice of stochastic calculus and numerical integration scheme for correct convergence. In Ref. \cite{banas2013computational}, the authors analyze the implicit midpoint scheme for a single spin, a finite ensemble of spins, and an infinite ensemble of ferromagnetic spins, but they do not consider the Slonczewski spin torque, which is essential since STT-based switching of nanomagnets has become immensely popular in recent devices owing to its efficiency and reliability. Reference \cite{banas2013computational} also does not offer any comparison of the implicit midpoint scheme with other numerical schemes or present any quantitative evidence as to why one method must be preferred over others with respect to solving the s-LLGS equation. Reference \cite{Pinna2013} presents the impact of thermal noise on the magnetization reversal, but it does not present the details of the numerical scheme employed to treat the thermal noise. In this research, we leverage these prior works to carefully study the impact of the specific numerical scheme to solve the s-LLGS equation on the macrospin evolution. We propose a new method called the RK4-Huen, which is particularly useful when the stochastic term in the SDE is much smaller than the deterministic term, which is usually the case in magnetic bodies. We also comment on the importance of the time step used for numerical integration on the accuracy of various numerical schemes. The second part of this paper focuses on SPICE simulations of nanomagnet dynamics. SPICE-based models are widely used to facilitate the design and optimization of complex magneto-logic networks. However, it is important to represent the s-LLGS equation in SPICE in a manner that the existing numerical integration schemes in SPICE can successfully handle the multiplicative white noise in the system. We quantify the performance (accuracy and stability) of SPICE solvers that use numerical schemes, such as trapezoidal and Euler, to solve the s-LLGS equation in both cartesian and spherical coordinate systems.\\ \noindent The remainder of this paper is organized as follows. Section~\ref{sec:eff_field} describes the formulation of the effective magnetic field acting on the macrospin. The details underlying the midpoint method are presented in Section~\ref{sec:midpoint}. In Section~\ref{sec:methods}, the accuracy of various numerical methods, including a method for small noise, to solve the s-LLGS equation is presented. We conclude the paper by comparing the results of the implicit midpoint method and those obtained from SPICE solvers and the NIST-standard micromagnetic tool OOMMF. \vspace{-2ex} \section{Description of the effective magnetic field} \label{sec:eff_field} To arrive at the expression for the effective field in the s-LLGS equation, we consider the following energies, per unit volume, in the energy landscape of the macrospin \cite{Bruno, Pinna2015,spaldin2010,2005Aquino}: \begin{enumerate}[(1)] \item Zeeman energy due to an externally applied field, $E_{zeeman} = -\mu_0 (\boldsymbol{H}_{app} \cdot \boldsymbol{M}),$ \vspace{-1ex} \item Uniaxial anisotropy energy, $E_{uniaxial} = -K_u\cos^2\theta,$ \vspace{-1ex} \item Shape anisotropy energy, $E_{shape} = \frac{1}{2}\mu_0 (N_xM_x^2 + N_yM_y^2 + N_zM_z^2),$ \vspace{-1ex} \item Thermal energy, $E_T.$ \end{enumerate} \noindent In the above set of equations, $\boldsymbol{H}_{app}$ is the external magnetic field, $K_u = \frac{1}{2}\mu_0M_sH_k $ is the uniaxial energy density ($H_k$ is the anisotropy field), $\theta $ is the angle between the easy axis and the magnetization $(\cos\theta = \boldsymbol{\hat{n}} \cdot \boldsymbol{m} = \boldsymbol{\hat{n}} \cdot \frac{\boldsymbol{M}}{M_s})$, and $N_x, N_y$ and $N_z$ are the geometry-dependent demagnetization coefficients of the nanomagnet. The exchange energy, which originates from the overlap of electron orbitals, is a very short range force that acts on the order of the exchange length of the magnetic material. Since the exchange lengths of typical magnetic materials are on the order of a few nm, and the size of the nanomagnets used in applications today have dimensions of few 10's of nm, the exchange energy becomes insignificant compared to long range interactions like the dipolar interaction, that results in the shape anisotropy term \cite{pinna2015spin}. Hence, for the purposes of this paper, we neglect the exchange energy term in the formulation of the effective field. \\ \noindent The thermal field $\boldsymbol{H}_T(t)$ can be expressed in terms of the Wiener process as $\boldsymbol{H}_T(t) dt = \nu d\boldsymbol{W}(t) $ \cite{Aquino2006}, where $\boldsymbol{W}(t)$ is the Wiener process, and $\nu =\sqrt{\frac{2\alpha K_bT}{\mu_0 M_s^2 V }}$ \cite{Sun2006, Sasikanth2012}. Here{\color{red}{,}} $K_bT$ is the thermal energy. The statistical properties of this thermal field discussed by Brown and Kubo are given as \cite{Brown1963, kubo1970}\\ \indent (1) The mean thermal field: $\langle \boldsymbol{H}_{T,i}(t) \rangle= 0,$\\ \indent (2) The correlation between the components of $\boldsymbol{H}_T(t)$ defined over a time interval $\tau$, \begin{equation} \langle \boldsymbol{H}_{T,i}(t)\boldsymbol{H}_{T,j}(t+\tau) \rangle = \frac{2K_bT\alpha}{\gamma \mu_0^2 M_s V}\delta_{ij}\delta(\tau), \end{equation} where $\delta_{ij}$ is the Kronecker delta function. To simulate the thermal effects numerically, we discretize the model in time \begin{equation} \label{eq_discretized_thermal_model} \boldsymbol{H}_T(t) \Delta t = \nu \Delta \boldsymbol{W}(t), \end{equation} where $\Delta \boldsymbol{W}(t) = \boldsymbol{W}(t + \Delta t) - \boldsymbol{W}(t)$. The normalized standard deviation of the thermal field is given by \begin{equation} \sigma = \sqrt{\frac{2\alpha K_bT}{\mu_0 M_s^2 V}} \sqrt{ \frac{\Delta t' }{\gamma \mu_0 M_s} }, \end{equation} where $\Delta t$ is the time step of the numerical method used and $t' = (\gamma \mu_0 M_s) t$. \noindent We then have \begin{equation} \boldsymbol{h}_T(t) \Delta t' = \sigma \xi_t, \end{equation} where the normalized thermal field $\boldsymbol{h}_T = \boldsymbol{H}_T / M_s$, and $\xi_t \sim \mathcal{N}(0,1)$ is a standard Gaussian vector.\\ \noindent The total energy of the macrospin (excluding the thermal energy) is \begin{equation} \begin{aligned} E_{total} &= V[E_{zeeman} + E_{uniaxial} + E_{shape}] \\ &= V\Big[-\mu_0 (\boldsymbol{H}_{app}\cdot \boldsymbol{M}) - K_u\cos^2\theta + \frac{1}{2}\mu_0 (N_xM_x^2 + N_yM_y^2 + N_zM_z^2)\Big]. \end{aligned} \end{equation} \noindent The total normalized effective field is then given as\\ \begin{equation} \label{heff_1} \begin{aligned} \boldsymbol{h}_{eff} &= \frac{\boldsymbol{H}_{eff}}{M_s} = \frac{-1}{\mu_0 M_s V}\nabla_{\boldsymbol{M}}E_{total}(\boldsymbol{M}) + \boldsymbol{h}_T\\ &= \boldsymbol{h}_{app} + \frac{H_k}{M_s}(\boldsymbol{\hat{n}} \cdot \boldsymbol{m}) \boldsymbol{\hat{n}} - \sum_i N_i\boldsymbol{m}_i +\boldsymbol{h}_T., \end{aligned} \end{equation} \noindent where we have normalized as $\boldsymbol{h}_{app} = \frac{\boldsymbol{H}_{app}}{M_s}, \boldsymbol{m}_i = \frac{\boldsymbol{M}_i}{M_s},$ and $\nabla_{\boldsymbol{M}}$ is the gradient with respect to the magnetization $\boldsymbol{M}$.\\ \vspace{-15pt} \section{The Implicit Midpoint method}\label{sec:midpoint} Using a numerical scheme based on the midpoint rule ensures that~(\ref{eq_llg_implicit1}) converges to the Stratonovich solution in the limit of $\Delta t \rightarrow$ 0. In this section, we define the implicit midpoint method, discretize the deterministic LLGS equation, and derive the Jacobian for use in the Gauss-Newton algorithm. \subsection{The deterministic case} For simplicity, we start the discussion with a one-dimensional (1D) deterministic ordinary differential equations (ODE) of the form \[ y'(t) = f(t, y(t)), \] with initial condition $y(0) = y_0$. The implicit midpoint update is given as \begin{equation} \label{eq_implicit1} y_{n+1} = y_n + \Delta t \cdot f\left( t_n + \frac{\Delta t }{2} , \frac{y_n + y_{n+1}}{2}\right ), \end{equation} where $t_0 = 0$ and $t_{n+1} = t_n + \Delta t$. In general, \eqref{eq_implicit1} is a nonlinear equation of $y_{n+1}$. To solve it for $y_{n+1}$ in the 1D case, one can apply Newton's method on the system \begin{equation} S(y) = y - y_n - \Delta t \cdot f\left( t_n + \frac{\Delta t }{2} , \frac{y_n + y}{2}\right ). \end{equation} \noindent In order to solve a non-linear system $\boldsymbol S:\mathbb{C}^n \to \mathbb{C}^m$, we can use a generalized version of Newton's method, the Gauss-Newton algorithm. Analogous to the 1D case, we can Taylor expand a differentiable system $\boldsymbol S$ \begin{equation} \boldsymbol S(x) = \boldsymbol S( \boldsymbol x_0) + J[\boldsymbol S](\boldsymbol x_0) \cdot (\boldsymbol x - \boldsymbol x_0) + \epsilon, \end{equation} where $J[\boldsymbol S](\boldsymbol x_0)$ is the Jacobian of $\boldsymbol S$ at $\boldsymbol x_0$. Note that $\boldsymbol x, \boldsymbol x_0 \in \mathbb{C}^n$, $\boldsymbol S(\boldsymbol x) \in \mathbb{C}^m$, and $J[\boldsymbol S](\boldsymbol x_0) \in \mathbb{C}^{m \times n}$, and $\epsilon = o(\|\boldsymbol x- \boldsymbol x_0\|_2^2)$. The $ij^{\text{th}}$ entry of the Jacobian of $\boldsymbol S$ is defined by \begin{equation} J[\boldsymbol S](\boldsymbol x)_{ij} = \frac{ \partial S_i( \boldsymbol x) }{\partial x_j}, \end{equation} where $S_i$ is the $i^{\text{th}}$ component of $\boldsymbol S$, and $x_j$ is the $j^{\text{th}}$ component of $\boldsymbol x$. Setting $\boldsymbol S( \boldsymbol x) = 0$, dropping the higher-order terms, and setting $\boldsymbol x_{n+1} = \boldsymbol x$, $\boldsymbol x_n = \boldsymbol x_0$, we can write \begin{equation} \boldsymbol x_{n+1} = \boldsymbol x_n - J[\boldsymbol S](\boldsymbol x_n)^{-1} \cdot \boldsymbol S(\boldsymbol x_n), \end{equation} if $J[\boldsymbol S]$ is invertible. As a result, we can use implicit methods on multi-dimensional ODE's, including the s-LLGS equation. \subsection{Discretizing the deterministic Landau-Lifshitz-Gilbert-Slonczewski equation} In order to apply the midpoint method, we need to discretize \eqref{eq_llg_implicit1}. Note that \eqref{eq_implicit1} takes the form \begin{equation} \label{eq_llg_implicit_step} \boldsymbol m_{n+1} = \boldsymbol m_n + \Delta t \cdot \boldsymbol f(t_n + \Delta t / 2, \boldsymbol m_{n+1/2}). \end{equation} Using the implicit formulation of $f$, we obtain \begin{equation} \begin{aligned} \boldsymbol f(t_n + \Delta t/2, \boldsymbol m_n, \boldsymbol m_{n+1}) =& (-\boldsymbol m_{n+1/2} \times \boldsymbol h_{n+1/2}) + \alpha (\boldsymbol m_{n+1/2} \times \Delta \boldsymbol m) \\ &- \boldsymbol m_{n+1/2} \times( \boldsymbol m_{n+1/2} \times \boldsymbol i_s), \end{aligned} \end{equation} where $\boldsymbol m_{n+1/2} = (\boldsymbol m_n + \boldsymbol m_{n+1})/2$, $\boldsymbol h_{n+1/2} = \boldsymbol h(\boldsymbol m_{n+1/2})$, and $\Delta \boldsymbol m = (\boldsymbol m_{n+1} - \boldsymbol m_n)/ \Delta t$. Since we are dealing with an implicit method, using the implicit definition of $d \boldsymbol m/dt$ from \eqref{eq_llg_implicit1} does not require any additional computation. In fact, using the implicit definition of $d\boldsymbol m/dt$ yields a more efficient method. This is due to the reduced complexity of the Jacobian of the system $\boldsymbol S_n$, which is used to solve for $\boldsymbol m_{n+1}$, \begin{equation} \label{eq_llg_implicit_system} \boldsymbol S_n(\boldsymbol m) = \boldsymbol m - \boldsymbol m_n - \Delta t \cdot \boldsymbol f(t_n + \Delta t/2, \boldsymbol m_n, \boldsymbol m ). \end{equation} \noindent Having set up the system $\boldsymbol S_n$, it remains to derive its Jacobian with respect to $\boldsymbol m$. \subsection{Derivation of the Jacobian} First, we state that for $\boldsymbol a, \boldsymbol b \in \mathbb{C}^3$, the cross product of $a$ and $b$ can be restated in terms of matrix-vector multiplication \begin{subequations} \begin{equation} \boldsymbol a \times \boldsymbol b = (a^{\times}) \boldsymbol b, \end{equation} where \begin{equation} a^{\times} = \begin{bmatrix} 0 & -a_3 & a_2 \\ a_3 & 0 & -a_1 \\ -a_2 & a_1 & 0 \end{bmatrix}. \end{equation} \end{subequations} \noindent The Jacobian of a cross product of two functions $\boldsymbol f,\boldsymbol g: \mathbb{C}^n \to \mathbb{C}^3$ is given by \cite{Choroszucha2010} \begin{equation} J[\boldsymbol f \times \boldsymbol g] = f^{\times} J[\boldsymbol g] - g^{\times} J[\boldsymbol f] . \end{equation} \noindent Using the above relation, one can derive the Jacobian as \begin{equation} \label{eq_jacobian} \begin{aligned} J[\boldsymbol S_n](\boldsymbol m) = \ &I + \Delta t/2 \left( \boldsymbol m_{n+1/2}^{\times} J[\boldsymbol h] - \boldsymbol h_{n+1/2}^\times \right) \\ & + \Delta t /2 \left[ \boldsymbol m_{n+1/2}^\times \boldsymbol i_s^\times - \left(\boldsymbol m_{n+1/2} \times \boldsymbol i_s \right)^\times \right] \\ &- \alpha \boldsymbol m_n^\times, \end{aligned} \end{equation} \noindent where $\boldsymbol m_{n+1/2} = (\boldsymbol m+\boldsymbol m_n)/2$, $\boldsymbol h_{n+1/2} = \boldsymbol h(\boldsymbol m_{n+1/2})$, and \begin{equation} J[\boldsymbol h] = \frac{H_k}{M_s} \begin{bmatrix} n_x^2 & n_x n_y & n_x n_z \\ n_y n_x & n_y^2 & n_y n_z \\ n_z n_x & n_z n_y & n_z^2 \end{bmatrix} - \begin{bmatrix} N_x & & \\ & N_y & \\ & & N_z \end{bmatrix}, \end{equation} \noindent where $\boldsymbol n= (n_x, n_y, n_z)^T$ is the easy axis, and $\boldsymbol N = (N_x, N_y, N_z)^T$ are the demagnetization coefficients. To summarize, we will use the Jacobian in \eqref{eq_jacobian} to solve $\boldsymbol S_n$ in \eqref{eq_llg_implicit_system} for the magnetization value $\boldsymbol m_{n+1}$ using the Gauss-Newton algorithm. \section{Methods to solve the s-LLGS equation}\label{sec:methods} In this section, we first obtain the Stratonovich SDE form of the s-LLGS equation. We then formulate the various numerical schemes for Stratonovich SDEs. Specifically, we focus on implicit midpoint, Heun, Euler-Heun, and a new method, namely the RK4-Heun. The accuracy of these methods is tested on an SDE with a known analytical solution. Finally, we perform numerical tests for these methods by applying them on the s-LLGS equation. \subsection{Stratonovich SDE form of the s-LLGS equation} \noindent A stochastic integral over a Wiener process $W(t): \mathbb{R}^+ \to \mathbb{R}$ \cite{Morters2010} is defined as \begin{equation} \int_a^b g(t ) dW(t) = \lim_{\Delta t \to 0} \sum_k g(\tau_k) (W(t_{k+1}) - W(t_k)), \end{equation} where $\{ t_k \}_k$ forms a partition of $[a,b]$, and $\tau_k \in [t_{k+1}, t_k]$. Contrary to the Riemann integral, different choices of $\tau_k$ yield different results of the integral. This is the case if $g$ is a function of both $t$ and the Wiener process $W(t)$. The most common choices are $\tau_k = t_k$ and $\tau_k = (t_{k+1} + t_k)/2$. The former yields the Ito calculus, while the latter results in the Stratonovich calculus. A widely used notation for the Stratonovich integral to distinguish it from the Ito integral is \begin{equation} \int_a^b g(t) \circ dW(t), \end{equation} which we will use in the following. Numerical schemes that attempt to approximate an SDE's solution have to be constructed for a specific calculus. In our case, this is the Stratonovich interpretation \cite{Gardiner1997}.\\ \noindent In a somewhat general form, we can write a Stratonovich SDE as \begin{equation} \label{eq_stochastic_def} dX_t = f(X_t, t) dt + g(X_t, t) \circ dW_t. \end{equation} The s-LLGS equation in \eqref{eq_llg_explicit1} can be restated in the general Stratonovich SDE form given in \eqref{eq_stochastic_def}. To this end, let $X(t) = \boldsymbol{m}_t$, and define \begin{subequations} \label{eq_fg_stochastic} \begin{equation} \label{eq_f_stochastic} \begin{aligned} f(\boldsymbol{m}_t, t) =& -\alpha' \ [ \boldsymbol{m}_t \times \boldsymbol{h} + \boldsymbol{m}_t \times (\boldsymbol{m}_t \times \boldsymbol{i}_s) \\ &+ \alpha (\boldsymbol{m}_t \times (\boldsymbol{m}_t \times \boldsymbol{h}) - \boldsymbol{m}_t \times \boldsymbol{i}_s )], \end{aligned} \end{equation} \begin{equation} \label{eq_g_stochastic} \begin{aligned} g(\boldsymbol{m}_t,t) =& - (\alpha' \nu) \ m_t^\times [ I + \alpha m_t^\times ], \end{aligned} \end{equation} \end{subequations} where $\alpha' = 1/(1 + \alpha^2)$, and $\nu$ is the standard deviation of the Wiener process, which is used to model the thermal field fluctuations. Together \eqref{eq_stochastic_def} and \eqref{eq_fg_stochastic} define the form of the s-LLGS equation we will be working with throughout this paper. Further, we will deal with two modes of convergence of numerical approximations to solutions of SDEs introduced in \cite{Kloeden_Platen}. Given a solution $X_t$ of \eqref{eq_stochastic_def}, an approximation $\tilde X_t$ is said to converge to $X_t$ in the {\it strong sense with order} $\gamma > 0$ if there is a $ C \in \mathbb{R}$ such that \begin{equation} \mathbb{E} \left( \left| \tilde X_t - X_t \right| \right) < C (\Delta t)^\gamma, \end{equation} holds for any discrete approximation $\tilde X_t$ with maximum step-size $\Delta t \>$\cite{Kloeden_Platen}. On the other hand, an approximation $\tilde X_t$ is said to converge to $X_t$ in the {\it weak sense with order} $\gamma > 0$ if there is a $ C \in \mathbb{R}$ such that \begin{equation} \left| \mathbb{E} ( p( \tilde X_t) ) - \mathbb{E} (p(X_t)) \right| < C (\Delta t)^\gamma, \end{equation} holds for any polynomial $p$ and any discretization with maximum step size $\Delta t$. Note that convergence in the strong sense is equivalent to convergence for each realization of $W_t$ (i.e. path), while convergence in the weak sense merely implies that the statistical properties of the approximation converge to those of the solution. \subsection{Numerical schemes for Stratonovich SDEs} In order to numerically solve \eqref{eq_stochastic_def}, the following must be computed: \begin{equation} \label{eq_stochastic_solution} X_t = \int_0^t f(X_t, t) dt + \int_0^t g(X_t,t) \circ dW(t) + X_0, \end{equation} where the first integral is a regular Riemann integral, and the second integral is a stochastic integral interpreted in the Stratonovich sense. In our case, we are dealing with three-dimensional (3D) vector integrals, and $W(t)$ is a 3D Wiener process. Below, we briefly outline the methods that we focus on in this work. \begin{enumerate} \item \textit{Euler-Heun}\\ Arguably the simplest method that converges to the Stratonovich solution is the Euler-Heun method defined by \begin{subequations} \begin{equation} X_{n+1} = X_n+ f(X_n, t_n) \Delta t + \frac{1}{2} \left[ g( \tilde X_{n+1}, t_{n+1}) + g(X_n, t_n)\right] \eta_n, \end{equation} \begin{equation} \tilde X_{n+1} = X_n + g(X_n, t_n)\eta_n, \end{equation} where $\eta_n = \sqrt{\Delta t} \xi_n$, $\Delta t \in \mathbb{R}^+$, and $\xi_n \sim \mathcal{N}(0,1)$. \end{subequations} \item \textit{Heun}\\ The Heun method is given by \begin{equation} X_{n+1} = X_n + \frac{1}{2} \left[ f(\tilde X_{n+1}, t_{n+1}) + f(X_n, t_n) \right] \Delta t+ \frac{1}{2}\left[ g( \tilde X_{n+1}, t_{n+1}) + g(X_n, t_n)\right] \eta_n, \end{equation} where \begin{equation} \begin{aligned} \tilde X_{n+1} = X_n + f(X_n, t_n) \Delta t + g(X_n, t_n) \eta_n. \end{aligned} \end{equation} \item \textit{Implicit midpoint}\\ The implicit midpoint rule is defined by \begin{equation} \begin{aligned} X_{n+1} = X_n + f(X_{n+1/2}, t_n+\Delta t/2) \Delta t + g(X_{n+1/2}, t_n+ \Delta t/2) \eta_n, \\ \end{aligned} \end{equation} where $X_{n+1/2} = (X_{n+1} + X_n) /2$. Note that this defines an implicit system just as in the deterministic case, and we can use the methods derived for the deterministic case to solve it. We use the Euler step \begin{equation} X^* = X_n + f(X_n, t_n) \Delta t + g(X_n, t_n) \eta_n, \end{equation} as the initial guess for the Gauss-Newton algorithm, to find $X_{n+1}$.\\ \end{enumerate} \vspace*{-3ex} \noindent All of the above methods are known to converge to the Stratonovich solution with a strong order $1/2$, and a weak order $1$ \cite{Kloeden_Platen,Milstein2002}. Compared to numerical methods for deterministic DEs, this order of convergence is low. However, stochastic higher-order methods generally require the approximation of iterated stochastic integrals, which is a detailed and time-intensive procedure \cite{Gaines_1993}. If an SDE has a special structure, like an additive or commutative noise term, this calculation can be simplified. Unfortunately, the noise term of the s-LLGS equation is multiplicative and non-commutative.\\ \noindent Further, it is important to note that semi-implicit numerical methods based on extrapolation, which are used to handle deterministic DEs, generally do not converge to solutions of SDEs. The reason for this is that extrapolations from the past destroy the Markov property (independent increments) of the Wiener process over which we integrate. As seen in Figure \ref{fig_adam_path}, a semi-implicit midpoint scheme using Adam's extrapolation \cite{Serpico2001}, where $X_{n+1/2} = \frac{1}{2}(3X_n-X_{n-1})$, does not converge to the solution of the test SDE, \begin{equation} \label{eq_test_eq} dX_t = X_t dt + X_t \circ dW_t, \end{equation} with solution \begin{equation} \label{eq_test_sol} X_t = e^{t+W_t}. \end{equation} \vspace*{-3ex} \begin{figure}[H] \begin{center} \includegraphics[scale=0.42]{adam_path_err} \end{center} \vspace*{-3ex} \caption{\footnotesize Average path-wise error of midpoint approximations using Adam's extrapolation compared to the analytical solution of \eqref{eq_test_eq}.} \label{fig_adam_path} \end{figure} \subsection{A method for small noise} In many stochastic models corresponding to physical systems, the noise term is usually much smaller than the drift term. Indeed, while the stochastic term plays a critical role in the s-LLGS equation, it is frequently one to two orders of magnitude smaller than the deterministic term, if $T$ is close to room temperature. This means a higher order approximation of the drift term alone could improve the accuracy of the entire approximating process, when the absolute error is dominated by the error in the deterministic term. This idea has been investigated for Ito SDEs in \cite{Mil_1997, Ro_2013}. Here, we propose a scheme for Stratonovich SDEs, which we are going to refer to as RK4-Heun. It is defined by \begin{equation} \begin{aligned} X_{n+1} &= X_n + D \Delta t + S \eta_n, \\ D &= (d_1 + 2d_2 + 2d_3 + d_4)/6, \\ S &= (s_1 + s_2)/2,\\ s_1 &= g(X_n,t_n),\\ s_2 &= g(X_n + f(X_n, t_n) \Delta t + s_1 \eta_n), \\ d_1 &= f(X_n,t_n),\\ d_2 &= f(X_n + (d_1 + s_1)/2, t_n + \Delta t/2), \\ d_3 &= f(X_n + (d_2 + s_1)/2, t_n + \Delta t/2), \\ d_4 &= f(X_n + d_3 + s_1, t_n + \Delta t). \end{aligned} \end{equation} \noindent Importantly, the RK4 stages are only used to approximate the deterministic part of the equation. Therefore, convergence to the Stratonovich solution is maintained. Further, we made the observation that making the additional update \begin{equation} \begin{aligned} X_{n+1}' &= X_n + D \Delta t + S' \eta_n, \\ S' &= (s_1 + s_2')/2, \\ s_2' &= g(X_{n+1}, t_{n+1}), \\ \end{aligned} \end{equation} after computing the above, leads to a significant increase in accuracy for the test equation \eqref{eq_test_eq}. The additional update includes the higher order approximation of the deterministic term in the Heun step of the stochastic term. As a consequence, it is not surprising that this leads to a better approximation. However, we have not yet investigated analytical justifications of this update. Future work will look into that and the possible limitations of this effect. \subsection{Numerical tests using a general SDE} To compare the accuracy of the various methods discussed above, we consider a modified version of the test SDE \eqref{eq_test_eq}, which can be solved analytically, \begin{equation} \label{eq_test_gen} dX_t = a X_t dt + b X_t \circ dW_t, \end{equation} where $a, b \in \mathbb{R}$. A Stratonovich solution to \eqref{eq_test_gen} is given by \begin{equation} \label{eq_integral_test_eq} X_t = X_0 + \int_0^t X_s ds + \int_0^t X_s \circ dW_s = e^{at+b W_t}, \end{equation} where $X_0 = 1$. We can now compare the results of different numerical schemes to this analytical solution. \noindent The implicit midpoint method for \eqref{eq_test_eq} is defined by \begin{equation} X_{t+1} = X_t + a \frac{1}{2} ( X_{t+1} + X_t) \Delta t + b \frac{1}{2}(X_{t+1} + X_t) \eta_t. \end{equation} This can be solved for $X_{t+1}$ analytically, according to \begin{equation} \begin{aligned} \label{eq_test_implicit} X_{t+1} = X_t \left( \frac{ 2 + c}{2 - c} \right), \\ c = a \Delta t + b \eta_t. \end{aligned} \end{equation} \noindent The experimental results for the path-wise error to the analytical solution, $X_{1}$, are shown in Figure \ref{fig_path_err_1} and \ref{fig_path_err_2} for two different choices of $b$ in \eqref{eq_test_gen}. The data shows that the RK4-Heun scheme is more accurate than all other methods for all step sizes considered. Interestingly, the empirical order of convergence, which is the slope of the error function in Figure \ref{fig_path_err}, is only constant for the Euler-Heun method. All other methods start out with a faster order (steeper slope) for larger step sizes. This is due to the fact that the deterministic part of the equation, and hence the error in its approximation, dominates the stochastic part. Therefore, the slope for big step sizes is more like the deterministic second order of convergence for the Heun and implicit midpoint schemes. We see this effect for smaller step sizes as the size $b$ of the stochastic part is decreased in Figure~\ref{fig_path_err_2}. \vspace{-2ex} \begin{figure}[H] \centering \begin{subfigure}{.45\linewidth} \centering \includegraphics[scale=0.4]{path_err_3} \vspace{-3ex} \caption{$a = 1, b = 0.1$} \label{fig_path_err_1} \end{subfigure} \qquad \begin{subfigure}{.45 \linewidth} \centering \includegraphics[scale=.4]{path_err_2} \vspace{-3ex} \caption{$a = 1, b = 0.01$} \label{fig_path_err_2} \end{subfigure} \vspace{-1ex} \caption{\footnotesize Average path-wise error of the numerical approximations corresponding to equation \eqref{eq_test_gen} with two different choices of $a$ and $b$. } \label{fig_path_err} \end{figure} \vspace{-3ex} \subsection{Numerical tests using the s-LLGS equation} \noindent Now, we are going to test the properties of these methods on the s-LLGS equation, defined by \eqref{eq_f_stochastic} and \eqref{eq_g_stochastic}. It is important to note that the norm of the magnetization vector $\boldsymbol{m}$ is preserved if the equation is interpreted in the Stratonovich sense. This follows from the fact that in the Stratonovich calculus, \[ d( \| m_t \|^2 ) = 2 \boldsymbol{m}_t \cdot d\boldsymbol{m}_t = 0, \] whereas in the Ito calculus, Ito's Lemma gives \vspace{-2ex} \[ d( \|m_t\|^2 ) = 2 \boldsymbol{m}_t \cdot d\boldsymbol{m}_t + \sum_{i,j = 1}^3 \frac{ \partial^2 \| \boldsymbol m\|^2}{\partial m_i \partial m_j}(\boldsymbol{m}_t) ~ \left( dm_{i,t} \cdot dm_{j,t} \right), \] which is, in general, not zero. Explicit schemes like Euler-Heun or Heun, while solving the Stratonovich equation, do not preserve the norm. Indeed, one has to take a very small step size $\Delta t$ so that the norm of the magnetization does not blow up. On the other hand, the implicit midpoint method preserves the norm. The contrasting behavior of the magnetization norm obtained using different stochastic calculi is highlighted in Figures 3a and 3b. Indeed the midpoint method that converges to the Stratonovich solution preserves the norm while the Ito solution does not. \vspace{-2ex} \begin{figure}[H] \centering \begin{minipage}{.5\textwidth} \centering \includegraphics[scale=0.39]{stratonovich} {(a)} \label{fig:test1} \end{minipage}% \begin{minipage}{.5\textwidth} \centering \vspace*{-2ex} \includegraphics[scale=0.39]{ito} (b) \label{fig:test2} \end{minipage} \vspace{-2ex} \caption{\footnotesize Time evolution of magnetization with (a) implicit midpoint converging to the Stratonovich solution and (b) Heun-Euler converging to the Ito solution. Conditions are zero spin current (only noise) and energy barrier $U = 10K_BT$. The magnetization norm is preserved with the former, but blows up with the latter. } \end{figure} \noindent The deviation of the magnetization norm from unity for different numerical techniques is depicted in Figure \ref{fig_norm_comparison}. The error tolerance for the Gauss-Newton algorithm in the implicit midpoint method was $10^{-12}$. For this reason, the norm is only preserved up to this order of magnitude. In practice, decreasing the error tolerance further only slows down the algorithm without significant benefit. In Figure \ref{fig_magnetization_difference}, we show the maximum norm of the difference of the magnetization vectors, calculated by the Heun, implicit midpoint, and RK4-Heun methods. The difference is quite pronounced at larger time steps as expected, and diminishes as the time step is reduced. We see that the explicit methods converge to the same path quickly, while the difference to the implicit scheme decreases at a slower rate.\\ \vspace{-4ex} \begin{figure}[H] \centering \begin{subfigure}{.45\textwidth} \centering \includegraphics[scale=0.4]{mag_norm_plot} \caption{\footnotesize Norm of magnetization vector in the course of several simulations with different $\Delta t$, calculated with three numerical schemes. Here, $1$ ns $\sim 66$ time units.} \label{fig_norm_comparison} \end{subfigure} \qquad \begin{subfigure}{.45\textwidth} \centering \includegraphics[scale=0.4]{traj_diff} \caption{\footnotesize Maximum norm of the difference of magnetization vectors as calculated with the implicit midpoint method, Heun, and RK4-Heun scheme. } \label{fig_magnetization_difference} \end{subfigure} \caption{\footnotesize Results of using the Heun, RK4-Heun, and implicit midpoint scheme on the s-LLGS equation.} \label{fig_LLG_bench} \end{figure} \noindent It would be interesting to see the discrepancy in the switching boundaries obtained from the Heun and implicit midpoint schemes. The 50$\%$ switching boundary is a useful metric since it is commonly used to benchmark solvers. To this end, we construct a contour plot (Figure \ref{3Dcontour}) by sweeping over spin current durations and amplitudes to ascertain the 50$\%$ and 90$\%$ switching boundaries from the two methods. The maximum discrepancy in the switching boundaries is observed at the 50$\%$ switching probability. However, the difference in the switching probability between the two methods diminishes in the high current, long time-scale regimes for the chosen time step of numerical integration. \vspace{-4ex} \begin{figure}[H] \includegraphics[scale=0.47]{contour} \vspace{-4ex} \caption{\footnotesize Contour plot for switching probability as a function of spin current amplitude and duration. 50$\%$ and 90$\%$ switching boundaries are depicted for implicit midpoint and explicit Heun schemes.} \label{3Dcontour} \vspace{0ex} \end{figure} \section{SPICE simulation of macrospin dynamics} Circuit-level simulations using nanomagnets are often conducted using SPICE-based models, which represent the nanomagnet using basic circuital elements such as resistors, capacitors, and current sources. It is, therefore, important to benchmark the accuracy of SPICE numerical solvers with respect to the implicit midpoint method to solve the s-LLGS equation. SPICE uses methods such as Euler, Gear, trapezoidal, or a combination of these, to numerically integrate DEs. Although the trapezoidal method is compatible with the Stratonovich stochastic calculus, other numerical methods in SPICE require that the s-LLGS equation be transformed into the equivalent Ito representation for correct convergence. While there are many versions of SPICE available \cite{ngspice2011,xyce2014,smartspice2010}, possibly with varying minimum time steps and specific algorithms to solve equations, we use NGSPICE for our simulatons in this paper. Figure~\ref{fig:nanomagnet_schem} shows the SPICE circuit implementation of a nanomagnet subject to thermal effects and spin torque, in one particular direction (x, y or z). The other two magnetization directions are implemented similarly. The magnetization $\boldsymbol{m} = \boldsymbol{M}/M_s$ is modeled as the node voltage of a capacitor, and the effective field inside the nanomagnet is represented as a dependent current source. Detailed circuit derivations can be found in prior works \cite{Sasikanth2012} and \cite{Phillip}. The s-LLGS equation was implemented in SPICE in both its cartesian and spherical coordinates form, and the results from these implementations are discussed below. \subsection{SPICE implementation of the s-LLGS equation in cartesian form} Large-scale Monte Carlo simulations were conducted using the circuit representation in Figure~\ref{fig:nanomagnet_schem} to generate the probability density function (PDF) plots of the magnetization reversal delays of an in-plane nanomagnet. The time step of integration is taken to be 1 ps, reducing which, leads to convergence issues in the SPICE nodal analysis. From the results shown in Figure~\ref{fig:spice_vs_matlab}, we find that SPICE solvers working with the cartesian form of the s-LLGS equation overestimate the mean delay of magnetization reversal as compared to the implicit midpoint method. In addition, unlike the implicit midpoint method, SPICE solvers are unable to preserve the magnetization norm with the cartesian s-LLGS eqauation, as shown in Figure \ref{SPICE_norm}. We also highlight the differences in the initial angle distribution obtained from various numerical schemes in Figure \ref{initial_angle}. In this figure, the macrospin is subject to thermal noise for a period of 1 ns and no external magnetic field or spin torque is applied. The s-LLGS equation conserves magnetization norm and has a Boltzmann initial angle distribution only if the Stratonovich stochastic calculus with the midpoint prescription is used. For other stochastic discretization schemes, none of these properties are ensured and a carefully chosen drift term has to added to recover the validity of these properties when other stochastic calculi are used \cite{roma2014numerical}. As expected, the discrepancy in results is exacerbated for Euler method due to the fact that it does not solve for the correct Stratonovich SDE equation and will require an alternative Ito representation of the s-LLGS equation. Whereas the trapezoidal method does converge to the Stratonovich solution for $\Delta t \rightarrow 0$, but it is limited by the minimum time step of the DE solver in SPICE. Hence we see a 200 ps difference in the mean reversal delay obtained with implicit midpoint and the trapezoidal solver of SPICE (with cartesian s-LLGS equation). It must be noted that these significant errors in magnetization reversal computed from SPICE will ultimately lead to inaccurate estimates of circuit-level performance metrics such as error-rate, latency, and energy dissipation of complex magneto-logic networks. \\ \vspace*{-4ex} \begin{figure}[H] \centering \begin{subfigure}[t]{0.5\textwidth} \centering \includegraphics[scale=0.47]{schematic} \caption{\footnotesize Schematic of nanomagnet used in the circuit\\ simulations.} \label{fig:nanomagnet_schem} \end{subfigure}% ~ \begin{subfigure}[t]{0.5\textwidth} \centering \includegraphics[scale=0.33]{SPICEvsMATLAB3-eps-converted-to.pdf} \caption{\footnotesize PDFs of reversal delays of a nanomagnet using implicit midpoint, SPICE trapezoidal solver (cartesian and spherical s-LLGS), and SPICE backward-Euler solver.} \label{fig:spice_vs_matlab} \end{subfigure} \vspace*{-1ex} \caption{\footnotesize The simulations were performed for an in-plane magnet using identical dimensions $40\times 40\times 1$ nm$^3$, parameters $M_s = 1.11e6$ A/m, $H_k = 1.11e5$ A/m, $\alpha = 0.01$, time step $1$ ps, and spin current $0.16$ mA.} \end{figure} \vspace*{-2ex} \begin{figure}[H] \centering \begin{minipage}{.5\textwidth} \centering \includegraphics[scale=0.4]{SPICEnorm-eps-converted-to.pdf} \caption{\footnotesize Norm of the magnetization vector in the course of a simulation, calculated with SPICE and implicit mid-point solvers. The parameters used are the same as for Figure~\ref{fig:spice_vs_matlab}. } \label{SPICE_norm} \end{minipage}% ~ \hspace*{2ex} \begin{minipage}{.5\textwidth} \centering \vspace{2ex} \includegraphics[scale=0.42]{SPICE_init2-eps-converted-to.pdf} \caption{\footnotesize Initial angle distribution of the nanomagnet after it has been subjected to only thermal noise for 1 ns, from SPICE and our implicit midpoint solver. \\} \label{initial_angle} \end{minipage} \end{figure} \subsection{SPICE implementation of the s-LLGS equation in spherical form} To mitigate the aforementioned errors in the SPICE-based solvers, we look at the implementation of the s-LLGS equation in spherical coordinates form. The spherical coordinates of the unit magnetization vector $\mathbf{m}$ at any position can be written as $(\theta, \phi)$, where $\theta$ is measured from the positive direction of $z$-axis, and $\phi$ is measured counterclockwise from the positive direction of $x$-axis. Further, the local orthogonal unit vectors in the directions of increasing $\theta$, and $\phi$ respectively are given by \begin{equation} \boldsymbol{\hat{\theta}} = [\cos\theta \cos\phi, \cos\theta \sin\phi, -\sin\theta]; \quad \boldsymbol{\hat{\phi}} = [-\sin\phi, \cos\phi, 0]. \end{equation} \noindent The s-LLGS equation in spherical coordinates can be expressed in terms of the $\theta$ and $\phi$ components of the effective field ($h_{\theta}, h_{\phi}$), and spin current ($i_{s\theta}, i_{s\phi}$) as \begin{equation} \begin{bmatrix} \frac{d\theta}{dt}\\ \sin\theta \cdot \frac{d\phi}{dt} \end{bmatrix} = \frac{1}{1+\alpha^2}\begin{bmatrix} h_{\phi} + i_{s\theta} + \alpha h_{\theta} - \alpha i_{s\phi}\\ i_{s\phi} - h_{\theta} + \alpha h_{\phi} + \alpha i_{s\theta} \end{bmatrix}, \label{eq:LLG_sph} \end{equation} \begin{wrapfigure}{l}{0.52\textwidth} \vspace{-5ex} \includegraphics[scale=0.31]{SPICE_init1-eps-converted-to.pdf} \caption{\footnotesize Initial angle distribution of the nanomagnet after it has been subjected to only thermal noise for 1 ns, from SPICE trapezoidal solver (spherical s-LLGS) and our implicit midpoint solver} \label{spherical_init} \vspace*{-1ex} \end{wrapfigure} \noindent where $(i_s/h)_{\theta} = \mathbf{(i_s/h)}\cdot\boldsymbol{\hat{\theta}}$, $(i_s/h)_{\phi} = \mathbf{(i_s/h)}\cdot\boldsymbol{\hat{\phi}}$, and $\cdot$ denotes dot product. The solution ($\theta$, $\phi$) of (\ref{eq:LLG_sph}) can be reverted back to cartesian coordinates by using $\mathbf{m} = [\sin\theta\cos\phi, \sin\theta\sin\phi, \cos\theta]$. Though a change of coordinate system does not change the solution of the system being solved, there are some subtle differences in the numerical implementation of the cartesian and spherical forms. An apparent difference is the need to solve for only two variables $\theta$ and $\phi$ in spherical coordinates, compared to three variables $m_x, m_y$ and $m_z$ in cartesian. This makes implementing s-LLGS equation in spherical coordinates less expensive, especially when using implicit methods, where the overhead cost of calculating the $\theta$ and $\phi$ components (few multiplications) is way less than solving the implicit step (which incurs fixed point iteration). This also forces the norm to be preserved to unity automatically as shown in Figure \ref{SPICE_norm}. Importantly, we can also see from Figures \ref{fig:spice_vs_matlab} and \ref{spherical_init} that the deviation in the delay distribution and initial angle distribution from the implicit midpoint solver is minimized for the case of the spherical s-LLGS implementation. \noindent Figures \ref{fig:carLLG} and \ref{fig:sphLLG} show the trajectory of magnetization of the cartesian and spherical coordinates implementation of the s-LLGS equation for a range of time steps. For smaller time steps, the cartesian and spherical forms yield identical solutions which are both accurate and stable. For moderate time steps, the cartesian form stretches the trajectory to outside the unit circle; the spherical form has kinks around $m_z = \pm 1$ due to singularity of $\frac{d\phi}{dt}$ in (\ref{eq:LLG_sph}) at $\theta = 0, \pi$. This results in underestimation of reversal delay at moderate time step. For larger time steps, the cartesian and spherical forms produce unstable solutions, which are unbounded and oscillatory in nature respectively. In short, there is a trade-off between differentiability and boundedness in the cartesian and spherical implementations, respectively. \vspace{4ex} \begin{figure}[H] \begin{subfigure}[t]{0.33\textwidth} \includegraphics[width=0.95\textwidth]{car_small} \caption{Small time step: 0.01} \label{fig:car_small} \end{subfigure}% ~ \begin{subfigure}[t]{0.33\textwidth} \includegraphics[width=0.95\textwidth]{car_medium} \caption{Moderate time step: 0.1} \label{fig:car_medium} \end{subfigure}% ~ \begin{subfigure}[t]{0.33\textwidth} \includegraphics[width=0.95\textwidth]{car_large} \caption{Large time step: 0.3} \label{fig:car_large} \end{subfigure} \caption{\footnotesize Cartesian form s-LLGS simulation of magnetization reversal from $-x$ to $+x$ direction for an applied field $H_{app}= M_s$ along positive x-axis for different time steps (normalized). The simulation parameters are same as that in Figure \ref{fig:spice_vs_matlab}. (a) Small time step yields both accurate and stable solution. (b) Moderate time step gives a stable solution which is stretched along the reversal direction, thereby underestimating the reversal delay (zero crossing of $m_x$). (c) Large time step results in unbounded and unstable solution.} \label{fig:carLLG} \end{figure} \begin{figure}[H] \begin{subfigure}[t]{0.33\textwidth} \includegraphics[width=0.95\textwidth]{sph_small} \caption{Small time step: 0.01} \label{fig:sph_small} \end{subfigure}% ~ \begin{subfigure}[t]{0.33\textwidth} \includegraphics[width=0.95\textwidth]{sph_medium} \caption{Moderate time step: 0.1} \label{fig:sph_medium} \end{subfigure}% ~ \begin{subfigure}[t]{0.33\textwidth} \includegraphics[width=0.95\textwidth]{sph_large} \caption{Large time step: 0.3} \label{fig:sph_large} \end{subfigure} \caption{\footnotesize Spherical form s-LLGS simulation of magnetization reversal from $-x$ to $+x$ direction for an applied field $H_{app}= M_s$ along positive x-axis for different time steps (normalized). The simulation parameters are same as that in Figure \ref{fig:spice_vs_matlab}. (a) Small time step yields both accurate and stable solution. (b) Moderate time step gives a stable solution with kinks and missing precessions around the unstable pole at $m_z=\pm 1$, and also underestimates the reversal delay. (c) Large time step results in bounded and oscillatory solution.} \label{fig:sphLLG} \end{figure} \noindent We have already seen that time step is crucial to obtain an accurate and a stable solution. To get a smooth and continuous trajectory of $\mathbf{m}$, $|d\mathbf{m}|$ should be small for a given time step $dt$. Since, $\frac{d\mathbf{m}}{dt} \propto \mathbf{h}_{\mathrm{eff}}, \mathbf{i}_{\mathrm{s}}$, we get the constraint \begin{equation} \{|\mathbf{h}_{\mathrm{eff}}|, |\mathbf{i}_{\mathrm{s}}|\}dt\ll 1\quad \text{or} \quad \mathrm{max}\{|\mathbf{h}_{\mathrm{eff}}|, |\mathbf{i}_{\mathrm{s}}|\}dt\ll 1. \label{eq:tscons1} \end{equation} From (\ref{heff_1}), the magnitude of effective field is bounded by \begin{align*} |\mathbf{h}_{\mathrm{eff}}| &= \left|\mathbf{h}_{app} + \frac{H_k}{M_s} (\hat{\mathbf{n}}\cdot \mathbf{m})\hat{\mathbf{n}} - \sum_i N_i m_i + \mathbf{h}_T\right| \\ &\leq |\mathbf{h}_{app}| + \left|\frac{H_k}{M_s}(\hat{\mathbf{n}}\cdot \mathbf{m})\hat{\mathbf{n}}\right| + \left|\sum_i N_i m_i \right| + |\mathbf{h}_T| \\ &= |\mathbf{h}_{app}| + \frac{H_k}{M_s} + 1 + \nu = |\mathbf{h}_{app}| + \varepsilon \end{align*} where $\varepsilon = \frac{H_k}{M_s} + 1 + \nu$. Therefore, the constraint in (\ref{eq:tscons1}) becomes \begin{equation} \mathrm{max}\{|\mathbf{h}_{\mathrm{ap}}| + \varepsilon, |\mathbf{i}_{\mathrm{s}}|\}dt\ll 1. \label{eq:tscons2} \end{equation} Suppose we take the LHS in (\ref{eq:tscons2}) to be 10 times smaller than 1, we get \begin{equation} dt_{\mathrm{crit}} = \frac{0.1}{\mathrm{max}\{|\mathbf{h}_{\mathrm{ap}}|+\varepsilon, |\mathbf{i}_{\mathrm{s}}|\}}. \end{equation} Note that the critical time step doesn't depend on $\alpha$, when we have chosen a $dt$ which satisfies the wider constraint in (\ref{eq:tscons1}). For the simulation results in Figures \ref{fig:car_small} and \ref{fig:sph_small}, we have $\varepsilon = 0.1+1+0 = 1.1$ and $|\mathbf{h}_{ap}| = 1$, which gives $\varepsilon = 2.1$, and $dt_{\mathrm{crit}} = 0.045$. Hence, we get a smooth trajectory in Figures \ref{fig:car_small} and \ref{fig:sph_small} since the time step chosen is smaller than this critical time step. But this is not the case for Figures 10(b-c) and 11(b-c), which results in unboundedness, discontinuities, and oscillations. \section{Benchmarking with OOMMF} The Object Oriented MicroMagnetic Framework (OOMMF) by NIST is a widely used open source, portable, public domain tool for micromagnetics. The Oxs (OOMMF eXtensible Solver) is an extensible micromagnetic computation engine capable of solving problems defined on three-dimensional grids of rectangular cells holding three-dimensional spins \cite{Donahue}. We use OOMMF's Oxs with evolver class \textquotedblleft SpinXferEvolve" to simulate a single magnet whose magnetization evolves under the application of a spin current. Unfortunately there are no default evolver classes from OOMMF that include the effects of thermal noise. However, there are third-party extensions capableof performing thermal simulations in \begin{figure}[H] \hspace*{-15ex} \includegraphics[scale=0.38]{MATLABvsOOMMF} \vspace*{-5ex} \caption{\footnotesize The evolution of the magnetization vector under STT (without thermal effects) using our implicit midpoint implementation and the OOMMF solver. } \label{MATLABvsOOMMF} \end{figure} \noindent OOMMF. While the extension \textquotedblleft Xf$\_$ThermSpinXferEvolve" \cite{Fong} implements the solution of the s-LLGS equation using the Heun method, it naively renormalizes the magnetization norm in every iteration. The numerical accuracy and stability of this operation are questionable, and there appears to be no physical reason for this. In this section, we compare the magnetization vectors (from OOMMF) in the deterministic case with the corresponding vectors obtained from our implementation of the implicit midpoint method, as shown in Figure \ref{MATLABvsOOMMF}. This exercise substantiates the claim that our implicit midpoint realization is very close to the NIST standard. \section{Outlook and Conclusion} While a discussion of strong higher-order methods was not the goal of this paper, there are notable developments, whose applicability in solving the s-LLGS equation will be investigated in the future. Namely, the efficient strong order stochastic Runge-Kutta schemes which were proposed in \cite{Ro_2010_2}. These schemes have the advantage that the number of evaluations of the drift term of an SDE only increases linearly with the dimension of the driving Wiener process. This is in contrast to previous methods, which relied on a quadratically increasing number of evaluations of the drift term. For the approximation of the general multiple stochastic integrals, a method introduced in \cite{Wi_2001, Wi_2001_2} can be used.\\ \noindent Further, is important to note that many properties of magnets that are of interest to engineers can be solved for using weakly convergent schemes. For example, the switching probability of a magnet at an arbitrary time is such a property. Importantly, there are known schemes which converge in the weak sense with orders up to 4 \cite{Kloeden_Platen}. Recently, a very efficient explicit scheme with weak convergence order 2 has been introduced in \cite{Ro_2009, Ro_2010}. Many of these schemes can be simplified so that no iterated stochastic integrals have to be computed. This is a time-intensive process because many random variables have to be computed at each time step, and it is one of the reasons why efficient high strong-order schemes are currently only available for equations with special structure. In order to take advantage of these higher order approximations, the authors are currently investigating weakly convergent higher-order schemes for the s-LLGS equation.\\ \noindent Summarising, in this paper, we examine the accuracy of (i) implicit midpoint, (ii) Heun, (iii) Euler-Heun, and (iv) RK4-Heun numerical methods to solve the s-LLGS equation of macrospin dynamics. We compare the performance of these methods in terms of the average path-wise error, preservation of the magnetization norm, and 50\% switching boundary of macrospin reversal. We clearly show the higher accuracy of the implicit midpoint method as compared to other methods. However, the RK4-Heun method offers significant benefits in precision when the stochastic term in the s-LLGS equation is 2-3 orders of magnitude lower than the deterministic term. We also demonstrate that SPICE-based numerical integration schemes, such as trapezoidal, Gear, and Euler, when solving the cartesian s-LLGS equation, perform poorly and lead to significant errors in the results as compared to the implicit midpoint method at the same time step. We discuss the spherical coordinates implementation of the s-LLGS equation in SPICE as a possible solution to mitigate this problem, and also evaluate the critical time step required to obtain stable and accurate solutions for these implementations. Through an exhaustive study, we provide guidelines that will help researchers analyze macrospin dynamics more accurately in the context of appropriate stochastic calculus and the use of numerical integration method. \section*{Acknowledgements} \noindent The authors would like to acknowledge Prof. Supriyo Datta and Dr. Kerem Camsari from Purdue, and Nickvash Kani from Georgia Tech for the many useful discussions.
ff3b8bca236045dd4906c1b5bccf49a844b9f9ea
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} A class of infinite-dimensional simple Lie algebras was introduced by R. Block \cite{B}. Generalizations of Block algebras, usually referred to as {\it Block type Lie algebras}, partially because of which are closely related to the Virasoro algebra, Virasoro-like algebra, or special cases of (generalized) Cartan type $S$ Lie algerbas or Cartan type $H$ Lie algerbas \cite{X3}, have been extensively studied (cf.~\cite{DZ2,WS,X1}). In this paper, we are interested in the representation theory for a class of Block type Lie algebras. Let $G$ be an abelian group, $\F$ a field of characteristic $0$, $\mu: G\rightarrow \F$ an additive map, and $\varphi: G\times G\rightarrow \F$ a bilinear skew-symmetric map. A Block type Lie algebra \cite{DZ0} is by definition a vector space $L$ with basis $\{L_x, x\in G\}$ and the following brackets: \begin{equation*} [L_x,L_y]=\big(\varphi(x,y)+\mu(x-y)\big)L_{x+y},\ \ \ \forall x,y\in G. \end{equation*} Taking $G$ to be the product $\Gamma\times \Delta$ of two nonzero additive subgroups $\Gamma$ and $\Delta$ of $\F$ and $(p,q)\in\mathbb F^2$, define $\mu (\alpha,a)=pa-q\alpha$ and $\varphi\big((\alpha,a), (\beta,b)\big)=\beta a-\alpha b$ for any $(\alpha,a), (\beta, b)\in \Gamma\times \Delta$. Then we get a Lie algebra $\mathfrak{B}(\Gamma,\Delta,p,q)$ with basis $\{L_{\a,a}\,|\,(\a, a)\in\g\times\d\}$ subject to the following relations: \begin{eqnarray*}\label{In} \aligned &[L_{\a,a},L_{\b,b}]=\big((\b+p)(a+q)-(\a+p)(b+q)\big)L_{\a+\b,a+b}, \ \forall (\alpha,a), (\beta,b)\in \g\times \Delta. \endaligned \end{eqnarray*} Note that the center of $\mathfrak{B}(\Gamma,\Delta,p,q)$ is \begin{equation*} Z(\mathfrak{B}(\Gamma,\Delta,p,q))=\left\{\begin{array}{llll}\F L_{-p,-q}&\mbox{if \ }(-p,-q)\in\g\times\d,\\[4pt] 0&\mbox{otherwise. \ } \end{array}\right.\end{equation*} The object of study is the {\it Block type Lie algebra} $$\mathrm{B}(\g, \d, p, q):=[\mathfrak{B}(\Gamma,\Delta,p,q),\mathfrak{B}(\Gamma,\Delta,p,q)]/Z(\mathfrak{B}(\Gamma,\Delta,p,q))$$ with basis $\{L_{\a,a}\,|\,(\a, a)\in\widetilde{\Gamma\times\d}\}$, where \begin{equation*} \widetilde{\Gamma\times\d}=\left\{\begin{array}{llll} (\Gamma\times\d)\setminus\{(-p, -q),(-2p, -2q)\}&\mbox{if \ }p\in\g,\\[4pt] (\Gamma\times\d)\setminus\{(-2p, -2q)\}&\mbox{if \ }p\not\in\g, 2p\in\g,\\[4pt] \Gamma\times\d&\mbox{otherwise. \ } \end{array}\right.\end{equation*} Here we restrict $q$ to be in $\Delta$ and use the same notation $L_{\alpha,a}$ to denote its coset in the quotient space. $\mathrm{B}(\Gamma,\Delta,p,q)$ can also be obtained by extending the basis $\{\ L_{\a,i}\mid (\a,i)\in\g\times \Z_{\geq-1}\}$ of the Lie algebra $B$ \cite{WS} satisfying relations $[L_{\a,i},L_{\b,j}]=(\b(i+1)-\a(j+1))L_{\a+\b,i+j}$ to the basis indexed by $\g\times \Delta$ and replacing $0$ by $p$ and $1$ by $q$ in the defining relations. Note that a class of (centerless) Block type Lie algebras $\mathrm B(q)$ (cf. \cite{SXX}) is, in fact, a half part of $\mathrm{B}(\Z, \Z, 0, q)$ and that the so-called generalized (centerless) Virasoro algebra \begin{equation*}{\rm Vir}(\g)={\rm span}\{L_{\a,0}\,|\,\a\in\g\}\end{equation*} is contained in $\mathrm B(\g,\Delta, p, q) (q\neq0)$ as a subalgebra. The unique (up to a scalar multiplicative factor) semisimple element is $L_{0,0},$ but the gradation of $\mathrm B(\g,\d,p,q)$ with respect to $L_{0,0}$ is too coarse to distinguish basis elements via their eigenvalues. In view of this, it becomes nontrivial to study the representation theory of $ \mathrm B(\g,\d,p,q)$. In this paper, we shall study Verma modules over $\mathrm B(\Gamma,\Delta,p,q)$, which are not weight modules in general. After introducing a total order ``$\succ$'' on $\g$ we define a class of Verma $\mathrm B(\Gamma,\Delta,p,q)$-modules $M(\Lambda, \succ)$. The aim of this paper is to give the equivalent conditions of these Verma modules being irreducible. A reduced method is adopted in \cite{HWZ} to deal with weight modules, namely, the irreducibility of a Verma module is reduced to that of the Verma module over some of its ``integral" subalgebra in the case of discrete order and the length of some weight vector is reduced to $1$ in the case of dense order. This method also works for our case by making full use of certain ``filtrations" (see Proposition \ref{l+1} and Theorem \ref{theo0123}). This technique provides a way to study irreducibilities of non-weight modules. This paper is organized as follows. In Section $2$, we introduce some basic notions, especially we define Verma modules over $\mathrm{B}(\g, \d, p, q)$. Besides, for any $\t\in\g_+$, a useful proposition is given to characterise the irreducibility of the Verma $\mathrm{B}(\Z\t,\d, p, q)$-module $M_{\t}(\Lambda, \succ)$. In Section 3, we generalize the techniques developed in \cite{WS} to non-weight modules and completely determine the irreducibility of the Verma modules over $\mathrm{B}(\g,\d, p, q)$. Throughout the paper, we denote by $\Z,\ \N,\ \Z_{+}$ the sets of integers, positive integers and nonnegative integers, respectively. \section{Preliminaries} Observe easily that $\mathrm{B}(\g, \d, p, q)$ is $\g$-graded \begin{equation*} \mathrm{B}(\g, \d, p, q)=\raisebox{-5pt}{${}^{\, \, \displaystyle\oplus}_{\alpha\in \g}$}{\mathrm{B}(\g, \d, p, q)}_{\alpha},\ \ \ {\mathrm{B}(\g, \d, p, q)}_{\alpha}={\rm span} \{L_{\alpha,a}\mid a\in \d\}\mbox{ \ for }\a\in\g. \end{equation*} For a total order ``$\succ$'' on $\g$ compatible with its group structure, $\g=\g_+\cup \{0\}\cup \g_-$ with $\g_{\pm}=\{\a\in\g\,|\,\pm\a\succ0\}$. Denote $$\mathrm{B}(\g, \d, p, q)_\pm=\oplus_{\pm\a\succ0}\mathrm{B}(\g, \d, p, q)_{\a}.$$ Then $\mathrm{B}(\g, \d, p, q)$ admits the triangular decomposition \begin{equation*}\mathrm{B}(\g, \d, p,q)=\mathrm{B}(\g, \d, p, q)_+\oplus\mathrm{B}(\g, \d, p, q)_0\oplus\mathrm{B}(\g, \d, p, q)_-.\end{equation*} Also, fix a total order ``$\succ$'' on $\d$ compatible with its group structure (no confusion should arise here when the same symbol ``$\succ$'' is used to denote the total orders on $\g$ and $\d$). Given any $$\underline{\a}=(\a_1,\ldots, \a_r),\ \underline{\a}^\prime=(\underline{\a}_{1}^\prime,\ldots, \underline{\a}_{r}^\prime)\in\g_{+}^{r}\ {\rm and}\ \underline{a}=(a_1,\ldots, a_r),\ \underline{a}^\prime=(\underline{a}_{1}^\prime,\ldots, \underline{a}_{r}^\prime)\in\d^{r},$$ define \begin{eqnarray*} \underline{\a}\succ \underline{\a}^\prime &\Longleftrightarrow& \exists s\in \{1,\ldots, r\} \mbox { \ such that \ } \a_s\succ \a^\prime_{s} \mbox { \ and \ } \a_t= \a^\prime_{t} \mbox { \ for \ } t>s.\\ \underline{a}\succ \underline{a}^\prime &\Longleftrightarrow& \exists s\in \{1,\ldots, r\} \mbox { \ such that \ } a_s\succ a_{s}^{\prime }\mbox { \ and \ } a_t= a_{t}^\prime \mbox { \ for \ } t>s. \end{eqnarray*} Then $\g_{+}^{r}\times\d^{r}$ carries a natural total order: $$(\underline{\a}, \underline{a})\succ(\underline{\a}^\prime, \underline{a}^\prime)\Longleftrightarrow \underline{\a}\succ \underline{\a}^\prime \mbox { \ or \ } \underline{\a}=\underline{\a}^\prime,\ \underline{a}\succ \underline{a}^\prime,\quad \forall\ (\underline{\a}, \underline{a}), (\underline{\a}^\prime, \underline{a}^\prime)\in\g_{+}^{r}\times\d^{r},$$ which is also compatible with its addition operator. We use $U\big(\mathrm{B}(\g, \d, p, q)\big)$ to denote the universal enveloping algebra of $\mathrm{B}(\g, \d, p, q)$. Given any $\Lambda\in\mathrm{B}(\g, \d, p, q)_{0}^{*}$, the dual space of $\mathrm{B}(\g, \d, p, q)_{0}$, let $I(\Lambda, \succ)$ be the left ideal of $U\big(\mathrm{B}(\g, \d, p, q)\big)$ generated by the following elements $$\{L_{\a,a}\,\mid\,(\a, a)\in\g_{+}\times\d\}\cup\{h-\Lambda(h)\cdot1\,\mid\,h\in\mathrm{B}(\g, \d, p, q)_0\}. $$ Then the {\it Verma $\mathrm{B}(\g, \d, p, q)$-module} with respect to the order ``$\succ$" is defined as \begin{equation*}M(\Lambda, \succ)=U(\mathrm{B}(\g, \d, p, q))/I(\Lambda, \succ).\end{equation*} By the PBW theorem, $M(\Lambda, \succ)$ has a basis $\mathcal{B}$ consisting of all vectors of the form \begin{equation}\label{ssz}L_{-\a_1, a_1}L_{-\a_2, a_2} \cdots L_{-\a_r, a_r}v_\Lambda,\ \ \ \ (\a_{s}, a_{s})\in\g_{+}\times\d,\end{equation} where $v_\Lambda$ is the coset of $1$ in $M(\Lambda, \succ)$, $r\in\Z_{+}$ and $(\a_{s}, a_{s})\preceq(\a_{s+1}, a_{s+1})$ for all $s=1,\ldots,r-1$. Set $$\mathscr{L}(L_{-\a_1, a_1}L_{-\a_2, a_2} \cdots L_{-\a_r, a_r}v_\Lambda)=r\mbox{ \ (called the {\it length} of $L_{-\a_1, a_1}L_{-\a_2, a_2} \cdots L_{-\a_r, a_r}v_\Lambda$)}.$$ It is important to observe that the set $\{v\in\mathcal{B}\mid \mathscr L(v)=r\}$ inherits a total order from $\g_{+}^{r}\times\d^{r}$ in the following sense: $$L_{-\a_1, a_1}L_{-\a_2, a_2} \cdots L_{-\a_r, a_r}v_\Lambda \succ L_{-\a_1^\prime, a_1^\prime}L_{-\a_2^\prime, a_2^\prime} \cdots L_{-\a_r^\prime, a_r^\prime}v_\Lambda\Longleftrightarrow (\underline{\a}, \underline{a})\succ(\underline{\a}^\prime, \underline{a}^\prime),$$ where $(\underline{\a},\underline{a})=(\a_1,\ldots, \a_r,a_1,\ldots, a_r), (\underline{\a}^\prime, \underline{a}^\prime)=(\underline{\a}_{1}^\prime,\ldots, \underline{\a}_{r}^\prime, \underline{a}_{1}^\prime,\ldots, \underline{a}_{r}^\prime)\in\g_{+}^{r}\times\d^{r}$. For each $r\in\Z_{+}$, denote $$H_r={\rm span}\{v\in\mathcal{B}\mid \mathscr L(v)=r\}\quad {\rm and}\quad V_r=\oplus_{i=0}^r H_i.$$ It is easy to see that $L_{\a, a}V_r\subseteq V_r$ for any $(\a, a)\in\g_{+}\times\d$. Take $0\neq u=\mbox{$\sum\limits_{i=1}^n$}f_{i}u_i\in M(\Lambda, \succ)$ for some $f_i\in\F$ and $u_i\in \mathcal{B}$. Clearly, $u\in V_s\setminus V_{s-1}$ for some $s\in\Z_{+}$. Set $${\rm Supp\,}(u)=\{u_i\in H_s\mid f_i\neq0, i=1,\ldots,n\}\mbox{ \ (called the {\it support} of $u$)}.$$ We define the {\em leading term} ${\rm lt}(u)$ of $u$ as $f_{i_0}u_{i_0}$ if ${\rm max\,}{\rm Supp\,}(u)=u_{i_0}$. Note that $M(\Lambda, \succ)=\oplus_{\a\preceq0}M(\Lambda,\succ)_{\a}$ is a $\g$-graded $\mathrm{B}(\g, \d, p, q)$-module, where $M(\Lambda,\succ)_0=\F v_\Lambda$ and $M(\Lambda,\succ)_{\a}$ is spanned by $L_{-\a_1, a_1}L_{-\a_2, a_2} \cdots L_{-\a_k, a_k}v_\Lambda\in \mathcal{B}$ with $\a_1+\a_2+\cdots +\a_k=-\a$ for any $\a\in\g_-$. We call a nonzero vector $u\in M(\Lambda,\succ)_{\a}$ a {\it weight vector} with {\it weight} ${\rm wt}(u)=\a$, even though $M(\Lambda,\succ)$ might not be a weight module. For any $\t\in\g_{+}$, denote by $\mathrm{B}(\Z\t,\d,p,q)={\rm span} \{L_{n\t,a}\mid n\in\Z, a\in\d\}$ a subalgebra of $\mathrm{B}(\g, \d, $ $ p, q)$ and by $M_{\t}(\Lambda, \succ)$ the $\mathrm{B}(\Z\t,\d,p,q)$-submodule of $M(\Lambda, \succ)$ generated by $v_\Lambda$, respectively. The irreducibility of $M_{\t}(\Lambda, \succ)$ can be characterised in the following way. \begin{prop}\label{l+1} The Verma $\mathrm{B}(\Z\t,\d,p,q)$-module $M_{\t}(\Lambda, \succ)$ is irreducible if and only if $$\{u\in H_1\cap M_\t(\Lambda,\succ)\mid L_{\t,k}u=0,\forall k\in\d\}=0.$$ \end{prop} \begin{proof}\ Suppose that there exists $0\neq u\in H_1\cap M_\t(\Lambda,\succ)$ such that $L_{\t,k}u=0$ for any $k\in\d$. Then $\mathrm B(\Z\t,\d,p,q)_+u=0$, since $\{L_{\t,k}\mid k\in\d\}$ generates $\mathrm B(\Z\t,\d,p,q)_+$. It follows that $v_\Lambda\notin \oplus_{i\geq1}H_i \supseteq U\big(\mathrm B(\Z\t,\d,p,q)_-\big)u \cong U\big(\mathrm B(\Z\t,\d,p,q)\big)u.$ Thus, $U\big(\mathrm B(\Z\t,\d,p,q)\big)u$ is a proper submodule of $M_{\t}(\Lambda, \succ)$ and therefore $M_{\t}(\Lambda, \succ)$ is reducible. Conversely, the idea is essentially from \cite[Theorem 6.5]{GLZ}. But for our case a little modification is needed, since modules considered here are not weighted in general. For this, we introduce a family of subspaces of $M_\t(\Lambda,\succ)$. To be explicit, for any $(l, w)\in\Z_+\times\Z_+,$ define $$V(l,w)=\mbox{$\bigoplus\limits_{\mathfrak b\in \mathcal B (M_\t(\Lambda,\succ))\atop \mathscr L(\mathfrak b)\leq l, {\rm wt}(\mathfrak b)\leq - w\t}$}\mathbb F\mathfrak b.$ Suppose on the contrary that $M_{\t}(\Lambda, \succ)$ is reducible. Let $W$ be any nonzero submodule of $M_\t(\Lambda,\succ)$. Take $(l_0, w_0)\in \Z_+\times\Z_+$ minimal $((l,w)<(l^\prime,w^\prime)\Longleftrightarrow l< l^\prime\ {\rm or}\ l= l^\prime, w< w^\prime)$ such that $W\cap V(l_0,w_0)\neq 0$ and take $0\neq v\in W\cap V(l_0,w_0)$. It is clear from the definitions of length and weight that $w_0\geq l_0$. In fact, $w_0=l_0$. Suppose not, then the leading term of $v$ has the form $L_{-\t,a_1}\cdots L_{-\t,a_{r-1}}L_{-i_r\t,a_r}\cdots L_{-i_{l_0}\t, a_{l_0}}v_\Lambda$ with $i_r\geq 2$. On the one hand, by choosing some $a_0\in\d$ such that $(p-i_{r}\t)(a_0+q)\neq (\t+p)(a_{r}+q)$ we have \begin{eqnarray*} 0\neq {\rm lt\,}(L_{\t,a_0}v)\in \mathbb F L_{-\t,a_1}\cdots L_{-\t,a_{r-1}}L_{-(i_r-1)\t,a_r+a_0}\cdots L_{-i_{l_0}\t, a_{l_0}}v_\Lambda, \end{eqnarray*} but on the other hand, $L_{\t,a_0}v\in V(l_0,w_0-1)$, contradicting the choice of $(l_0,w_0)$. Write $$v=\mbox{$\sum\limits_{i=1}^r$}f_i v_i+\sum\limits_{i=r+1}^nf_iv_i+\tilde v+\overline v,$$ where $v_i\in {\rm Supp}(v) \ \cap M_\t(\Lambda,\succ)_{-l_0\t} (1\leq i\leq r)$ such that ${\rm min\,}{\rm Supp\,}(v)=v_{1}$, $v_i\in \mathcal B\big(M_{\t}(\Lambda, \succ)\big)\cap H_{l_0-1} \cap M_\t(\Lambda,\succ)_{-l_0\t}$ for $r+1\leq i\leq n$, $\tilde v\in (H_{l_0}\oplus H_{l_0-1})\cap \oplus_{\a\prec -l_0\t} M_\t(\Lambda,\succ)_\a$ and $\overline v\in V_{l_0-2}$. Assume that for $1\leq i\leq r$, each $v_i$ is of the form $v_i=L_{-\t,b_{i,1}}L_{-\t,b_{i,2}}\cdots L_{-\t,b_{i,l_0}}v_{\L}$. For any $k\in\d$ we compute \begin{eqnarray}\label{A11110} &&L_{\t,k}v_i\nonumber\\ &=& [L_{\t,k}, L_{-\t,b_{i,1}}L_{-\t,b_{i,2}}\cdots L_{-\t,b_{i,l_0}}]v_{\L}\nonumber\\ &\!\!\!=\!\!\!&\sum_{d=1}^{l_0} L_{-\t,b_{i,1}}\cdots L_{-\t,b_{i,d-1}}[L_{\t,k},L_{-\t,b_{i,d}}]L_{-\t,b_{i,d+1}}\cdots L_{-\t,b_{i,l_0}}v_{\L} \nonumber\\ &\!\!\!\equiv\!\!\!&\sum_{d=1}^{l_0}\L([L_{\t,k},L_{-\t,b_{i,d}}]) L_{-\t,b_{i,1}}\cdots L_{-\t,b_{i,d-1}}\widehat{L_{-\t,b_{i,d}}}L_{-\t,b_{i,d+1}}\cdots L_{-\t,b_{i,l_0}}v_{\L}+\sum_{d=1}^{l_0}\sum_{e=d+1}^{l_0}x_{i,e,d}(k) \nonumber\\ &\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!& L_{-\t,b_{i,1}}\cdots L_{-\t,b_{i,d-1}}\widehat{L_{-\t,b_{i,d}}}\cdots \widehat{L_{-\t,b_{i,e}}}\cdots L_{-\t,b_{i,l_0}}L_{-\t,b_{i,d}+b_{i,e}+k}v_{\L}\quad( {\rm mod\ } V_{l_0-2} ), \end{eqnarray} where `` $\widehat{~~~~~ }$" means deleting the factor and $x_{i,e,d}(k)\ (i=1,\ldots,r)$ are polynomial functions defined by $$\big[[L_{\t,k},L_{-\t,b_{i,d}}],L_{-\t,b_{i,e}}\big]=x_{i,e,d}(k)L_{-\t,b_{i,d}+b_{i,e}+k},\ \ \ \ \forall k\in\d.$$ For $r+1\leq i \leq n$, each $v_i$ is of the form $v_i=L_{-\t,b_{i,1}}L_{-\t,b_{i,2}}\cdots L_{-2\t,b_{i,l_0-1}}v_{\L}$. And we have \begin{eqnarray}\label{A111} \!\!\!\!\!\! L_{\t,k}v_i&\!\!\!=\!\!\!& [L_{\t,k}, L_{-\t,b_{i,1}}L_{-\t,b_{i,2}}\cdots L_{-2\t,b_{i,l_0-1}}]v_{\L}\nonumber\\ &\!\!\!\equiv\!\!\!&x_{i}(k)L_{-\t,b_{i,1}}L_{-\t,b_{i,2}}\cdots L_{-\t,b_{i,l_0-1}+k}v_{\L}\quad( {\rm mod\ } V_{l_0-2} ), \end{eqnarray} where $x_{i}(k)$ are polynomial functions defined by $$[L_{\t,k}, L_{-2\t,b_{i,l_0-1}}]=x_{i}(k)L_{-\t,b_{i,l_0-1}+k},\ \ \ \ \forall k\in\d.$$ Denote \begin{eqnarray*} w_1(k)\!\!\!\!&=&\sum_{i=1}^rf_i\sum_{d=1}^{l_0}\L([L_{\t,k},L_{-\t,b_{i,d}}]) L_{-\t,b_{i,1}}\cdots L_{-\t,b_{i,d-1}}\widehat{L_{-\t,b_{i,d}}}L_{-\t,b_{i,d+1}}\cdots L_{-\t,b_{i,l_0}}v_{\L}\quad{\rm and}\\ w_2(k)\!\!\!\!&=&\sum_{i=1}^rf_i\sum_{d=1}^{l_0}\sum_{e=d+1}^{l_0}x_{i,e,d}(k)L_{-\t,b_{i,1}}\cdots L_{-\t,b_{i,d-1}}\widehat{L_{-\t,b_{i,d}}}\cdots \widehat{L_{-\t,b_{i,e}}}\cdots L_{-\t,b_{i,l_0}}L_{-\t,b_{i,d}+b_{i,e}+k}v_{\L}\\ && +\sum_{i=r+1}^nf_i x_{i}(k)L_{-\t,b_{i,1}}L_{-\t,b_{i,2}}\cdots L_{-\t,b_{i,l_0-1}+k}v_{\L}. \end{eqnarray*}Rewrite $w_2(k)$ as $$\sum_{\underline{l}}x_{\underline{l}}(k)L_{-\t,l_{1}}L_{-\t,l_{2}}\cdots L_{-\t,l_{l_0-2}}L_{-\t,l_{l_0-1}+k}v_{\L},$$ where $x_{\underline l}(k)$ are polynomial functions in $k$. Then it follows from the choice of $v$, \eqref{A11110} and \eqref{A111} that for all $k\in\d$ we have \begin{eqnarray*}\label{A} 0&=&L_{\t,k}v=\sum_{i=1}^rf_iL_{\t,k}v_i+\sum_{i=r+1}^nf_iL_{\t,k}v_i+L_{\t,k}\tilde v+L_{\t,k}\overline v\nonumber\\ &\equiv& w_1(k)+w_2(k)+w_3(k)+w_4(k)\quad ( {\rm mod\ } V_{l_0-2} ), \end{eqnarray*} where $w_3(k)\in H_{l_0-1}$, $w_4(k)\in H_{l_0}$ such that $L_{\t,k}\tilde v\equiv w_3(k)+w_4(k)\ ({\rm mod}\ V_{l_0-2}).$ Then we must have $w_1(k)+w_2(k)+w_3(k)=0$, since $w_j(k)\in H_{l_0-1}$ for $j=1,2,3.$ Furthermore, $w_3(k)=0$ since the weight of $w_3(k)$ is different from that of $w_1(k)+w_2(k)$. Thus, $w_1(k)+w_2(k)=0$. It is important to observe that $w_1(k)$ and $w_2(k)$ are linearly independent whenever $k$ is large enough. So in particular, $x_{\underline{l}}=0$, i.e., $w_2(k)=0$, which in turn forces $w_1(k)=0$ for all $k\in\d$. This implies the coefficient of $L_{-\t,b_{1,1}}L_{-\t,b_{1,2}}\cdots L_{-\t,b_{1,l_0-1}}v_{\L}$ should be zero, i.e., $$\sum_{i\in I}r_if_i\L([L_{\t,k},L_{-\t,b_{i,l_0}}])=0, \ \ \ \ \forall k\in\d,$$where $I=\{1\leq i \leq r\mid b_{i,t}=b_{1,t},\forall 1\leq t \leq t_{0}-1\}$ and $r_i=\sharp \{1\leq j \leq l_0\mid b_{i,j}=b_{i,l_0}\}$. Then the vector $\sum_{i\in I}r_if_iL_{-\t,b_{i,l_0}}v_\Lambda$ lies in $\{u\in H_1\cap M_\t(\Lambda,\succ)\mid L_{\t,k}u=0,\forall k\in\d\}$ and is nonzero since $r_1f_1\neq0.$ This completes the proposition. \end{proof} \section{Main result} For a given order ``$\succ$'' on $\g$, set $\Phi(\a)= \{\b\in\g\mid 0\prec\b\prec\a\} \mbox{ \ for \ } \a\in\g_+$. The order ``$\succ$'' is called {\it dense} if $ \Phi(\a)\neq \emptyset \mbox{ \ for all \ } \a\in\g_+$ and otherwise, {\it discrete} if $\Phi(\t)=\emptyset \mbox{ \ for some \ } \t\in\g_+$. The aim of this paper is to find out the necessary and sufficient conditions for the Verma $\mathrm{B}(\g, \d, p, q)$-module $M(\Lambda, \succ)$ to be irreducible. \begin{theo}\label{theo0123} \begin{itemize}\parskip-3pt \item[(i)] With respect to a discrete order ``$\succ$" of $\g$, the Verma $\mathrm{B}(\g, \d, p, q)$-module $M(\Lambda, \succ)$ is irreducible if and only if $$\{u\in H_1\cap M_\t(\Lambda,\succ)\mid L_{\t,k}u=0,\forall k\in\d\}=0,$$ where $\tau={\rm min}\{\a\mid \a\in\g_+\}$. \item[(ii)] With respect to a dense order ``$\succ$" of $\g$, the Verma $\mathrm{B}(\g, \d, p, q)$-module $M(\Lambda, \succ)$ is irreducible if and only if $\Lambda\neq0$. Moreover, in the case $\Lambda=0$, the unique maximal proper submodule $$M^\prime(0, \succ)=\mbox{$\sum\limits_{k\geq1\atop 0\prec (\a_{1}, a_{1})\preceq\cdots \preceq(\a_{k}, a_{k})}$}\F L_{-\a_{1}, a_{1}}\cdots L_{-\a_k, a_k}v_{0}$$ of $M(0, \succ)$ is irreducible if and only if for any $x,y\in\g_{+}$, there exists a positive integer $n$ such that $nx\succ y$. \end{itemize} \end{theo} \begin{proof}(i) Suppose that the order ``$\succ$'' is discrete. Then $\Gamma=\Z\t\cup H_+\cup H_-,$ where $H_\pm=\{\a\in\g\mid \pm a\succ \Z\t\}.$ Suppose that $M_\t(\Lambda,\succ)$ is reducible. Let $S$ be a nonzero proper $\mathrm B(\Z\t,\d,p,q)$-submodule of $M_\t(\Lambda,\succ)$. Consider the $U(\mathrm B(\g,\d,p,q))$-submodule $W$ generated by $S$. Note by PBW theorem and the fact $U(\mathrm B(H_+,\d,p,q))S=0$ derived from the definition of $H_+$ that \begin{eqnarray*}W&=& U\big(\mathrm B(\g,\d,p,q)\big)S\\ &\cong& U(\mathrm B(H_-,\d,p,q))U(\mathrm B(\Z\t,\d,p,q))U(\mathrm B(H_+,\d,p,q))S\\ &\cong& U(\mathrm B(H_-,\d,p,q))S.\end{eqnarray*} Then $W$ is proper, since $v_\Lambda\notin W.$ This shows the reducibility of $M(\Lambda,\succ)$. Thus, we have established that the irreducibility of $M(\Lambda,\succ)$ implies that of $M_\t(\Lambda,\succ)$. Conversely, we show that the irreducibility of $M_\t(\Lambda, \succ)$ implies that of $M(\Lambda,\succ)$. Let $W$ be any nonzero $\mathrm B(\g,\d,p,q)$-submodule of $M(\Lambda,\succ)$. It is enough to show $W\cap M_\t(\Lambda,\succ)\neq 0$. Take $r\in\Z_+$ be minimal such that $W\cap V_r\neq0$. Choose any $0\neq y\in W\cap V_r$. Now it follows from the approach used in \cite{HWZ} (see also \cite{WS}), i.e., by repeatedly applying some $L_{h,k}\ (h\in H_+)$ to $y$, that we can get $w=y_r+\overline y\in U\big(\mathrm B(\g,\d,p,q)\big)y$ with $0\neq y_r\in M_\t(\Lambda,\succ)\cap H_r$ and $\overline y\in V_{r-1}$. We are done if $\overline y\in M_\t(\lambda,\succ)$. Otherwise, choosing $h^\prime\in H_+$ and $k^\prime\in \d$ and by using the approach again we may assume that $L_{h^\prime,k^\prime} \overline y\neq0.$ Then $L_{h^\prime,k^\prime}w=L_{h^\prime,k^\prime}\overline y,$ since $ L_{h^\prime,k^\prime} M_\t(\Lambda,\succ)=0$. Whence $0\neq L_{h^\prime,k^\prime} \overline y\in W\cap V_{r-1},$ contradicting the minimality of $r$. To sum up, we have proved that the $\mathrm{B}(\g, \d,p, q)$-module $M(\Lambda, \succ)$ is irreducible if and only if the $\mathrm{B}(\Z\t,\d,p,q)$-module $M_{\t}(\Lambda, \succ)$ is irreducible. Then (i) follows from Proposition \ref{l+1}. \bigskip (ii) Let us consider the case in which the order ``$\succ$'' is dense. Let $S$ be any nonzero submodule of $M(\Lambda, \succ).$ Take $r\in\Z_+$ minimal such that $0\neq S\cap V_{r}$. \begin{clai}\label{claim1}There exists some $0\neq v\in S$ such that \begin{equation*} v\equiv \mbox{$\sum\limits_{\underline k=(k_1, \ldots, k_r)\in\d^r}$}f_{\underline{k}}L_{-\varepsilon_1, k_1}\cdots L_{-\varepsilon_r, k_r}v_\Lambda\quad( {\rm mod\ } V_{r-1} ),\end{equation*} where $0\prec \varepsilon_1 \prec\cdots\prec \varepsilon_r$ and $f_{\underline{k}}\in\F$. \end{clai} Choose any $0\neq u\in S\cap V_{r}$ and write \begin{equation*}u\equiv\mbox{$\sum\limits_{(\underline \a,\underline a)\in \g_+^r\times \d^r \atop0\prec(\a_{1}, a_{1})\preceq(\a_{2}, a_{2})\preceq\cdots\preceq(\a_{r}, a_{r}) }$}f_{\underline{\a}, \underline{a}} L_{-\a_1, a_1}L_{-\a_2, a_2} \cdots L_{-\a_{r}, a_{r}}v_\Lambda\quad ( {\rm mod}\ V_{r-1} ) \mbox{ \ for some \ } f_{\underline{\a}, \underline{a}}\in\F.\end{equation*} Without loss of generality, we assume that $r>1$, since it can be treated in a similar but simpler way for the case $r=1$. Denote \begin{eqnarray*}& I=\{(\underline{\a}, \underline{a})=(\a_1,\ldots,\a_r,a_1,\ldots,a_r)\in\g_+^r\times\d^r\mid f_{\underline{\a}, \underline{a}}\neq0\}\\ &{\rm and}\quad \gamma_1={\rm max}\{\a_r\mid (\underline\a,\underline a)\in I\}.\end{eqnarray*} Since ``$\succ$'' is dense, there exists some $\varepsilon_1\in\g_+$ such that $$\varepsilon_1\prec{\rm min} \{\a_1\mid (\underline\a,\underline a)\in I\}$$ $${\rm and}\quad\{x\in\g_+\mid \gamma_1-\varepsilon_1\prec x \prec \gamma_1\}\cap \{\a_r, \a_{r-1}\mid (\underline{\a}, \underline{a})\in I\}=\emptyset.$$ Note that the choice of $\v_1$ forces $\gamma_1-\varepsilon_1-\a_k\succ0$ if $\a_k\neq \gamma_1, k\in\{1,\ldots,r\}$. For any $k_1^\prime\in\d$, we see that \begin{eqnarray}\label{L-action} \!\!\!\!\!\! u_0&\!\!\!:=\!\!\!& L_{\gamma_1-\varepsilon_1, k_1^\prime} u\nonumber\\ &\!\!\!\equiv\!\!\!&\mbox{$\sum\limits_{(\underline\a,\underline a)\in I}$} f_{\underline\a,\underline a}L_{\gamma_1-\varepsilon_1, k_1^\prime} L_{-\a_1, a_1}\cdots L_{-\a_{r}, a_{r}}v_\Lambda\ ( {\rm mod\ } V_{r-1} )\nonumber\\ &\!\!\!\equiv\!\!\!&\mbox{$\sum\limits_{(\underline\a,\underline a)\in I^\prime}$}\sum_{l=s_{\underline\a}}^r f_{\underline\a,\underline a} L_{-\a_1, a_1}\cdots L_{-\a_{s_{\underline\a}-1},a_{s_{\underline\a}-1}} L_{-\gamma_1,a_{s_{\underline\a}}}\cdots[L_{\gamma_1-\varepsilon_1, k_1^\prime}, L_{-\gamma_1,a_l}]\cdots L_{-\gamma_1, a_{r}}v_\Lambda\ ( {\rm mod\ } V_{r-1} )\nonumber\\ &\!\!\!\equiv&\!\!\!\!\mbox{$\sum\limits_{(\underline\a,\underline a)\in I^\prime}$}\sum_{l=s_{\underline\a}}^r f_{\underline\a,\underline a}^{\prime} L_{-\varepsilon_1, {k_1^\prime+a_l}} L_{-\a_1, a_1}\cdots L_{-\gamma_1,a_{s_{\underline\a}}}\cdots \widehat{L_{-\gamma_1,a_l}}\cdots L_{-\gamma_1, a_{r}}v_\Lambda\quad ( {\rm mod\ } V_{r-1} ), \end{eqnarray} where \begin{eqnarray*}& I^\prime=\{(\underline\a,\underline a)\in I\mid \exists 1\leq s_{\underline\a}\leq r\ {\rm such\ that}\ \a_{s_{\underline\a}-1}\prec\gamma_1, \a_{s_{\underline\a}}=\a_{s_{\underline\a}+1}=\cdots =\a_r=\gamma_1\}\\ &{\rm and\ }\quad f_{\underline\a,\underline a}^{\prime}=(r-s_{\underline \a}+1)f_{\underline\a,\underline a}\big((p-\gamma_1)(k_1^\prime+q)-(\gamma_1-\varepsilon_1+p)(a_l+q)\big).\end{eqnarray*} Note that the terms $$L_{-\varepsilon_1, {k_1^\prime+a_l}} L_{-\a_1, a_1}\cdots L_{-\gamma_1,a_{s_{\underline\a}}}\cdots \widehat{L_{-\gamma_1,a_l}}\cdots L_{-\gamma_1, a_{r}}v_\Lambda$$ occur in \eqref{L-action} all have the standard form as in \eqref{ssz} and therefore are linearly independent. Hence by choosing suitable $k_1^\prime$ we know the resulting vector $u_0\neq0.$ It is worthwhile to point out that both choices of $\varepsilon_1$ and $k_1^\prime$ are infinite. Repeating the preceding procedure we see that there exist $(\gamma_2,\ldots,\gamma_r)\in\g_+^{r-1}$, $(k_2^\prime,\ldots,k_r^\prime)\in\d^{r-1}$ and $0\prec\varepsilon_2\prec\cdots\prec\varepsilon_r$ such that \begin{eqnarray*}0&\neq& L_{\gamma_r-\varepsilon_r,k_r^\prime}\cdots L_{\gamma_2-\varepsilon_2,k_2^\prime}u_0\\ &\equiv&\sum_{\underline k=(k_1,\ldots,k_{r})\in\d^r}f_{\underline k}L_{-\varepsilon_1,k_1}L_{-\varepsilon_2,k_2}\cdots L_{-\varepsilon_r,k_r}v_{\Lambda}\quad ( {\rm mod\ } V_{r-1} ).\end{eqnarray*} Indeed, $\varepsilon_2$ can be chosen to satisfy $\varepsilon_1\prec\varepsilon_2.$ So the Claim \ref{claim1} is true. \begin{clai}\label{clai2} There exists some $0\neq w\in S$ such that \begin{equation*} w\equiv L_{-\varepsilon_1, k_1}\cdots L_{-\varepsilon_r, k_r}v_\Lambda\quad ({\rm mod\ } V_{r-1} ),\end{equation*} where $0\prec \varepsilon_1 \prec\cdots\prec \varepsilon_r$ and $(k_1,\ldots,k_r)\in(\d\setminus\{-q\})^r$. \end{clai} Choose $$0\neq w\equiv \mbox{$\sum\limits_{\underline k=(k_1,\ldots,k_r)\in \d^r}$}f_{\underline{k}}L_{-\varepsilon_1, k_1}\cdots L_{-\varepsilon_r, k_r}v_\Lambda\quad ( {\rm mod\ } V_{r-1} )$$ as in Claim \ref{claim1} such that $\sharp {\rm Supp}(w)$ is minimal. The claim is equivalent to showing that $\sharp {\rm Supp}(w)=1.$ Suppose on the contrary that $\sharp {\rm Supp}(w)\geq2$. Let ${\rm lt\,}(w)=f_{\underline{b}}L_{-\varepsilon_1, b_1}\cdots L_{-\varepsilon_r, b_r}v_\Lambda$. Then $\{1\leq i\leq r\mid k_i\neq b_i, f_{\underline{k}}\neq0\}$ is not an empty set. Denote $s={\rm max}\{1\leq i \leq r\mid k_i\neq b_i, f_{\underline{k}}\neq0\}$ and choose $L_{-\varepsilon_1, k_1^*}\cdots L_{-\varepsilon_s, k_s^*}L_{-\varepsilon_{s+1}, b_{s+1}}\cdots L_{-\varepsilon_r, b_r}v_\Lambda\in {\rm Supp}(w)$ such that $k_{s}^{*}\neq b_s$. Note that $w$ has the form \begin{eqnarray*} w&\!\!\!\equiv\!\!\!&f_{\underline{b}}L_{-\varepsilon_1, b_1}\cdots L_{-\varepsilon_s, b_s}\cdots L_{-\varepsilon_r, b_r}v_\Lambda+ \nonumber\\ &\!\!\!\!\!\!&f_{\underline{k}^{*}}L_{-\varepsilon_1, k_{1}^{*}}\cdots L_{-\varepsilon_s, k_{s}^{*}}\cdots L_{-\varepsilon_r, b_r}v_\Lambda +\mbox{ \ other terms }\quad ({\rm mod\ } V_{r-1} ), \end{eqnarray*} where $f_{\underline{b}}f_{\underline{k}^{*}}\neq0$. Choose $\delta_r\in \g_+$ such that ${\rm max}\{\epsilon_r-\epsilon_1, \epsilon_{r-1}\}\prec \delta_r\prec \epsilon_r$ and $b_r(\varepsilon_{r}-p)\neq (p+\delta_{r})(b_r+q)$. We have \begin{eqnarray*} w_r&\!\!\!:=& L_{\delta_r, -b_r-q}w\nonumber\\ &\equiv\!\!\!&L_{\delta_r, -b_r-q}\Big\{f_{\underline{b}}L_{-\varepsilon_1, b_1}\cdots L_{-\varepsilon_s, b_s}\cdots L_{-\varepsilon_r, b_r}v_\Lambda+f_{\underline{k}^{*}}L_{-\varepsilon_1, k_{1}^{*}}\cdots L_{-\varepsilon_s, k_{s}^{*}}\cdots L_{-\varepsilon_r, b_r}v_\Lambda\nonumber\\ &&+\mbox{ \ other terms}\Big\} \ ( {\rm mod\ } V_{r-1} )\nonumber\\ &\equiv&f_{\underline{b}}^{r} L_{-\varepsilon_r+\delta_r, -q}L_{-\varepsilon_1, b_1}\cdots L_{-\varepsilon_s, b_s}\cdots L_{-\varepsilon_{r-1}, b_{r-1}}v_\Lambda+ \nonumber\\ && f_{\underline{k}^{*}}^{r}L_{-\varepsilon_r+\delta_r, -q}L_{-\varepsilon_1, k_{1}^{*}}\cdots L_{-\varepsilon_s, k_{s}^{*}}\cdots L_{-\varepsilon_{r-1}, b_{r-1}} v_\Lambda +\mbox{ \ other terms}\quad ({\rm mod\ } V_{r-1} ), \end{eqnarray*}where $f_{\underline{b}}^{r}f_{\underline{k}^{*}}^{r}\neq0$. Inductively, there exists $(\delta_{s},\ldots,\delta_{r-1})\in\g_+^{r-s}$ such that \begin{equation*} {\rm max}\{\varepsilon_{i-1},\epsilon_i+\delta_{i+1}-\epsilon_{i+1}\}\prec \delta_{i} \prec \varepsilon_{i} \quad{\rm and}\quad b_i(\varepsilon_{i}-p)\neq(\delta_{i}+p)(b_i+q). \end{equation*} Then \begin{eqnarray*} w_{s}&\!\!\!:=& L_{\delta_{s}, -b_{s}-q}L_{\delta_{s+1}, -b_{s+1}-q}\cdots L_{\delta_{r-1}, -b_{r-1}-q}w_r\nonumber\\ &\equiv&f_{\underline{b}}^{s}L_{-\epsilon_s+\delta_s,-q}\cdots\cdots\cdot\cdot L_{-\varepsilon_r+\delta_r, -q}\ L_{-\varepsilon_1, b_1}\cdots L_{-\varepsilon_{s-1}, b_{s-1}}v_\Lambda+ \nonumber\\ && f_{\underline{k}^{*}}^{s}L_{-\epsilon_s+\delta_s,k_s^{*}-b_s-q}\cdots L_{-\varepsilon_r+\delta_r, -q}\ L_{-\varepsilon_1, k_{1}^{*}}\cdots L_{-\varepsilon_{s-1}, k_{s-1}^*}v_\Lambda +\mbox{ \ other terms}\quad ({\rm mod\ } V_{r-1} ). \end{eqnarray*} Similarly, choosing suitable $(\delta_1,\ldots,\delta_{s-1})\in\g_+^{s-1}$ and $(k^\prime_1,\ldots,k^\prime_{s-1})\in \Delta^{s-1}$ we get another nonzero vector: \begin{eqnarray*} w_0\!\!\!&:=&L_{\delta_{s-1},k^\prime_{s-1}}\cdots L_{\delta_1,k^\prime_1}w_s\\ &\equiv&f_{\underline{b}}^{0} L_{-\varepsilon_1+\delta_1, b_1+k^\prime_1}\cdots L_{-\varepsilon_{s-1}+\delta_{s-1}, b_{s-1+k^\prime_{s-1}}} L_{-\epsilon_s+\delta_s,-q}\cdots L_{-\varepsilon_r+\delta^r, -q}\v_\Lambda+ \nonumber\\ && f_{\underline{k}^{*}}^{0}L_{-\varepsilon_1+\delta_1, k_{1}^{*}+k^\prime_1}\cdots L_{-\varepsilon_{s-1}+\delta_{s-1}, k_{s-1}^*+k^\prime_{s-1}} L_{-\epsilon_s+\delta_s,k_s^{*}-b_s-q}\cdots L_{-\varepsilon_r+\delta^r, -q}v_\Lambda\\ && +\mbox{ \ other terms \ } ( {\rm mod\ } V_{r-1} ). \end{eqnarray*}In fact, we can require that both $f_{\underline{b}}^{0}f_{\underline{k}^{*}}^{0}\neq0$ and $0\prec\epsilon_1-\delta_1\prec \epsilon_2-\delta_2\prec\cdots\prec \epsilon_r-\delta_r$ hold. At this moment, choose $-p\neq\delta\in\g_+$ such that $ \epsilon_{s-1}-\delta_{s-1}\prec \delta\prec \epsilon_s-\epsilon_{s-1}-\delta_s+\delta_{s-1}$, now we have \begin{eqnarray*} 0&\neq& L_{\delta,-q}w_{0}\nonumber\\ &\equiv&f_{\underline{k}^{*}}^{\prime}L_{-\varepsilon_1+\delta_1, k_{1}^{*}+k^\prime_1}\cdots L_{-\varepsilon_{s-1}+\delta_{s-1}, k_{s-1}^*+k^\prime_{s-1}} L_{-\epsilon_s+\delta_s+\delta,k_s^{*}-b_s-2q}\cdots L_{-\varepsilon_r+\delta_r, -q}v_\Lambda\quad (f_{\underline{k}^{*}}^{\prime}\neq0)\\ && +\mbox{ \ other terms}\quad ({\rm mod\ } V_{r-1} ). \end{eqnarray*} Note that $L_{\delta,-q}w_{0}\in V_r\setminus V_{r-1}$ and $\sharp{\rm Supp}(L_{\delta,-q}w_{0})<\sharp{\rm Supp}(w)$, contradicting the minimality of $\sharp{\rm Supp}(w)$. Hence, $\sharp {\rm Supp}(w)=1.$ Finally, using the same arguments as in Claim \ref{claim1} and choosing suitable $k_1^\prime,\ldots,k_r^\prime$, one can assume that $(k_1,\ldots,k_r)\in(\d\setminus\{-q\})^r$. Thus our claim holds. \begin{clai}\label{clai3} $r\leq 1$. In fact, there exists some $\pm p\neq\eta\in\g_+$ such that $L_{-\eta, k}v_\Lambda\in S$ for all $k\in\d\setminus\{-2q\}$. \end{clai} Suppose that $r\geq 2$. By Claim \ref{clai2}, there exists some $0\neq w\in S$ such that $$w= L_{-\varepsilon_1, k_{1}}\cdots L_{-\varepsilon_r, k_{r}}v_\Lambda+\mbox{$\sum\limits_{0\leq l<r\atop 0\prec(\a_{1}, a_{1})\preceq(\a_{2}, a_{2})\preceq\cdots\preceq(\a_{l}, a_{l})}$}g_{\underline{\a}, \underline{a}}L_{-\a_{1}, a_{1}}\cdots L_{-\a_l, a_l}v_\Lambda,$$ where $0\prec \varepsilon_1\prec\cdots\prec \varepsilon_r$, $(k_1,\ldots,k_r)\in(\d\setminus\{-q\})^r$ and $g_{\underline{\a}, \underline{a}}\in\F$. Denote $I_0=\{\underline{\a}=(\a_1, \ldots, \a_l)\mid g_{\underline{\a}, \underline{a}}\neq0 \mbox { \ for some \ } \underline{a}\}$. Let $\lambda=-\sum_{i=1}^r\epsilon_i$. Fix some $b\in\d$, for any $0\prec\varepsilon\prec{\rm min}\{\varepsilon_1, \a_1\mid \underline{\a}\in I_0\}$ one has \begin{eqnarray*}\label{12345} \!\!\!\!\!\! L_{-\lambda-\varepsilon, b}w&\!\!\!=\!\!\!&L_{-\lambda-\varepsilon, b}\{ L_{-\varepsilon_1, k_{1}}\cdots L_{-\varepsilon_r, k_{r}}v_\Lambda+\mbox{$\sum\limits_{0\leq l<r\atop 0\prec(\a_{1}, a_{1})\preceq(\a_{2}, a_{2})\preceq\cdots\preceq(\a_{l}, a_{l})}$}g_{\underline{\a}, \underline{a}}L_{-\a_{1}, a_{1}}\cdots L_{-\a_l, a_l}v_\Lambda\} \nonumber\\ &=&\left\{F(-\lambda-\varepsilon)+\mbox{$\sum\limits_{a_{1}+\cdots+a_{l}=k_{1}+\cdots +k_{r}\atop \a_1+\cdots+\a_l=\lambda}$}g_{\underline{\a}, \underline{a}}H_{\underline{\a}, \underline{a}}(-\lambda-\varepsilon)\right\}L_{-\varepsilon, b+k_{1}+\cdots +k_{r}}v_\Lambda \nonumber\\ &&~~~~+\mbox{$\sum\limits_{a_{1}+\cdots+a_{l}\neq k_{1}+\cdots +k_{r}\atop{\rm or}\ \a_1+\cdots+\a_l\neq \lambda}$}v_{-\a_1-\cdots-\a_l-\lambda-\v}, \end{eqnarray*} where each $v_{-\a_1-\cdots-\a_l-\lambda-\v}\in V_{r-1}$ is some weight vector with weight $-(\a_1+\cdots+\a_l+\lambda+\v)$, $F(x)$ and $H_{\underline{\a}, \underline{a}}(x)$ are polynomials given as follows: \begin{equation*} F(x)=\left|\begin{array}{cc} b+q & k_{1}+q\\ x+p & p-\varepsilon_{1} \end{array}\right| \left|\begin{array}{cc} b+q+k_{1} & k_{2}+q\\ x+p-\varepsilon_{1} & p-\varepsilon_{2} \end{array}\right| \cdots \left|\begin{array}{cc} b+q+k_{1}+\cdots +k_{r-1} & k_{r}+q\\ x+p-\varepsilon_{1}-\cdots-\varepsilon_{r-1} & p-\varepsilon_{r} \end{array}\right|. \end{equation*} \begin{equation*} H_{\underline{\a}, \underline{a}}(x)=\left|\begin{array}{cc} b+q & a_{1}+q\\ x+p & p-\a_{1} \end{array}\right| \left|\begin{array}{cc} b+q+a_{1} & a_{2}+q\\ x+p-\a_{1} & p-\a_{2} \end{array}\right| \cdots \left|\begin{array}{cc} b+q+a_{1}+\cdots +a_{l-1} & a_{l}+q\\ x+p-\a_{1}-\cdots-\a_{l-1} & p-\a_{l} \end{array}\right|. \end{equation*} Since ${\rm deg}F(x)=r>{\rm deg}H_{\underline{\a}, \underline{a}}(x)$, we can find some $\v\in\g_+$ with $\varepsilon\prec{\rm min}\{\varepsilon_1, \a_1\mid \underline{\a}\in I_0\}$ such that $$F(-\lambda-\varepsilon)+\mbox{$\sum\limits_{a_{1}+\cdots+a_{l}=k_{1}+\cdots +k_{r}\atop \a_1+\cdots+\a_l=\lambda}$}g_{\underline{\a}, \underline{a}}H_{\underline{\a}, \underline{a}}(-\lambda-\varepsilon)\neq0\quad{\rm and\ thereby}\quad L_{-\lambda-\varepsilon, b}w\neq 0.$$ But then a contradiction arises, since $L_{-\lambda-\varepsilon, b}w\in V_{r-1}$. So $r\leq 1$, as desired. Consider the case $r=1$. Then by the reduced methods used as in Claims \ref{claim1} and \ref{clai2} we may assume that $\sharp{\rm Supp}(w)=1$. Thus, we can choose a nonzero vector of form $L_{-\eta^\prime,b}v_\Lambda+fv_\Lambda\in S$, where $\eta^\prime\in\g_+$, $b\in\Delta$ and $f\in \mathbb F$. Taking $p\neq\zeta\in\g_+$ and $\pm p\neq\eta\in\g_+$ such that $\eta\prec\zeta\prec\eta^\prime$ and $b(\eta^\prime-p)\neq(\eta^\prime-\zeta+p)(b+q)$, one has \begin{eqnarray*} &\!\!\!\!\!\!& L_{\zeta-\eta, k+q}L_{\eta^\prime-\zeta, -q-b}\big(L_{-\eta^\prime, b}v_{\Lambda}+fv_\Lambda\big)\nonumber\\ &\!\!\!\!\!\!&= (p-\zeta)(k+2q)\big(b(\eta^\prime-p)-(\eta^\prime-\zeta+p)(b+q)\big)L_{-\eta, k}v_{\Lambda}\in S, \end{eqnarray*} which means $L_{-\eta, k}v_{\Lambda}\in S$ for any $k\in\d\setminus\{-2q\}$. This completes Claim \ref{clai3}. \bigskip Now we are ready to prove the first statement. For the case $q\neq0$, by Claim \ref{clai3} we have \begin{eqnarray*} &L_{\eta, k+q}L_{-\eta, -q}v_{\Lambda}=(p-\eta)(k+2q)\Lambda(L_{0, k})v_{\Lambda} \mbox{ \ for \ any\ } k\in\d\\ &{\rm and}\quad L_{\eta, -3q}L_{-\eta,q}v_{\Lambda}=-4pq\Lambda(L_{0, -2q})v_{\Lambda}, \end{eqnarray*} which mean $v_{\Lambda}\in S$ if $\Lambda\neq0$. Similarly, for the case $q=0$ one has \begin{eqnarray*} &L_{\eta, 0}L_{-\eta,k}v_{\Lambda}=-k(p+\eta)\Lambda(L_{0, k})v_{\Lambda} \mbox{ \ for\ any \ } k\in\d\setminus\{0\},\\ &{\rm and}\quad L_{\eta, 1}L_{-\eta,-1}v_{\Lambda}=2p\Lambda(L_{0, 0})v_{\Lambda}, \end{eqnarray*} which also force $v_{\Lambda}\in S$ if $\Lambda\neq0$. Thus in either case $M(\Lambda, \succ)$ is irreducible. Conversely, it is obvious that $$M^\prime(0, \succ)=\mbox{$\sum\limits_{k\geq1\atop 0\prec (\a_{1}, a_{1})\preceq\cdots \preceq(\a_{k}, a_{k})}$}\F L_{-\a_{1}, a_{1}}\cdots L_{-\a_k, a_k}v_{\Lambda}$$ is the unique maximal proper $U\big(\mathrm{B}(\g, \d, p, q)\big)$-submodule if $\Lambda=0$. Now suppose that there exists a pair $(x, y)\in\g_{+}\times \g_+$ such that $n x\preceq y$ for all $n\in\mathbb N$. It is easy to verify that $$W^\prime=\sum_{x\succ z\in\g_+, k\in\Delta}U\big(\mathrm{B}(\g, \d, p, q)\big)L_{-z, k}v_{0}$$ is a submodule of $M^\prime(0, \succ)$, which is proper since $L_{-y, k}v_{0}\not\in W^\prime$ for any $k\in\d$. Conversely, assume that for any $x, y\in\g_+$, there exists $n\in\mathbb N$ such that $nx\succ y$. Let $S^\prime$ be any nonzero submodule of $M^\prime(0,\succ)$. By Claim 3, there exists some $\eta\in\g_+$ such that $L_{-\eta, k}v_{0}\in S^\prime$ for $k\in\d\setminus\{-2q\}$. Furthermore, thanks to the density of ``$\succ$'', we can choose $\eta$ such that $p\notin\N\eta$. According to our assumption, for any $z\in\g_{+}$, there exists some $n\in\N$ such that $n\eta\succ z$. If $q\neq0$, then $$q^{n-1}(p-\eta)\cdots (p-(n-1)\eta)L_{-n\eta, -q}v_{0}=[\ \,\stackrel{n-1}{\overbrace{\!\!\!L_{-\eta, 0} ,[L_{-\eta, 0},...,[L_{-\eta, 0},\!\!\!\!\!\!\!}}\ \ \ \ L_{-\eta, -q}]...]]v_0\in S^\prime,$$ from which we get $L_{-n\eta, -q}v_{0}\in S^\prime$. Then we have \begin{equation*}\label{equation1} (p-n\eta)(k+2q)L_{-z, k}v_{0}=L_{n\eta-z, k+q} L_{-n\eta, -q}v_{0}\in S^\prime\, \mbox{ \ for \ } k\in\d. \end{equation*} Thus, $L_{-z, k}v_{0}\in M^\prime(0, \succ)$ for all $z\in\g_{+}$ and $k\in\d\setminus\{-2q\}$. While for $k=-2q$, \begin{equation*}\label{equation11}2q(z-2p)L_{-z, -2q}v_{0}=L_{z, -3q} L_{-2z, q}v_{0}\in S^\prime.\end{equation*} To sum up, $L_{-z,k}v_0\in S^\prime$ for any $z\in\g_+$ and $k\in\d$. We are going to show this also holds for the other case $q=0$. Assume that the $n$ above is even. On the one hand, by a direct calculation we obtain \begin{eqnarray*} && (-2)^{\frac{n}{2}}\prod_{i=1}^{\frac{n}{2}}(p-i\eta)\prod_{j=1}^{\frac{n}{2}}(p-2j\eta)L_{-n\eta, 0}v_{0}\nonumber\\ &=& [\ \,\stackrel{n-1}{\overbrace{\!\!\!L_{-\eta, -1} ,[L_{-\eta, 1},...,[L_{-\eta, 1},[L_{-\eta, -1},\!\!\!\!\!\!\!}}\ \ \ \ L_{-\eta, 1}]]...]]v_{0}\in S^\prime, \end{eqnarray*} forcing $L_{-n\eta, 0}v_{0}\in S^\prime$, from which we get \begin{equation*}\label{equation2} (p-n\eta)kL_{-z, k}v_{0}=L_{n\eta-z, k} L_{-n\eta, 0}v_{0}\in S^\prime\, \mbox{ \ for \ } k\in\d. \end{equation*}And we also have \begin{equation*}\label{equation3} (2p-z)L_{-z, 0}v_{0}=L_{z, 1} L_{-2z, -1}v_{0}\in S^\prime. \end{equation*} So in either case, we have obtained that $L_{-z, k}v_{0}\in S^\prime$ for any $z\in\g_+$ and $k\in\d$. This implies $M^\prime(0, \succ)\subseteq S^\prime$, proving the irreducibility of $M^\prime(0, \succ)$. \end{proof}
47dc44b04a243ece2883d554ba8e78e04ab3b68b
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Results and Discussion} { Consider the optical response of a symmetric quadrumer when it is excited by a plane wave, whose propagation direction and electric field polarization lie in the plane of the quadrumer ($s$-polarization), as shown in Fig.~\ref{fig:schematic}a. } \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.5\textwidth]{schematic.pdf}} \caption{(a) Schematic of the problem: a symmetric nanosphere quadrumer is excited by an in-plane $s$-polarized plane wave {(both propagation and polarization vectors lie in the plane of the quadrumer)}. (b) The combined magnetic response: a collective optically-induced magnetic resonance of the whole structure is accompanied by the magnetic responses of the individual particles. } \label{fig:schematic} \end{figure} The conventional magnetic response to such an excitation is supported by a collective circulation of electric displacement current around all four nanoparticles. However, if the nanoparticles are made of a high-index dielectric material such as silicon, the individual nanoparticles are also able to couple directly with the applied magnetic field, sustaining internal circulating polarization currents. In this sense, a silicon quadrumer can produce magnetic responses to both electric {\it and} magnetic components of the applied plane wave, as is depicted in Fig.~\ref{fig:schematic}b. { In this figure, the red arrows denote an electric dipole, and the blue circular arrows denote the circulation of displacement current that supports an out-of-page magnetic dipole. Importantly, } an interplay may be induced between the two magnetic responses through the locally-enhanced, $z$-polarized magnetic field sustained by the collective circulation of displacement current, which establishes a coupling channel to the magnetic responses of the individual nanoparticles. We begin by understanding this magnetic interplay, for which we will model the interactions between nanoparticles using the {\it Coupled Electric and Magnetic Dipole Approximation}~\cite{Mulholland1994} (CEMDA). The mathematical description of this dipole model in free space is based on the following equations: \begin{subequations} \label{eq:dipole equations} \begin{align} \mathbf{p}_{i} = \alpha_{E}\epsilon_{0}\mathbf{E_{0}}(\mathbf{r}_{i})& +\alpha_{E}k^{2} \left(\underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{p}_{j} -\frac{1}{c_0}\,\nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{m}_{j}\right) ,\\ \mathbf{m}_{i} = \alpha_{H}\mathbf{H_{0}}(\mathbf{r}_{i})& +\alpha_{H}k^{2} \left(\underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{m}_{j} +c_0\,\nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{p}_{j}\right), \end{align} \end{subequations} where $\mathbf{p}_i$ ($\mathbf{m}_i$) is the electric (magnetic) dipole moment of the $i^{\mathrm{th}}$ particle, $\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) $ is the free space dyadic Greens function between the $i^{\mathrm{th}}$ and $j^{\mathrm{th}}$ dipole, $\alpha_{E}$ ($\alpha_{H}$) is the electric (magnetic) polarizability of a particle, $c_0$ is the speed of light and $k$ is the free-space wavenumber. In Fig.~\ref{fig:CSTvsCEMDA}a, we show that this dipole model is accurate in modeling full-wave simulations of the silicon nanosphere quadrumers performed using CST Microwave Studio, confirming that, in this spectral range, the optical response of such nanoparticle quadrumers is dominated by dipole interactions among the individual particles. { This is the case because higher order coupling in such systems will occur only with smaller interparticle separations~\cite{Albella2013} (see also the Supporting Information).} Fig.~\ref{fig:CSTvsCEMDA}b shows the electric field distribution at the collective magnetic resonance of the cluster, confirming that the response of the silicon quadrumer includes a collective circulation of electric field around the particles. However, we can also see the additional dynamic produced by dielectric nanoparticles: the magnetic response induced by circulating transverse electric dipoles is accompanied by a magnetic response from each of the individual nanoparticles. This is the unique effect we are interested in. \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.9\textwidth]{CSTvsCEMDA2.pdf}} \caption{ (a) Extinction calculated using both the CEMDA and CST Microwave Studio for a quadrumer composed of $150\,\mathrm{nm}$ (diameter) silicon nanospheres, each separated by $50\,\mathrm{nm}$. (b) Associated near-field profile of the induced electric field at the silicon quadrumer's Fano resonance $(\lambda =570 \,\mathrm{nm})$; the superimposed cyan arrows indicate the direction of circulation for the electric field both external and internal to the individual spheres. Electric field amplitude is normalized to the amplitude of the incident { plane wave, whose propagation direction ($\mathbf{k}$) and polarization ($\mathbf{E}$) are both parallel to the plane of the quadrumer}. } \label{fig:CSTvsCEMDA} \end{figure} The magnetic response in an individual silicon nanosphere is an internal circulation of polarization current, which can be seen in Fig.~\ref{fig:CSTvsCEMDA}b. In the CEMDA we choose to homogenize this circulating current distribution into a discrete source of magnetization current; the magnetic dipole. This implies that the key to strong interaction between individual and collective magnetic responses resides in the strong coupling between induced electric and magnetic dipoles in each inclusion. To describe this coupling, we can consider the eigenmodes of the quadrumer system. As we will show, the eigenmodes of the full, electromagnetic, system can be constructed from the eigenmodes of the decoupled electric and magnetic dipole systems. Therefore, we are going to break apart our dipole system of Eq.~\ref{eq:dipole equations} into two decoupled equations: one equation for the electric dipoles and one equation for the magnetic dipoles \begin{subequations} \label{eq:decoupled equations} \begin{align} \mathbf{p}_{i} = \alpha_{E}\epsilon_{0}\mathbf{E_{0}}(\mathbf{r}_{i})& +\alpha_{E}k^{2} \underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{p}_{j} \;,\\ \mathbf{m}_{i} = \alpha_{H}\mathbf{H_{0}}(\mathbf{r}_{i})& +\alpha_{H}k^{2} \underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{m}_{j} \;. \end{align} \end{subequations} We can then define two sets of eigenmodes from these equations: one for the electric dipoles and one for the magnetic dipoles. Using state notation, we can refer to the electric dipole eigenmodes as $\left | \mathbf{e} \right \rangle$ and the magnetic dipole eigenmodes as $\left | \mathbf{h} \right \rangle$. The eigenmode approach in the dipole approximation offers a huge simplification when combined with symmetry, because it allows us to determine certain eigenmodes without calculation. The key is that, for both electric and magnetic dipole systems, the dipole approximation restricts the number of eigenmodes for each equation to twelve (three dimensions by four dipoles). Each of these eigenmodes can only transform according to a single irreducible representation of the quadrumer's symmetry group,\cite{Dresselhaus2008} referred to as $\mathrm{D_{4h}}$. Further details regarding the $\mathrm{D_{4h}}$ symmetry group's irreducible representations and the implications for eigenmodes are provided in the { Supporting Information}. For the analysis here, it suffices to say that there are only eight irreducible representations in $\mathrm{D_{4h}}$ for twelve eigenmodes, and therefore the eigenspace associated with certain irreducible representations must be one-dimensional. Moreover, given that any vector in a one-dimensional space is an eigenvector by default, we are able to derive a number of eigenmodes by simply finding dipole moment profiles that transform according to certain irreducible representations. Eight such dipole moment profiles are shown in Fig.~\ref{fig:eigenmodes}. Each of these dipole moment profiles is the sole basis vector for a single irreducible representation and is therefore an eigenmode of the electric or magnetic dipole equations in Eq.~\ref{eq:decoupled equations}, irrespective of wavelength or the choice of material, size, or any parameter which conserves the symmetry of the quadrumer. It is worth noting that this same procedure for finding eigenmodes is applicable to other symmetries when using the dipole approximation, particularly the $\mathrm{D}_{n\mathrm{h}}$ symmetry groups for small values of $n$. \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.75\textwidth]{eigenmodes2.pdf}} \caption{Basis vectors for the optical response of the quadrumer's electric and magnetic dipoles, named according to their associated irreducible representation. Due to finite dimensions and symmetry constraints, each basis vector shown here is an eigenmode of the electric or magnetic dipole equation (Eq.~\ref{eq:decoupled equations}a or Eq.~\ref{eq:decoupled equations}b). } \label{fig:eigenmodes} \end{figure} We need to now consider the interaction between electric and magnetic dipole systems. An eigenmode of either electric or magnetic dipoles can be substituted into Eq.~\ref{eq:dipole equations} to determine the resulting state of magnetic or electric dipoles ($ | \mathbf{m_{(e)}} \rangle$ or $ | \mathbf{p_{(h)}} \rangle$, respectively) that it will induce due to magneto-electric coupling: \begin{subequations}\label{eq:cross} \begin{align} \mathbf{m_{(e)}}_{i} =\alpha_{H} k^{2} {c_0} \underset{j\neq i}{{\sum}} \nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{e}_{j}\;,\\ \mathbf{p_{(h)}}_{i} =\frac{-\alpha_{E} k^{2}} {c_0} \underset{j\neq i}{{\sum}} \nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{h}_{j}\;. \end{align} \end{subequations} However, it has been shown that the operation describing the coupling between states of electric and magnetic dipole moments, must commute with the geometry's symmetry operations.~\cite{HopkinsLiu2013} Subsequently the $\left | \mathbf{e}\right \rangle$ and $ | \mathbf{m_{(e)}} \rangle$ dipole moments must both transform according to the same irreducible representation, and similarly for the $\left | \mathbf{h}\right \rangle$ and $ | \mathbf{p_{(h)}} \rangle$ dipole moments. Therefore, if we consider the eigenmodes of electric and magnetic dipoles in Fig.~\ref{fig:eigenmodes}, the requirement for symmetry conservation specifies which electric dipole eigenmode can couple into which (if any) magnetic dipole eigenmode, and {\it vice versa}. As a result, the $\mathrm{A_{2g}}$ or $\mathrm{B_{1g}}$ eigenmodes ({\it cf.} Fig.~\ref{fig:eigenmodes}) are only able to magneto-electrically couple into the other $\mathrm{A_{2g}}$ or $\mathrm{B_{1g}}$ eigenmode, whereas the $\mathrm{A_{1g}}$ and $\mathrm{B_{2g}}$ eigenmodes are not able to couple into any eigenmodes due to a symmetry mismatch. This conclusion allows us to take our analysis of the decoupled electric and magnetic dipole equations, and use it to determine the full, {\it electromagnetic}, eigenmodes of the real system. Indeed, we would not necessarily expect an eigenmode of this complex scattering system to be a current distribution described by purely electric or magnetic dipoles; we would expect it to be a combination of both. If we now restrict ourselves to considering only the $\mathrm{A_{2g}}$ (magnetic-like) eigenmodes of an all-dielectric quadrumer, the two eigenmodes from the decoupled electric and magnetic dipole equations ({\it cf.} Fig.~\ref{fig:eigenmodes}) form basis vectors for the $\mathrm{A_{2g}}$ eigenmodes, $| \mathbf{v}_x \rangle$, of the electromagnetic system. \begin{align} | \mathbf{v}_x \rangle = a_x \left ( \!\! \begin{array}{c} | \mathbf{e} \rangle \\ 0 \end{array} \!\!\right) + b_x \left (\!\! \begin{array}{c} 0 \\ | \mathbf{h} \rangle \end{array} \!\! \right)\,,\label{eq:EMeigenmode} \end{align} where $a_x$ and $b_x$ are complex scalars. Notably, we should expect to have two distinct electromagnetic eigenmodes here because there are two distinct basis vectors and, subsequently, a two-dimensional eigenspace for the quadrumer's response. To derive expressions for $a_x$ and $b_x$, we can write the electromagnetic eigenvalue equation for Eq.~\ref{eq:dipole equations}: \begin{subequations} \label{eq:eigenmode equations} \begin{align} a_x \mathbf{e}_{i} &= a_x \lambda_x \alpha_{E}\epsilon_{0}\mathbf{e}_i+\alpha_{E}k^{2} \left(\underset{j\neq i}{{\sum}}a_x\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{e}_{j} - \frac{b_x}{{c_0}} \,\nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{h}_{j}\right) ,\\ b_x \mathbf{h}_{i} &= b_x \lambda_x \alpha_{H}\mathbf{h}_{i} +\alpha_{H}k^{2} \left(\underset{j\neq i}{{\sum}}b_x\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{h}_{j} +a_x {c_0}\, \nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{e}_{j}\right), \end{align} \end{subequations} where $\lambda_x$ is the eigenvalue of the electromagnetic eigenmode. However, Eqs.~\ref{eq:eigenmode equations}a and \ref{eq:eigenmode equations}b are not independent\footnote{ Further details of this relation are provided in the { Supporting Information}. } because one can be obtained from the other using duality transformations.~\cite{Jackson} In other words, a solution for $a_x$ and $b_x$, that satisfies {\it either} Eq.~\ref{eq:eigenmode equations}a or Eq.~\ref{eq:eigenmode equations}b, must also satisfy the complete Eq.~\ref{eq:eigenmode equations} and describe an electromagnetic eigenmode of Eq.~\ref{eq:EMeigenmode}. So, if the basis vectors $\left | \mathbf{e}\right \rangle$ and $\left | \mathbf{h} \right \rangle $ are normalized to a magnitude of one, we can project Eq.~\ref{eq:eigenmode equations}a and Eq.~\ref{eq:eigenmode equations}b onto $ \mathbf{e}_{i}$ and $ \mathbf{h}_{i}$ (respectively) and sum over all $i$, to obtain the two solutions for Eq.~\ref{eq:eigenmode equations}: \begin{align} \frac{a_1}{b_1} &= \frac{ \underset{i}{\sum} \, \mathbf{e}_i^{*}\cdot \mathbf{p_{(h)}}_{i} } {\alpha_{\scriptscriptstyle E} \epsilon_0 (\lambda_{ \mathbf{e} } - \lambda_1)} \;, \label{eq:result1} \\[1 ex] \frac{b_2}{a_2} &= \frac{\underset{i}{\sum} \mathbf{h}_i^{*}\cdot \mathbf{m_{(e)}}_{i} } {\alpha_{\scriptscriptstyle H}(\lambda_{ \mathbf{h}} - \lambda_2)} \;, \label{eq:result2} \end{align} where $\lambda_{ \mathbf{e} }$ and $\lambda_{ \mathbf{h}} $ are the eigenvalues of $| \mathbf{e} \rangle$ and $| \mathbf{h} \rangle$ in the decoupled electric and magnetic dipole equations in Eq.~\ref{eq:decoupled equations}: \begin{subequations} \label{eq:decoupled eigenequation} \begin{align} \alpha_{E}\epsilon_{0}\lambda_{ \mathbf{e}} & = 1 -\alpha_{E}k^{2} \underset{i,\,j\neq i}{{\sum}}\mathbf{e}_i^{*}\cdot \hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{e}_{j} \;,\\ \alpha_{H}\lambda_{ \mathbf{h}} & = 1 -\alpha_{H}k^{2} \underset{i,\,j\neq i}{{\sum}}\mathbf{h}_i^{*}\cdot \hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{h}_{j} \;. \end{align} \end{subequations} The two ratios in Eq.~\ref{eq:result1} and Eq.~\ref{eq:result2} each describe a distinct eigenmode for the electromagnetic system according to Eq.~\ref{eq:EMeigenmode}. It is worth noting that, if either numerator or denominator go to zero in Eq.~\ref{eq:result1} or Eq.~\ref{eq:result2}, the starting basis vectors of the decoupled electric or magnetic dipole equations were already eigenmodes of the full system. In regard to the $\mathrm{A_{2g}}$ (magnetic-like) eigenmodes of an all-dielectric quadrumer, our derivations show that, when both eigenmodes are far from resonance and magneto-electric coupling can be neglected, one eigenmode will be purely composed of electric dipoles and the other will be purely composed of magnetic dipoles. However, as they approach resonance, magneto-electric coupling cannot be neglected and the two ratios of $a_x$ and $b_x$ will need to be calculated to determine their form as electromagnetic eigenmodes (see Fig.~\ref{fig:interference}c). In such a situation, the two eigenmodes, $| \mathbf{v}_1\rangle$ and $| \mathbf{v}_2\rangle$, are explicitly nonorthogonal to each other with the overlap defined by: \begin{align} \langle \mathbf{v}_1|\mathbf{v}_2 \rangle = a_1 ^* a_2 + b_1 ^* b_2\,. \label{eq:overlap} \end{align} We can therefore expect interference between these eigenmodes as they approach resonance.~\cite{HopkinsPoddubny2013} \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.8\textwidth]{{decomposition2.pdf}}} \caption{Electromagnetic eigenmode decomposition of the silicon quadrumer, calculated using the CEMDA, showing (a) the contribution of the $\mathrm{A_{2g}}$ (magnetic-like) eigenmodes to the overall extinction. In (b) we decompose the $\mathrm{A_{2g}}$ response in terms of its two eigenmodes, showing the creation of a Fano resonance from interference between them. To support this we also show (c) the real components of the dipole moment profiles for each $\mathrm{A_{2g}}$ eigenmode, calculated at the wavelength of the Fano feature. } \label{fig:interference} \end{figure} Indeed, as shown in Fig.~\ref{fig:interference}, the Fano resonance feature in the silicon quadrumer of Fig.~\ref{fig:CSTvsCEMDA} is produced entirely by the interference between these two eigenmodes. To serve as a broader qualification of this magnetic Fano resonance feature, we also investigate its parameter dependence. In Fig.~\ref{fig:sweep}a and \ref{fig:sweep}b, we vary the size of the gap between nanoparticles and the size of each individual nanoparticle, respectively. { It can be seen that the Fano resonance is dependent on a number of these coupling parameters, but in a way that is different to electric Fano resonances in plasmonic oligomers. Perhaps most notably, the magnetic Fano resonance is significantly shifted with small changes in size of the constituent nanoparticles, and closely spaced nanoparticles are favorable for electric Fano resonances in plasmonic oligomers, but can destroy the magnetic Fano resonance here.~\cite{Hentschel2010, RahmaniLukyanchuk2013} More specifically, the magnetic Fano resonance feature is heavily dependent on the spacing between neighboring nanoparticles; increasing or decreasing the spacing can quickly diminish the Fano feature. } On the other hand, varying the size of the particles while holding the gap between particles constant is able to conserve the Fano resonance and shift it spectrally (shown in the shaded region of Fig.~\ref{fig:sweep}b). This behavior can be expected because the action of increasing the size of the particles while holding the gap constant is very similar to uniformly scaling Maxwell's equations, and silicon permittivity has relatively minor dispersion in this spectral range.~\cite{Palik} Otherwise, the final parameter we consider in Fig.~\ref{fig:sweep}c, is the angle of incidence to address the practical limitations for in-plane excitation of nanoscale optical structures. It can be seen that the Fano resonance will persist up to roughly 45\degree~incidence, beyond which the magnetic response is dominated by the normal-incidence responses. \begin{figure}[!t] \centering \centerline{\includegraphics[width=0.9\textwidth]{{sweep.pdf}}} \caption{Simulation results, from CST Microwave Studio, of the extinction caused by the silicon quadrumer in Fig.~\ref{fig:CSTvsCEMDA}, when changing (a) the size of the gap between nanoparticles, (b) the size of the individual nanoparticles and (c) the angle of incidence of the applied plane wave, with respect to the quadrumer plane (maintaining $s$-polarization).} \label{fig:sweep} \end{figure} \newpage {\it Experimental verification.} To verify and validate the theoretical arguments presented above, we look for the existence of magnetic Fano resonances in a high index cluster experimentally. In this regard, one practical option is to mimic the scattering properties of silicon nanoparticles using MgO-$\mathrm{TiO_2}$ ceramic spheres characterized by dielectric constant of 16 and dielectric loss factor of $(1.12  -  1.17)\times10^{-4}$, measured at {$9 - 12$ GHz}. These ceramic spheres in the microwave range therefore have very similar properties to silicon nanospheres in the optical range and they are subsequently a useful macroscopic platform on which to prototype silicon nanostructures. Here, they allow us to perform a `proof of concept' investigation into the properties of an isolated quadrumer with much more signal than would be expected from a single silicon nanosphere quadrumer. Indeed, such spheres have been used previously to predict the behavior of silicon nanoantennas.{~\cite{Filonov2012, Geffrin2012, Filonov2014}} \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.9\textwidth]{{experiment.pdf}}} \caption{The (a) CST simulation results and (b) experimental measurements of extinction for a quadrumer made of four MgO-$\mathrm{TiO_2}$ ceramic spheres. We also show the extinction from a single MgO-$\mathrm{TiO_2}$ sphere for reference. In both simulation and experiment we observe the existence of a sharp Fano resonance occurring at 5.4 GHz. } \label{fig:experiment} \end{figure} The MgO-$\mathrm{TiO_2}$ quadrumer consists of four dielectric spheres with diameter $d=15\mathrm{mm}$, and the size of the gap between the particles is $s=5\, \mathrm{ mm}$. The experimentally measured, and numerically calculated, total scattering of the quadrumer structure are shown in Fig.~\ref{fig:experiment}. It can clearly be seen that a magnetic Fano resonance is produced at 5.4 GHz, in both simulation and experiment. This is the first example of a magnetic-magnetic Fano resonance in a single symmetric metamolecule. Notably, this Fano resonance occurs in a spectral range where the single particle is not at resonance, which demonstrates its collective nature. Indeed, it appears near the intersection of the single particle's electric and magnetic scattering contributions, reflecting that the overlap of eigenmodes (Eq.~\ref{eq:overlap}) is dependent on both electric and magnetic dipole polarizabilities. \section{Conclusions} We have presented a comprehensive study of the interplay between the collective optically-induced magnetic responses of all-dielectric quadrumers and the individual magnetic responses of their constituent dielectric nanoparticles. We have been able to establish the theoretical basis behind the interaction between collective and individual magnetic responses in all-dielectric structures, providing a quantitative prediction of the interference between the quadrumer's magnetic responses leading to magnetic-magnetic Fano resonance features. We have also been able to experimentally observe the existence of a sharp magnetic-magnetic Fano resonance in a dielectric quadrumer. Such Fano resonance features demonstrate the unique potential that suitably designed dielectric nanoclusters have for scattering engineering at the nanoscale, opening exciting and unexplored opportunities for dielectric nanophotonics. \section{Methods} To fasten together the MgO-$\mathrm{TiO_2}$ ceramic spheres for the experiment, we used a custom holder made of a styrofoam material with dielectric permittivity of 1 (in the microwave frequency range). To approximate plane wave excitation, we employed a rectangular horn antenna (TRIM $0.75 - 18$ GHz ; DR) connected to the transmitting port of a vector network analyzer (Agilent E8362C). The quadrumer was then located in the far-field of the antenna, at a distance of approximately 2.5 m, and a second horn antenna (TRIM $0.75 - 18$ GHz ; DR) was used as a receiver to observe the transmission through the quadrumer. The extinction measurement shown in Fig.~\ref{fig:experiment}b was then obtained as the difference between the measured transmission and unity transmission ({\it i.e.}, with no quadrumer). For the theory, CST simulations were performed assuming plane wave excitation on a quadrumer located in free space. The CEMDA simulations, seen in Fig.~\ref{fig:CSTvsCEMDA}a, used electric and magnetic dipole polarizabilities that were derived from the scattering coefficients of Mie theory.~\cite{Mie1908} In all simulations, silicon permittivity data was taken from Palik's Handbook~\cite{Palik} and the MgO-$\mathrm{TiO_2}$ spheres were assumed be dispersionless with a dielectric constant of 16 and dielectric loss factor of $(1.12  -  1.17)\times10^{-4}$. { \section{Acknowledgements} This work was supported by the Australian Research Council. F.M. and A.A. have been supported through the U.S. Air Force Office of Scientific Research with grant No. FA9550-13-1-0204, P02. The measurements were supported by the Government of the Russian Federation (grant 074-U01), the Ministry of Education and Science of the Russian Federation, Russian Foundation for Basic Research, Dynasty Foundation (Russia). } \bibliographystyle{apsrev4-1} \section{Results and Discussion} { Consider the optical response of a symmetric quadrumer when it is excited by a plane wave, whose propagation direction and electric field polarization lie in the plane of the quadrumer ($s$-polarization), as shown in Fig.~\ref{fig:schematic}a. } \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.5\textwidth]{schematic.pdf}} \caption{(a) Schematic of the problem: a symmetric nanosphere quadrumer is excited by an in-plane $s$-polarized plane wave {(both propagation and polarization vectors lie in the plane of the quadrumer)}. (b) The combined magnetic response: a collective optically-induced magnetic resonance of the whole structure is accompanied by the magnetic responses of the individual particles. } \label{fig:schematic} \end{figure} The conventional magnetic response to such an excitation is supported by a collective circulation of electric displacement current around all four nanoparticles. However, if the nanoparticles are made of a high-index dielectric material such as silicon, the individual nanoparticles are also able to couple directly with the applied magnetic field, sustaining internal circulating polarization currents. In this sense, a silicon quadrumer can produce magnetic responses to both electric {\it and} magnetic components of the applied plane wave, as is depicted in Fig.~\ref{fig:schematic}b. { In this figure, the red arrows denote an electric dipole, and the blue circular arrows denote the circulation of displacement current that supports an out-of-page magnetic dipole. Importantly, } an interplay may be induced between the two magnetic responses through the locally-enhanced, $z$-polarized magnetic field sustained by the collective circulation of displacement current, which establishes a coupling channel to the magnetic responses of the individual nanoparticles. We begin by understanding this magnetic interplay, for which we will model the interactions between nanoparticles using the {\it Coupled Electric and Magnetic Dipole Approximation}~\cite{Mulholland1994} (CEMDA). The mathematical description of this dipole model in free space is based on the following equations: \begin{subequations} \label{eq:dipole equations} \begin{align} \mathbf{p}_{i} = \alpha_{E}\epsilon_{0}\mathbf{E_{0}}(\mathbf{r}_{i})& +\alpha_{E}k^{2} \left(\underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{p}_{j} -\frac{1}{c_0}\,\nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{m}_{j}\right) ,\\ \mathbf{m}_{i} = \alpha_{H}\mathbf{H_{0}}(\mathbf{r}_{i})& +\alpha_{H}k^{2} \left(\underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{m}_{j} +c_0\,\nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{p}_{j}\right), \end{align} \end{subequations} where $\mathbf{p}_i$ ($\mathbf{m}_i$) is the electric (magnetic) dipole moment of the $i^{\mathrm{th}}$ particle, $\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) $ is the free space dyadic Greens function between the $i^{\mathrm{th}}$ and $j^{\mathrm{th}}$ dipole, $\alpha_{E}$ ($\alpha_{H}$) is the electric (magnetic) polarizability of a particle, $c_0$ is the speed of light and $k$ is the free-space wavenumber. In Fig.~\ref{fig:CSTvsCEMDA}a, we show that this dipole model is accurate in modeling full-wave simulations of the silicon nanosphere quadrumers performed using CST Microwave Studio, confirming that, in this spectral range, the optical response of such nanoparticle quadrumers is dominated by dipole interactions among the individual particles. { This is the case because higher order coupling in such systems will occur only with smaller interparticle separations~\cite{Albella2013} (see also the Supporting Information).} Fig.~\ref{fig:CSTvsCEMDA}b shows the electric field distribution at the collective magnetic resonance of the cluster, confirming that the response of the silicon quadrumer includes a collective circulation of electric field around the particles. However, we can also see the additional dynamic produced by dielectric nanoparticles: the magnetic response induced by circulating transverse electric dipoles is accompanied by a magnetic response from each of the individual nanoparticles. This is the unique effect we are interested in. \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.9\textwidth]{CSTvsCEMDA2.pdf}} \caption{ (a) Extinction calculated using both the CEMDA and CST Microwave Studio for a quadrumer composed of $150\,\mathrm{nm}$ (diameter) silicon nanospheres, each separated by $50\,\mathrm{nm}$. (b) Associated near-field profile of the induced electric field at the silicon quadrumer's Fano resonance $(\lambda =570 \,\mathrm{nm})$; the superimposed cyan arrows indicate the direction of circulation for the electric field both external and internal to the individual spheres. Electric field amplitude is normalized to the amplitude of the incident { plane wave, whose propagation direction ($\mathbf{k}$) and polarization ($\mathbf{E}$) are both parallel to the plane of the quadrumer}. } \label{fig:CSTvsCEMDA} \end{figure} The magnetic response in an individual silicon nanosphere is an internal circulation of polarization current, which can be seen in Fig.~\ref{fig:CSTvsCEMDA}b. In the CEMDA we choose to homogenize this circulating current distribution into a discrete source of magnetization current; the magnetic dipole. This implies that the key to strong interaction between individual and collective magnetic responses resides in the strong coupling between induced electric and magnetic dipoles in each inclusion. To describe this coupling, we can consider the eigenmodes of the quadrumer system. As we will show, the eigenmodes of the full, electromagnetic, system can be constructed from the eigenmodes of the decoupled electric and magnetic dipole systems. Therefore, we are going to break apart our dipole system of Eq.~\ref{eq:dipole equations} into two decoupled equations: one equation for the electric dipoles and one equation for the magnetic dipoles \begin{subequations} \label{eq:decoupled equations} \begin{align} \mathbf{p}_{i} = \alpha_{E}\epsilon_{0}\mathbf{E_{0}}(\mathbf{r}_{i})& +\alpha_{E}k^{2} \underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{p}_{j} \;,\\ \mathbf{m}_{i} = \alpha_{H}\mathbf{H_{0}}(\mathbf{r}_{i})& +\alpha_{H}k^{2} \underset{j\neq i}{{\sum}}\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{m}_{j} \;. \end{align} \end{subequations} We can then define two sets of eigenmodes from these equations: one for the electric dipoles and one for the magnetic dipoles. Using state notation, we can refer to the electric dipole eigenmodes as $\left | \mathbf{e} \right \rangle$ and the magnetic dipole eigenmodes as $\left | \mathbf{h} \right \rangle$. The eigenmode approach in the dipole approximation offers a huge simplification when combined with symmetry, because it allows us to determine certain eigenmodes without calculation. The key is that, for both electric and magnetic dipole systems, the dipole approximation restricts the number of eigenmodes for each equation to twelve (three dimensions by four dipoles). Each of these eigenmodes can only transform according to a single irreducible representation of the quadrumer's symmetry group,\cite{Dresselhaus2008} referred to as $\mathrm{D_{4h}}$. Further details regarding the $\mathrm{D_{4h}}$ symmetry group's irreducible representations and the implications for eigenmodes are provided in the { Supporting Information}. For the analysis here, it suffices to say that there are only eight irreducible representations in $\mathrm{D_{4h}}$ for twelve eigenmodes, and therefore the eigenspace associated with certain irreducible representations must be one-dimensional. Moreover, given that any vector in a one-dimensional space is an eigenvector by default, we are able to derive a number of eigenmodes by simply finding dipole moment profiles that transform according to certain irreducible representations. Eight such dipole moment profiles are shown in Fig.~\ref{fig:eigenmodes}. Each of these dipole moment profiles is the sole basis vector for a single irreducible representation and is therefore an eigenmode of the electric or magnetic dipole equations in Eq.~\ref{eq:decoupled equations}, irrespective of wavelength or the choice of material, size, or any parameter which conserves the symmetry of the quadrumer. It is worth noting that this same procedure for finding eigenmodes is applicable to other symmetries when using the dipole approximation, particularly the $\mathrm{D}_{n\mathrm{h}}$ symmetry groups for small values of $n$. \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.75\textwidth]{eigenmodes2.pdf}} \caption{Basis vectors for the optical response of the quadrumer's electric and magnetic dipoles, named according to their associated irreducible representation. Due to finite dimensions and symmetry constraints, each basis vector shown here is an eigenmode of the electric or magnetic dipole equation (Eq.~\ref{eq:decoupled equations}a or Eq.~\ref{eq:decoupled equations}b). } \label{fig:eigenmodes} \end{figure} We need to now consider the interaction between electric and magnetic dipole systems. An eigenmode of either electric or magnetic dipoles can be substituted into Eq.~\ref{eq:dipole equations} to determine the resulting state of magnetic or electric dipoles ($ | \mathbf{m_{(e)}} \rangle$ or $ | \mathbf{p_{(h)}} \rangle$, respectively) that it will induce due to magneto-electric coupling: \begin{subequations}\label{eq:cross} \begin{align} \mathbf{m_{(e)}}_{i} =\alpha_{H} k^{2} {c_0} \underset{j\neq i}{{\sum}} \nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{e}_{j}\;,\\ \mathbf{p_{(h)}}_{i} =\frac{-\alpha_{E} k^{2}} {c_0} \underset{j\neq i}{{\sum}} \nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \cdot \mathbf{h}_{j}\;. \end{align} \end{subequations} However, it has been shown that the operation describing the coupling between states of electric and magnetic dipole moments, must commute with the geometry's symmetry operations.~\cite{HopkinsLiu2013} Subsequently the $\left | \mathbf{e}\right \rangle$ and $ | \mathbf{m_{(e)}} \rangle$ dipole moments must both transform according to the same irreducible representation, and similarly for the $\left | \mathbf{h}\right \rangle$ and $ | \mathbf{p_{(h)}} \rangle$ dipole moments. Therefore, if we consider the eigenmodes of electric and magnetic dipoles in Fig.~\ref{fig:eigenmodes}, the requirement for symmetry conservation specifies which electric dipole eigenmode can couple into which (if any) magnetic dipole eigenmode, and {\it vice versa}. As a result, the $\mathrm{A_{2g}}$ or $\mathrm{B_{1g}}$ eigenmodes ({\it cf.} Fig.~\ref{fig:eigenmodes}) are only able to magneto-electrically couple into the other $\mathrm{A_{2g}}$ or $\mathrm{B_{1g}}$ eigenmode, whereas the $\mathrm{A_{1g}}$ and $\mathrm{B_{2g}}$ eigenmodes are not able to couple into any eigenmodes due to a symmetry mismatch. This conclusion allows us to take our analysis of the decoupled electric and magnetic dipole equations, and use it to determine the full, {\it electromagnetic}, eigenmodes of the real system. Indeed, we would not necessarily expect an eigenmode of this complex scattering system to be a current distribution described by purely electric or magnetic dipoles; we would expect it to be a combination of both. If we now restrict ourselves to considering only the $\mathrm{A_{2g}}$ (magnetic-like) eigenmodes of an all-dielectric quadrumer, the two eigenmodes from the decoupled electric and magnetic dipole equations ({\it cf.} Fig.~\ref{fig:eigenmodes}) form basis vectors for the $\mathrm{A_{2g}}$ eigenmodes, $| \mathbf{v}_x \rangle$, of the electromagnetic system. \begin{align} | \mathbf{v}_x \rangle = a_x \left ( \!\! \begin{array}{c} | \mathbf{e} \rangle \\ 0 \end{array} \!\!\right) + b_x \left (\!\! \begin{array}{c} 0 \\ | \mathbf{h} \rangle \end{array} \!\! \right)\,,\label{eq:EMeigenmode} \end{align} where $a_x$ and $b_x$ are complex scalars. Notably, we should expect to have two distinct electromagnetic eigenmodes here because there are two distinct basis vectors and, subsequently, a two-dimensional eigenspace for the quadrumer's response. To derive expressions for $a_x$ and $b_x$, we can write the electromagnetic eigenvalue equation for Eq.~\ref{eq:dipole equations}: \begin{subequations} \label{eq:eigenmode equations} \begin{align} a_x \mathbf{e}_{i} &= a_x \lambda_x \alpha_{E}\epsilon_{0}\mathbf{e}_i+\alpha_{E}k^{2} \left(\underset{j\neq i}{{\sum}}a_x\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{e}_{j} - \frac{b_x}{{c_0}} \,\nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{h}_{j}\right) ,\\ b_x \mathbf{h}_{i} &= b_x \lambda_x \alpha_{H}\mathbf{h}_{i} +\alpha_{H}k^{2} \left(\underset{j\neq i}{{\sum}}b_x\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{h}_{j} +a_x {c_0}\, \nabla\times\hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{e}_{j}\right), \end{align} \end{subequations} where $\lambda_x$ is the eigenvalue of the electromagnetic eigenmode. However, Eqs.~\ref{eq:eigenmode equations}a and \ref{eq:eigenmode equations}b are not independent\footnote{ Further details of this relation are provided in the { Supporting Information}. } because one can be obtained from the other using duality transformations.~\cite{Jackson} In other words, a solution for $a_x$ and $b_x$, that satisfies {\it either} Eq.~\ref{eq:eigenmode equations}a or Eq.~\ref{eq:eigenmode equations}b, must also satisfy the complete Eq.~\ref{eq:eigenmode equations} and describe an electromagnetic eigenmode of Eq.~\ref{eq:EMeigenmode}. So, if the basis vectors $\left | \mathbf{e}\right \rangle$ and $\left | \mathbf{h} \right \rangle $ are normalized to a magnitude of one, we can project Eq.~\ref{eq:eigenmode equations}a and Eq.~\ref{eq:eigenmode equations}b onto $ \mathbf{e}_{i}$ and $ \mathbf{h}_{i}$ (respectively) and sum over all $i$, to obtain the two solutions for Eq.~\ref{eq:eigenmode equations}: \begin{align} \frac{a_1}{b_1} &= \frac{ \underset{i}{\sum} \, \mathbf{e}_i^{*}\cdot \mathbf{p_{(h)}}_{i} } {\alpha_{\scriptscriptstyle E} \epsilon_0 (\lambda_{ \mathbf{e} } - \lambda_1)} \;, \label{eq:result1} \\[1 ex] \frac{b_2}{a_2} &= \frac{\underset{i}{\sum} \mathbf{h}_i^{*}\cdot \mathbf{m_{(e)}}_{i} } {\alpha_{\scriptscriptstyle H}(\lambda_{ \mathbf{h}} - \lambda_2)} \;, \label{eq:result2} \end{align} where $\lambda_{ \mathbf{e} }$ and $\lambda_{ \mathbf{h}} $ are the eigenvalues of $| \mathbf{e} \rangle$ and $| \mathbf{h} \rangle$ in the decoupled electric and magnetic dipole equations in Eq.~\ref{eq:decoupled equations}: \begin{subequations} \label{eq:decoupled eigenequation} \begin{align} \alpha_{E}\epsilon_{0}\lambda_{ \mathbf{e}} & = 1 -\alpha_{E}k^{2} \underset{i,\,j\neq i}{{\sum}}\mathbf{e}_i^{*}\cdot \hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{e}_{j} \;,\\ \alpha_{H}\lambda_{ \mathbf{h}} & = 1 -\alpha_{H}k^{2} \underset{i,\,j\neq i}{{\sum}}\mathbf{h}_i^{*}\cdot \hat{G}_{0}(\mathbf{r}_{i},\mathbf{r}_{j}) \mathbf{h}_{j} \;. \end{align} \end{subequations} The two ratios in Eq.~\ref{eq:result1} and Eq.~\ref{eq:result2} each describe a distinct eigenmode for the electromagnetic system according to Eq.~\ref{eq:EMeigenmode}. It is worth noting that, if either numerator or denominator go to zero in Eq.~\ref{eq:result1} or Eq.~\ref{eq:result2}, the starting basis vectors of the decoupled electric or magnetic dipole equations were already eigenmodes of the full system. In regard to the $\mathrm{A_{2g}}$ (magnetic-like) eigenmodes of an all-dielectric quadrumer, our derivations show that, when both eigenmodes are far from resonance and magneto-electric coupling can be neglected, one eigenmode will be purely composed of electric dipoles and the other will be purely composed of magnetic dipoles. However, as they approach resonance, magneto-electric coupling cannot be neglected and the two ratios of $a_x$ and $b_x$ will need to be calculated to determine their form as electromagnetic eigenmodes (see Fig.~\ref{fig:interference}c). In such a situation, the two eigenmodes, $| \mathbf{v}_1\rangle$ and $| \mathbf{v}_2\rangle$, are explicitly nonorthogonal to each other with the overlap defined by: \begin{align} \langle \mathbf{v}_1|\mathbf{v}_2 \rangle = a_1 ^* a_2 + b_1 ^* b_2\,. \label{eq:overlap} \end{align} We can therefore expect interference between these eigenmodes as they approach resonance.~\cite{HopkinsPoddubny2013} \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.8\textwidth]{{decomposition2.pdf}}} \caption{Electromagnetic eigenmode decomposition of the silicon quadrumer, calculated using the CEMDA, showing (a) the contribution of the $\mathrm{A_{2g}}$ (magnetic-like) eigenmodes to the overall extinction. In (b) we decompose the $\mathrm{A_{2g}}$ response in terms of its two eigenmodes, showing the creation of a Fano resonance from interference between them. To support this we also show (c) the real components of the dipole moment profiles for each $\mathrm{A_{2g}}$ eigenmode, calculated at the wavelength of the Fano feature. } \label{fig:interference} \end{figure} Indeed, as shown in Fig.~\ref{fig:interference}, the Fano resonance feature in the silicon quadrumer of Fig.~\ref{fig:CSTvsCEMDA} is produced entirely by the interference between these two eigenmodes. To serve as a broader qualification of this magnetic Fano resonance feature, we also investigate its parameter dependence. In Fig.~\ref{fig:sweep}a and \ref{fig:sweep}b, we vary the size of the gap between nanoparticles and the size of each individual nanoparticle, respectively. { It can be seen that the Fano resonance is dependent on a number of these coupling parameters, but in a way that is different to electric Fano resonances in plasmonic oligomers. Perhaps most notably, the magnetic Fano resonance is significantly shifted with small changes in size of the constituent nanoparticles, and closely spaced nanoparticles are favorable for electric Fano resonances in plasmonic oligomers, but can destroy the magnetic Fano resonance here.~\cite{Hentschel2010, RahmaniLukyanchuk2013} More specifically, the magnetic Fano resonance feature is heavily dependent on the spacing between neighboring nanoparticles; increasing or decreasing the spacing can quickly diminish the Fano feature. } On the other hand, varying the size of the particles while holding the gap between particles constant is able to conserve the Fano resonance and shift it spectrally (shown in the shaded region of Fig.~\ref{fig:sweep}b). This behavior can be expected because the action of increasing the size of the particles while holding the gap constant is very similar to uniformly scaling Maxwell's equations, and silicon permittivity has relatively minor dispersion in this spectral range.~\cite{Palik} Otherwise, the final parameter we consider in Fig.~\ref{fig:sweep}c, is the angle of incidence to address the practical limitations for in-plane excitation of nanoscale optical structures. It can be seen that the Fano resonance will persist up to roughly 45\degree~incidence, beyond which the magnetic response is dominated by the normal-incidence responses. \begin{figure}[!t] \centering \centerline{\includegraphics[width=0.9\textwidth]{{sweep.pdf}}} \caption{Simulation results, from CST Microwave Studio, of the extinction caused by the silicon quadrumer in Fig.~\ref{fig:CSTvsCEMDA}, when changing (a) the size of the gap between nanoparticles, (b) the size of the individual nanoparticles and (c) the angle of incidence of the applied plane wave, with respect to the quadrumer plane (maintaining $s$-polarization).} \label{fig:sweep} \end{figure} \newpage {\it Experimental verification.} To verify and validate the theoretical arguments presented above, we look for the existence of magnetic Fano resonances in a high index cluster experimentally. In this regard, one practical option is to mimic the scattering properties of silicon nanoparticles using MgO-$\mathrm{TiO_2}$ ceramic spheres characterized by dielectric constant of 16 and dielectric loss factor of $(1.12  -  1.17)\times10^{-4}$, measured at {$9 - 12$ GHz}. These ceramic spheres in the microwave range therefore have very similar properties to silicon nanospheres in the optical range and they are subsequently a useful macroscopic platform on which to prototype silicon nanostructures. Here, they allow us to perform a `proof of concept' investigation into the properties of an isolated quadrumer with much more signal than would be expected from a single silicon nanosphere quadrumer. Indeed, such spheres have been used previously to predict the behavior of silicon nanoantennas.{~\cite{Filonov2012, Geffrin2012, Filonov2014}} \begin{figure}[!h] \centering \centerline{\includegraphics[width=0.9\textwidth]{{experiment.pdf}}} \caption{The (a) CST simulation results and (b) experimental measurements of extinction for a quadrumer made of four MgO-$\mathrm{TiO_2}$ ceramic spheres. We also show the extinction from a single MgO-$\mathrm{TiO_2}$ sphere for reference. In both simulation and experiment we observe the existence of a sharp Fano resonance occurring at 5.4 GHz. } \label{fig:experiment} \end{figure} The MgO-$\mathrm{TiO_2}$ quadrumer consists of four dielectric spheres with diameter $d=15\mathrm{mm}$, and the size of the gap between the particles is $s=5\, \mathrm{ mm}$. The experimentally measured, and numerically calculated, total scattering of the quadrumer structure are shown in Fig.~\ref{fig:experiment}. It can clearly be seen that a magnetic Fano resonance is produced at 5.4 GHz, in both simulation and experiment. This is the first example of a magnetic-magnetic Fano resonance in a single symmetric metamolecule. Notably, this Fano resonance occurs in a spectral range where the single particle is not at resonance, which demonstrates its collective nature. Indeed, it appears near the intersection of the single particle's electric and magnetic scattering contributions, reflecting that the overlap of eigenmodes (Eq.~\ref{eq:overlap}) is dependent on both electric and magnetic dipole polarizabilities. \section{Conclusions} We have presented a comprehensive study of the interplay between the collective optically-induced magnetic responses of all-dielectric quadrumers and the individual magnetic responses of their constituent dielectric nanoparticles. We have been able to establish the theoretical basis behind the interaction between collective and individual magnetic responses in all-dielectric structures, providing a quantitative prediction of the interference between the quadrumer's magnetic responses leading to magnetic-magnetic Fano resonance features. We have also been able to experimentally observe the existence of a sharp magnetic-magnetic Fano resonance in a dielectric quadrumer. Such Fano resonance features demonstrate the unique potential that suitably designed dielectric nanoclusters have for scattering engineering at the nanoscale, opening exciting and unexplored opportunities for dielectric nanophotonics. \section{Methods} To fasten together the MgO-$\mathrm{TiO_2}$ ceramic spheres for the experiment, we used a custom holder made of a styrofoam material with dielectric permittivity of 1 (in the microwave frequency range). To approximate plane wave excitation, we employed a rectangular horn antenna (TRIM $0.75 - 18$ GHz ; DR) connected to the transmitting port of a vector network analyzer (Agilent E8362C). The quadrumer was then located in the far-field of the antenna, at a distance of approximately 2.5 m, and a second horn antenna (TRIM $0.75 - 18$ GHz ; DR) was used as a receiver to observe the transmission through the quadrumer. The extinction measurement shown in Fig.~\ref{fig:experiment}b was then obtained as the difference between the measured transmission and unity transmission ({\it i.e.}, with no quadrumer). For the theory, CST simulations were performed assuming plane wave excitation on a quadrumer located in free space. The CEMDA simulations, seen in Fig.~\ref{fig:CSTvsCEMDA}a, used electric and magnetic dipole polarizabilities that were derived from the scattering coefficients of Mie theory.~\cite{Mie1908} In all simulations, silicon permittivity data was taken from Palik's Handbook~\cite{Palik} and the MgO-$\mathrm{TiO_2}$ spheres were assumed be dispersionless with a dielectric constant of 16 and dielectric loss factor of $(1.12  -  1.17)\times10^{-4}$. { \section{Acknowledgements} This work was supported by the Australian Research Council. F.M. and A.A. have been supported through the U.S. Air Force Office of Scientific Research with grant No. FA9550-13-1-0204, P02. The measurements were supported by the Government of the Russian Federation (grant 074-U01), the Ministry of Education and Science of the Russian Federation, Russian Foundation for Basic Research, Dynasty Foundation (Russia). } \bibliographystyle{apsrev4-1}
5f5a36d9001473ed2f90eaddfcc7552fbdbfe869
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} The original motivation for warped extra-dimensions was to address the hierarchy problem, so that the fundamental scale of gravity is exponentially reduced along the extra dimension, from the Planck mass scale to the TeV scale. Thus, the TeV scale becomes the natural scale of the Higgs sector if this one is localized near the TeV boundary of the extra dimension, as first introduced by Randall and Sundrum (RS) \cite{RS1}. If SM fields are allowed to propagate in the extra dimension \cite{Davoudiasl:1999tf}, the scenario can also address the flavor puzzle of the SM, explaining fermion masses and mixings from the geographical location of fields along the extra dimension. However, processes mediated by the heavy resonances of the 5D bulk fields, Kaluza-Klein (KK) modes, generate dangerous contributions to electroweak and flavor observables (including dangerous deviations to the $Zb\bar{b}$ coupling) \cite{Burdman,RSeff,Agashe}, pushing the KK mass scale to $5-10$ TeV \cite{Weiler}. A popular mechanism to lower the KK scale involves using a custodial gauge $SU(2)_R$ symmetry \cite{custodial}, which ensures a small contribution to electroweak precision parameters, lowering the KK scale bound to about 3 TeV. Alternatively, one can study scenarios in which the metric is slightly modified from the RS metric background ($AdS_5$). This can be achieved quite naturally from the backreaction on the metric caused by a 5D scalar field stabilizing the original $AdS_5$ warped background \cite{Goldberger:1999uk}. When the 5D Higgs is sufficiently leaking into the bulk and when the metric background is modified near the TeV boundary, the scenario allows for KK scales as low as 1-2 TeV, with precision electroweak and flavor constraints under control \cite{Cabrer}. An inconvenience is that these scenarios are typically hard to probe experimentally as the couplings of all particles are very suppressed \cite{Falkowski:2008fz,deBlas:2012qf}. Still, it has been shown that it can still lead to interesting deviations in Higgs phenomenology, as the Higgs couplings can receive sufficient radiative corrections from the many KK fermions of the model \cite{Frank:2015zwd}. Run 2 LHC data at $\sqrt{s}=13$ TeV shows signals of a new resonance in the diphoton distribution at an invariant mass of 750 GeV with a 3.9 significance at ATLAS \cite{ATLAS750}, with 3.2 fb$^{-1}$ and 3.4 combined significance at CMS \cite{CMS:2016owr}, with 2.6 fb$^{-1}$(combining run 1 and run 2 results). ATLAS reports 14 events and CMS, 10. The experimental data is summarized in Table \ref{tab:siglims}. \begin{table} \begin{center} \begin{tabular}{|c|c|c|} \hline\hline Channel & $8~{\rm TeV}, \sigma_{max}$ & $13~{\rm TeV}, \sigma_{max}$\\ \hline\hline $\gamma\gamma$ &$.21\pm.22~{\rm fb}$ \cite{ATLAS750} & $5.5\pm 1.5~{\rm fb}$ \cite{ATLAS750} \\ & $.63\pm .31~{\rm fb}$ \cite{CMS:2016owr} &$4.8\pm 2.1~{\rm fb}$ \cite{CMS:2016owr} \\ \hline \hline \hline $t {\overline t}$ & $700~{\rm fb}$ \cite{Aad:2015fna} & $\sim 2000~{\rm fb}$ \\ \hline $jj$ & $2.1~{\rm pb}$ \cite{Khachatryan:2016ecr}& $\sim 10~{\rm pb}$ \\ \hline $hZ$ & $19~{\rm fb}$ \cite{Aad:2015wra} & $\sim 100~{\rm fb}$ \\ \hline\hline \end{tabular} \caption{\label{tab:siglims} LHC diphoton cross sections at ATLAS and CMS and relevant signal limits for the $750$ GeV CP odd scalar resonance considered here. At 13 TeV, these are the rough implied bounds from the 8 TeV limits.} \end{center} \end{table} In light of all this, we propose here a simple and economic explanation within warped extra-dimensional models. It would require the presence of a 5D bulk Higgs, and because the mass of the new resonance is 750 GeV, the Higgs should be as much delocalized as possible from the TeV brane (but still close enough to address the hierarchy problem). The reason is that the masses of the Higgs KK excitations will increase as the Higgs is pushed towards the brane, getting infinitely heavy in the limit of a brane Higgs. Out of these excitations some are CP odd scalars, making them a natural candidate for the signal since they do not couple at tree-level to $ZZ$ or $WW$. We will show that if the typical mass of the KK gluon (typically the lightest and most visible KK particle) is around 1-2 TeV, it is very simple to obtain a 750 GeV CP odd Higgs with the help of small (and positive) brane localized kinetic terms of the 5D Higgs. Since the CP odd scalars do not couple at tree-level to $ZZ$ or $WW$, the largest coupling is going to be to pairs of tops. As will be shown, this coupling can be naturally small in wide regions of the allowed parameter space. This way, the radiative coupling to gluons, large enough for producing CP odd scalars, can also dominate the decays and the (also) radiative decay into photons can then receive enough branching fraction. Explanations of the 750 GeV signal within warped scenarios have been put forward previously, with the resonance interpreted as a radion \cite{Ahmed:2015uqt}, (and/or dilaton \cite{Cao:2016udb}), as a KK graviton \cite{Falkowski:2016glr,Hewett:2016omf,Carmona:2016jhr}, a 5D field-related axion \cite{Chakrabarty:2016hxi} or as an additional 5D singlet scalar added to the model \cite{Bauer:2016lbe}. The explanation proposed here, while preserving minimality and agreement with the diphoton excess, is also satisfied naturally in a significant region of the parameter space. We proceed as follows. In Sec. \ref{sec:background} we describe briefly the warped scenario, followed by its Higgs and gauge sector in Sec. \ref{sec:higgsgauge}, and of the CP-odd sector in more detail in Sec \ref{sec:cpodd}. Within that section we look at the fermion couplings in \ref{subsec:fermion}, the $\gamma \gamma$ and $glu-glu$ couplings in \ref{subsec:gggamgam} and to $Zh$ couplings in \ref{subsec:zh}. Our numerical estimates are presented in \ref{subsec:numerical} and we conclude in Sec. \ref{sec:discussion}. We leave some of the details for the Appendix. \section{The background metric} \label{sec:background} The (stable) static spacetime background is: \bea ds^2 = e^{-2\sigma(y)}\eta_{\mu\nu} dx^\mu dx^\nu - dy^2, \label{RS} \eea where the extra coordinate $y$ ranges between the two boundaries at $y=0$ and $y=y_{1}$, and where $\sigma(y)$ is the warp factor responsible for exponentially suppressing mass scales at different slices of the extra dimension. In the original RS scenario, $\sigma(y)=ky$, with $k$ the curvature scale of the $AdS_5$ interval that we take of the same order as $M_{Pl}$. Nevertheless this configuration is not stable as it contains a massless radion, a result of having the length of the interval not fixed. In more general warped scenarios with stabilization mechanism, $\sigma(y)$ is a more general (growing) function of $y$. We consider here the specific case where a 5D bulk stabilizer field backreacts on the $AdS_5$ metric producing the warp factor \cite{Cabrer,Carmona}. \bea\label{genericmetric} \sigma(y) = ky - {1\over \nu^2} \text{log}\left(1-{y\over y_s}\right), \eea where $y = y_s$ is the position of a metric singularity, which stays beyond the physical interval considered here, i.e. $y_s> y_{1}$. In these modified metric scenarios, the Planck-TeV hierarchy is reproduced with a shorter extra-dimensional length due to a stronger warping near the TeV boundary, so that whereas in RS we have $k y_{1}\simeq 35$, in the modified scenarios we can have $k y_{1}\simeq 20-30$. The appeal of this particular modification lies on the possibility of allowing for light KK particles ($\sim 1$ TeV), while keeping flavor and precision electroweak bounds at bay. This happens when the Higgs profile leaks sufficiently out of the TeV brane so that all of its couplings to KK particles are suppressed compared to the usual RS scenario \cite{Cabrer, Falkowski:2008fz,Carmona}. We thus fix the Higgs localization to a point where it is maximally pushed away from the IR brane, while still solving the hierarchy problem (i.e. making sure that we are not reintroducing a new fine-tuning of parameters within the Higgs potential parameters \cite{Cabrer,Quiros:2013yaa}.) \section{Gauge and Higgs sector} \label{sec:higgsgauge} The matter content of the model is that of a minimal 5D extension of the Standard Model, so that we assume the usual strong and electroweak gauge groups $SU(3)_c\times SU(2)_L\times U(1)_Y$, with all fields propagating in the bulk. The fermions of the model are also bulk fields, with different 5D bulk masses, so that their zero mode wavefunctions are localized at different sides of the interval. This way the scenario also addresses the flavor puzzle of the SM, since hierarchical masses and small mixing angles for the SM fermions become a generic feature due to fermion localization and small wavefunction overlaps \cite{Frank:2015sua}. In the electroweak and Higgs sector we consider the following action \bea \hspace{-1cm}&&S=\int d^4xdy \sqrt{g}\ \left(-\frac{1}{4} F^2_{MN} + |D^M H|^2 - V(H) \right) \ \ \ \ \\ \hspace{-1cm} && + \sum_{i=1}^{2}\int d^4xdy \sqrt{g}\ \delta(y-y_i) \left(\frac{d_i}{k} |D^M H|^2 - \lambda_i(H)\right) \label{5Daction} \eea where the capital index $M$ will be used to denote the $5$ spacetime directions, while the Greek index $\mu$ will be used for the 4D directions. The coefficients $d_i$ (in units of $k$) are essentially free parameters encoding the importance of brane localized kinetic terms associated with the bulk Higgs field. These terms will allow for a slight modification of the spectrum of the KK Higgs excitations, particularly useful in reproducing a $750$ GeV CP-odd excitation. These brane kinetic terms can be thought of as exactly localized operators, or as bulk operators that happen to be dynamically localized due to couplings to some localizer VEV\footnote{In order to avoid tachyons and/or ghosts, the sign of the purely brane localized brane kinetic terms will be kept positive, i.e. $d_i>0$.}. The 5D Higgs doublet is expanded around a nontrivial VEV profile $v(y)$ as \bea \label{Hexpansion} H=\frac{1}{\sqrt{2}} e^{i g_5 \Pi}\left(\begin{matrix} 0\\ v(y) +h(x,y) \end{matrix} \right) \eea and the covariant derivative is $D_M= \partial_M + i g_5 A_M$ with \bea \label{Aexpansion} A_M=\left( \begin{matrix} s_W A_M^{em}+\frac{c_W^2-s_W^2}{2c_W}Z_M & \frac{1}{\sqrt{2}}W_M^+\\ \frac{1}{\sqrt{2}}W_M^-& -\frac{1}{2c_W} Z_M \end{matrix} \right) \eea and CP-odd and charged Higgs part is \bea \label{Piexpansion} \Pi=\left( \begin{matrix} \frac{c_W^2-s_W^2}{2c_W}\Pi_z & \frac{1}{\sqrt{2}} \Pi^+\\ \frac{1}{\sqrt{2}}\Pi^-& -\frac{1}{2c_W} \Pi_z \end{matrix} \right) \eea with the weak angle defined like in the SM, i.e. $s_W/c_W=g'_5/g_5$, where $g_5$ and $g_5'$ are the 5D coupling constants of $SU(2)_L$ and $U(1)_Y$. The extraction of degrees of freedom in this context has been performed in \cite{Quiros:2013yaa,Falkowski:2008fz,Archer:2012qa} and we outline here the main results. The effect of brane kinetic terms in the Higgs sector is new and its derivation is outlined in the Appendix. The 5D equations of motion for all these fields are coupled (except for the case of the real Higgs excitation $h(x,y)$) and in order to decouple them, one can partially fix the gauge, or add a gauge fixing term to the previous 5D action. For example, in the CP-odd case, the fields $Z_\mu(x,y)$, $Z_5(x,y)$ and $\Pi_z(x,y)$ must be unmixed. The partial gauge fixing constraint\footnote{There is still be some gauge freedom left, so that the towers of 4D Goldstone bosons that appear can be gauged away.} \bea \partial^\mu Z_\mu - M^2_z(y) \Pi_z +(e^{-2\sigma}Z_5)'=0 \eea manages to decouple the fields $Z_\mu$ from $Z_5$ and $\Pi_z$ in the bulk. We defined here $M_z(y)=\frac{g_5}{2c_W} v(y) e^{-\sigma(y)}$. However, the presence of the Higgs brane kinetic terms, proportional to $d_i$ in the action, forces us to extend the gauge choice on the branes, producing a lifting of $Z_5$ field so that the decoupling is maintained at the boundaries\footnote{In the absence of brane kinetic terms, $Z_5$ must have vanishing boundary conditions (Dirichlet) if $Z_\mu$ is to have Neumann conditions and thus develop a zero mode KK excitation in the effective 4D theory.}. The appropriate boundary condition at the IR brane is \bea Z_5(x,y_1) =- \frac{d_1}{k} M^2_z(y_1) e^{2\sigma(y_1)} \Pi_z(x,y_1) \eea where $y_1$ denotes the position of the boundary (note that if the brane kinetic term parameter $d_1$ tends to zero, the condition on $Z_5$ becomes Dirichlet, as expected). With this type of gauge choice, the 5D fields $Z_\mu$, $W_\mu$ and $A_\mu$ have independent 5D equations of motion. In order to extract the effective 4D degrees of freedom, we expand the gauge fields as $Z_\mu(x,y)=Z^n_\mu(x) f^n_z(y)$, $W_\mu(x,y)=W^n_\mu(x) f^n_w(y)$ and $A_\mu(x,y)=A^n_\mu(x) f^n_\gamma(y)$ (summation over $n$ is understood) and where $Z^0_\mu(x)$, $W^0_\mu(x)$ and $A^0_\mu(x)$ are the $Z$, $W$ and $\gamma$ gauge bosons of the SM. The extradimensional profiles $f^n_z(y)$, $f^n_w(y)$ and $f^n_\gamma(y)$ are solutions of \bea\label{gaugeeq} \left(e^{-2\sigma} f'_a\right)'+(m^2_{n} - M_a^2(y)) f_a =0 \eea where $a=z,w,\gamma$ and where $M_z(y)=\frac{g_5}{2c_W} v(y) e^{-\sigma(y)}$, as defined before, $M_w(y)=\frac{g_5}{2} v(y) e^{-\sigma(y)}$, and $M_\gamma=0$. The boundary conditions for these profiles are\footnote{We ignore here possible brane localized gauge kinetic terms and keep only the effects from Higgs brane kinetic terms. We include everything in the derivation outlined in the Appendix.} \bea\label{gaugebc} \frac{d_i}{k} M_a^2 f_a(y_i) = - e^{-2\sigma}f'_a(y_i). \eea The CP-even Higgs field is expanded as $h(x,y)=h^n(x) h^n_y(y)$ and the equations for the Higgs profiles are, with $h_y \equiv h_y^n(y)$: \bea \left(e^{-4\sigma} h'_y\right)'+(m^2_{h_n} - \mu^2_{bulk}) h_y =0, \eea where $\displaystyle \mu^2_{bulk} = \frac{\partial^2 V }{\partial H^2}\Big|_{H=v}$. The boundary conditions are \bea \left(\mu^2_{brane_i} - \frac{d_i}{k}\ m^2_{h_n} e^{2\sigma}\right) h_y = -e^{-4\sigma} h'_y, \eea with $\displaystyle \mu^2_{brane_i} = \frac{\partial^2 \lambda_i }{\partial H^2}\Big|_{H=v}$. Note that the mode $h^0(x)$ is interpreted as the SM Higgs boson. There are still some degrees of freedom left, and their 5D equations of motion still happen to be mixed. One of the coupled systems involves $Z_5$ and $\Pi_z$ and the other coupled system involves $\Pi^{\pm}$ and $W_5^\pm$. In order to disentangle these systems one must perform a mixed expansion, so that the decoupling of fields will happen KK level by KK level. The mixed expansions are, in the CP-odd sector, \bea\label{Z5decomp} \hspace{-.6cm}Z_5(x,y)&=& G^n(x)\frac{f'_{G_n}(y)}{m^2_{G_n}} + \Pi_n(x)\frac{e^{2\sigma}}{m^2_{\pi_n}}X_\pi(y) \\ \hspace{-.6cm}\Pi_z(x,y)&=& G^n(x)\frac{f_{G_n}(y)}{m^2_{G_n}} + \Pi_n(x) \frac{1}{m^2_{\pi_n}M^2_{z} }X_{\pi}'(y), \label{Pidecomp} \eea and in the charged scalar sector they are \bea \hspace{-.6cm}W_5^\pm(x,y) &=& G_n^\pm(x) \frac{f_{G_n^\pm}(y)}{m^2_{G_n^\pm}} + \Pi^\pm_n(x)\frac{e^{2\sigma}}{m^2_{\pi_n^\pm}}X_\pm(y)\ \ \\ \hspace{-.6cm}\Pi^\pm(x,y)&=& G_n^\pm(x) \frac{f_{G_n^\pm}(y)}{m^2_{G_n^\pm}} +\Pi^\pm_n(x) \frac{1}{m^2_{\pi_n^\pm}M^2_{w} }X_\pm'(y).\ \ \eea where $M_z(y)$ and $M_w(y)$ were defined below Eq~(\ref{gaugeeq}). The effective 4D physical fields are the tower of CP-odd neutral scalars $\Pi_n(x)$ and the tower of charged scalars $\Pi^\pm_n(x)$. Their associated extra-dimensional profiles $X_\pi(y)$ and $X_\pm(y)$ obey the equations \bea \left(\frac{1}{M_a^2(y)} X_a' \right)'+ \left(\frac{m^2_{\pi_a}}{M_a^2(y)}- 1 \right) e^{2\sigma} X_a=0 \label{Xeq} \eea where $M_a(y)=(M_z(y),M_{w}(y))$ and $X_a=(X_\pi,X_{\pm})$. The boundary conditions are \bea\label{Xbc} \frac{d_i}{k}\ X'_a= -X_a, \eea and note that vanishing Higgs brane kinetic terms implies Dirichlet boundary conditions for $X_a$. We checked that these bulk equations agree with \cite{Falkowski:2008fz,Quiros:2013yaa,Archer:2012qa}, the only new addition being the boundary conditions imposed by the presence of Higgs brane kinetic terms. In order for these 4D scalars to be canonically normalized, we require \bea \frac{1}{m^2_a}\int dy\ e^{2\sigma} \frac{X_a^2}{M_a^2}=1 \eea and this condition includes the effect of Higgs brane kinetic terms. The remaining 4D fields are $G_n(x)$ and $G_n^\pm(x)$, which are Goldstone bosons at each KK level. The profile wavefunctions $f_{G_a}(y)$ obey the same differential equations as the gauge profiles, Eq.(\ref{gaugeeq}), as well as the same boundary conditions, Eq.(\ref{gaugebc}). The spectrum is thus identical to the gauge bosons spectrum level by level. These fields appear in the effective 4D action coupled to ($\partial^\mu Z^n_\mu$) or ($\partial^\mu W^n_\mu$), and of course there is a leftover gauge freedom allowing us to gauge them away (i.e., they are pure gauge). We wish to identify the lightest CP-odd scalar $\Pi_0(x)$ with the observed diphoton peak at the LHC, so that we need to fix its mass $m_{\Pi_0}= 750$~GeV. In order to have an idea of the effects of the Higgs brane kinetic terms on the CP-odd scalar spectrum, we consider two different parameter points, one in which the RS background metric is recovered with $\nu=10$ and $y_s=4\times y_{1}$ ($\nu$ is the exponent appearing in the modified metric and if relatively large, the location of spurious singularity is sent away from the boundary, recovering essentially the $AdS_5$ metric). The other case is the situation where the metric modification allows for TeV size KK masses, which are safe from precision electroweak constrains. The parameters chosen there are $\nu=0.5$ and $y_s=1.04\times y_{1}$. In both parameter points, we fix the KK mass of the first gluon excitation to be $1500$ GeV\footnote{Of course, this RS point is presented for comparison only, since such light KK masses will produce too large deviations in the precision electroweak observables.}. \begin{figure}[t] \center \begin{center} \includegraphics[height=9.5cm]{XmassplotRS.pdf}\ \ \includegraphics[height=9.5cm]{Xmassplot.pdf} \end{center} \vspace{-.2cm} \caption{Mass spectrum of the first CP-odd Higgs scalars as a function of the brane kinetic term coefficient $d_{1}$ in the RS metric limit ($\nu=10$ and $y_s=4\ y_{1}$; left panel) and within a noticeably IR-modified $AdS_5$ metric ($\nu=0.5$ and $y_s=1.04\ y_{1}$; right panel). In both cases, the first KK gluon mass is fixed at 1500 GeV. One can see that in RS it is not possible to obtain a 750 GeV CP-odd scalar mass, whereas within the modified metric, a brane kinetic coefficient $d_{1}\simeq 0.51$ does produce a mass of 750 GeV, shown here as a dot. } \label{fig:Xmassplot} \vspace{-.2cm} \end{figure} In Fig. \ref{fig:Xmassplot} we show the spectrum of the first 4 KK levels of CP-odd Higgs bosons $\Pi_n(x)$, for $n=0,1,2,3$, as a function of the brane kinetic term $d_{1}$ in units of the curvature $k\sim M_{Pl}$. The effects of the UV localized brane kinetic term are warped suppressed and so we do not consider them here anymore. We can see that in the RS limit, it is not possible to reduce sufficiently the lightest CP-odd mass to 750 GeV, as its mass tends asymptotically from about $2500$ GeV without brane kinetic term to about $1750$ GeV for large brane kinetic terms. On the other hand, with the modified metric it becomes possible to reduce greatly the lightest CP-odd mass with relatively small brane kinetic term coefficients (which in this particular case tends asymthotically to about 400 GeV). Parameter points such that the metric modification lies between the two considered will actually have an intermediate behavior, with a lightest CP-odd mass having increasing asymptotic values as one recovers the RS background. Finally also note that the spectrum for the charged scalars is essentially the same since their differential equations and boundary conditions are identical except for the functions $M_z(y)$ and $M_w(y)$, which differ by about $10\%$. $M_w(y)$ produces a deviation from the CP-odd scalar spectrum of less than $5\%$. This means that the scenario under consideration should also contain a lightest charged Higgs scalar with a mass of about $750$ GeV also. The next question to ask is how big is the effect of the Higgs brane kinetic term on the gauge bosons, and in particular on the lowest ones, i.e. the SM $W$ and $Z$ bosons. These terms represent an additional (brane localized) contribution to the mass of the gauge bosons. In principle their mass is generated here from a bulk Higgs mechanism, unless the brane kinetic terms are overly important (not the limit we are working with here). We can quickly estimate its effect on the lowest lying gauge fields. These are essentially flat (like all gauge zero modes) and thus their wave function is $f^0_z\sim 1/\sqrt{y_{1}}$. The contribution of a brane localized mass squared term is $\delta m^2_z \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} d_{1}\ v^2/y_{1}\sim d_{1}\times 700$ GeV$^2 $. For IR brane kinetic term coefficient $d_{1}$ of ${\cal O}(1)$, this represents naively at most some $10\% $ contribution to the overall mass squared of either $W$ or $Z$. In the particular case of the modified metric with a brane coefficient $d_{1}\simeq 0.51$, and metric parameters $\nu =0.5$ and $y_s=1.04 \times y_1$ (which produces a light CP-odd scalar of $750$ GeV), the exact numerical effect on the zero mode gauge boson masses ($W$ and $Z$) is a shift of $3$ GeV with respect to the no-brane-kinetic-term limit. Of course, in the presence of brane kinetic terms, one redefines the VEV normalization constant, and the value of $g_5$, in order to correctly account for the SM gauge boson masses and electroweak couplings. \section{CP-odd Higgs couplings} \label{sec:cpodd} As a $750$ GeV CP-odd Higgs scalar $\Pi_0(x)$ is allowed in the spectrum, thanks to the effect of small brane localized Higgs kinetic terms, we now study its couplings to SM particles in order to see if the observed excess at the LHC can be associated with this excitation. Of course being a CP-odd scalar its tree-level couplings to $ZZ$ and $WW$ are zero, making it an ideal candidate for the observed exotic events. We thus need to focus on its tree-level couplings to fermions (and top quark in particular), to $Zh$ (where $h$ is the $125$ GeV Higgs) and to its radiative couplings to photons and gluons. We study these in the subsequent subsections. \subsection{Fermion couplings} \label{subsec:fermion} The couplings of $\Pi_0(x)$ to fermions arise from two sources in the action. First source comes from the 5D Higgs Yukawa couplings, and second, from the gauge fermion couplings. This is because the physical field $\Pi_0(x)$ contains some of CP-odd Higgs scalar, and some of $Z_5$ excitation, where $Z_5$ is the fifth component of the 5D vector boson $Z_M$. However the 5D Yukawa coupling allows for direct coupling of $\Pi_0(x)$ to two zero-mode fermions, whereas the gauge-fermion coupling allows only couplings between fermion zero-modes and higher KK fermion levels. As we will see, it is important to keep both couplings, since after electroweak symmetry breaking the physical SM fermions (top quarks in particular) are mostly zero-modes but also contain a small amount of higher KK excitations, and could thus inherit some of the original gauge-fermion coupling, especially if the tree-level Yukawa coupling between $\Pi_0(x)$ and zero-mode top quarks is suppressed (as it can be). The relevant terms in the action are the 5D Higgs Yukawa couplings and the fermion gauge interaction term, \bea S_{ff\Pi_0} \subset \int d^4xdy \sqrt{g}\left[ Y_u HQU + Y_d HQD + h.c.+ \bar{Q} {\cal D}\hspace{-3mm}/\ Q + \bar{U} {\cal D}\hspace{-3mm}/\ U + \bar{D} {\cal D}\hspace{-3mm}/\ D \right] \eea where $Q,U,D$ represent the 5D fermion $SU(2)_L$ doublets, up-type and down-type singlets (with generation indices and isospin indices suppressed). The kinetic terms contain the 5D covariant derivative and from them we extract the terms containing the CP-odd component $Z_5(x,y)$, and from the Higgs Yukawa couplings we extract the terms containing the CP-odd Higgs component $\Pi_z(x,y)$. We follow the approach of \cite{Frank:2015zwd,Diaz-Furlong:2016ril} and compute these couplings by considering only the effects of three full KK levels, i.e. computing $21 \times 21$ fermion Yukawa coupling matrices (with 3 $up$ and 3 $down$ families, each containing zero modes and 3 KK levels with an $SU(2)_L$ doublet and a singlet in each level, i.e., 3 zero modes plus $3\times 3 \times 2$ KK modes). Note that we are interested in the couplings of the $750$ GeV CP-odd scalar $\Pi_0(x)$ to SM fermions (top quarks primarily), but we also need its couplings with the rest of KK fermions, since these interactions will be crucial to generate large enough radiative couplings to photons and gluons. We first write the effective 4D up-type quark mass matrix as \bea \left(\begin{matrix} q^0_L(x)\ Q_L(x)\ U_L(x) \end{matrix}\right)\ M_u\ \left(\begin{matrix} u^0_R(x)\\ Q_R(x)\\ U_R(x) \end{matrix}\right) \eea in a basis where $q^0_L(x)$ and $u_R^0(x)$ represent three zero-mode flavors each (doublets and singlets of $SU(2)_L$), and $Q_L(x)$ and $Q_R(x)$ represent three flavors and three KK levels of the vector-like KK up-type doublets, and $U_L(x)$ and $U_R(x)$ represent three flavors and three KK levels of vector-like KK up-type singlets. The mass matrix is thus \bea M_u= \left(\begin{matrix} (y^{0}_{u})_{3\times3} & (0)_{3\times 9} & (Y^{qU})_{3\times 9}\\ (Y^{Qu})_{9 \times3} & (M_Q)_{9\times 9} & (Y_1)_{9 \times 9}\\ (0)_{9 \times3} & (Y_2)_{9\times 9} & (M_U)_{9\times 9} \end{matrix}\right) \label{mumatrix} \eea with the down sector mass matrix ${M}_d$ computed in the same way. The submatrices are obtained by evaluating the overlap integrals \bea\label{YYY1KK} y^0_u= \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{v(y)}{\sqrt{2}} q^{0,i}_{L}(y)u^{0,j}_{R}(y) \\ Y^{qU}= \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{v(y)}{\sqrt{2}} q^{0,i}_{L}(y)U^{n,j}_{R}(y) \\ Y^{Qu} = \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{v(y)}{\sqrt{2}} Q^{m,i}_{L}(y)u^{0,j}_{R}(y)\\ Y_1 = \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{ v(y)}{\sqrt{2}} Q^{m,i}_{L}(y)U^{n,j}_{R}(y) \\ Y_2 = \frac{(Y^{5D^*}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{ v(y)}{\sqrt{2}} Q^{m,i}_{R}(y)U^{n,j}_{L}(y)\, , \eea where the indices $m$ and $n$ track the KK level and $i,j=1,2,3$ are 5D flavor indices. The diagonal matrices $(M_Q)_{9\times 9}$ and $(M_U)_{9\times 9}$ are constructed with the masses of all the KK quarks involved. The masses and the profiles of the KK fermions appearing in these overlap integrals ($Q_L(y)$, $Q_R(y)$, $U_L(y)$ and $U_R(y)$) are obtained by solving differential equations for the fermion profiles \bea \partial_y\left(e^{(2c-1)\sigma(y)}\partial_y\left( e^{-(c+2)\sigma(y)}\right)\right)f(y)+e^{(c-1)\sigma(y)} m_n^2 f(y) =0 \label{fermionprofile} \eea where $f(y)$ is the KK profile. The mass eigenvalues $m_n$ are found by imposing Dirichlet boundary conditions on the wrong chirality modes. As mentioned before, we have included 3 full KK levels so that the mass matrices in the gauge basis are $21 \times 21$ dimensional matrices, which are not diagonal. One needs to diagonalize them, and by doing so, move to the quark physical basis where all the fermions couplings can then be extracted. In the CP-odd scalar sector, we can write the effective 4D Yukawa-type couplings to fermions in the same gauge basis as before \bea \left(\begin{matrix} q^0_L(x)\ Q_L(x)\ U_L(x) \end{matrix}\right) {\bf Y}_\pi \left(\begin{matrix} u^0_R(x)\\ Q_R(x)\\ U_R(x) \end{matrix}\right) \ \Pi_0(x) \eea where now the $21 \times 21$ coupling matrix ${\bf Y}_{\pi}$ is given by \bea {\bf Y}_{\pi}= \left(\begin{matrix} (y^{0}_{\pi qu})_{3\times3} & (a_{\pi qQ})_{3\times 9} & (Y_{\pi qU})_{3\times 9}\\ (Y_{\pi Qu})_{9 \times3} & (a_{\pi QQ})_{9\times 9} & (Y_{1}^{\pi})_{9 \times 9}\\ (a_{\pi uU})_{9 \times3} & (Y_{2}^{\pi})_{9\times 9} & (a_{\pi UU})_{9\times 9} \end{matrix}\right) \label{Ypimatrix} \eea The submatrices are obtained by the overlap integrals \bea\label{YYPI} y^{0}_{\pi qu} &=& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} q^{0,i}_{L}(y)u^{0,j}_{R}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y_{\pi qU}&=& i\frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} q^{0,i}_{L}(y)U^{n,j}_{R}(y)\frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y_{\pi Qu}& =& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q^{m,i}_{L}(y)u^{0,j}_{R}(y)\frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y^\pi_1& =& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q^{m,i}_{L}(y)U^{n,j}_{R}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y^\pi_2& =& i \frac{(Y^{5D^*}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q^{m,i}_{R}(y)U^{n,j}_{L}(y)\frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \, , \eea and \bea a_{\pi qQ}&=& \frac{g^{5D}_{L}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} q^{0,i}_{L}(y)Q^{n,j}_{R}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \\ a_{\pi uU}&=&\frac{g^{5D}_{R}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} u^{0,i}_{R}(y)U^{n,j}_{L}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \\ a_{\pi QQ}&=&\frac{g^{5D}_{L}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} Q^{m,i}_{L}(y)Q^{n,j}_{R}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \\ a_{\pi UU}&=&\frac{g^{5D}_{R}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} U^{m,i}_{R}(y)U^{n,j}_{L}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \eea where the $g^{5D}_{L,R}$ coupling are given by \bea g^{5D}_{L} = {g^{5D}\over \cos \theta_W } \left(T_3 - Q_q \sin^2 \theta_W \right) \eea \bea g^{5D}_{R} = {g^{5D}\over \cos \theta_W} Q_q \sin^2 \theta_W , \eea with $Q_q$ the charge of the quark, (here ${2 \over 3}$), $\theta_W$ the weak angle and $T_3 = {1\over 2}$. Note that when the interaction originates in the 5D Yukawa couplings, the profile to use is the one coming from the CP-odd Higgs component, i.e proportional to $X_\pi'(y)$. When the interaction originates in the gauge fermion coupling and thus comes from the $Z_5$ component, the profile to use is proportional to $X_\pi(y)$, with $X_\pi$ being the solution of Eq.~(\ref{Xeq}), using the decompositions of Eqs.~(\ref{Z5decomp}) and (\ref{Pidecomp}). When the fermion matrix in (\ref{mumatrix}) is diagonalized, the coupling matrix of fermions with the CP-odd field $\Pi_0(x)$ in (\ref{Ypimatrix}) is rotated, and we can then extract all the physical Yukawa couplings. All these couplings are needed later in order to compute the radiative couplings of $\Pi_0(x)$ with gluons and photons. \begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{yPiPoint6.pdf}\ \ \includegraphics[height=7cm]{yPiPoint5.pdf} \includegraphics[height=7cm]{yPiPoint4.pdf}\ \ \includegraphics[height=7cm]{yPiPoint2.pdf} \end{center} \vspace{-.2cm} \caption{Yukawa couplings between zero-mode fermions and the two lightest neutral scalars of the scenario, the $125$ GeV Higgs and the $750$ GeV CP-odd $\Pi_0$. The couplings are evaluated relative to the 5D bulk Higgs Yukawa coupling $Y_5$ and are shown as a function of the fermion bulk mass parameter $c_q$ (in the case $c_q=-c_u$ for simplicity) for different overall KK scales, $M_{KKglu_1}=3900$ GeV (upper left panel), $M_{KKglu_1}=2400$ GeV (upper right panel), $M_{KKglu_1}=1400$ GeV (lower left panel) and $M_{KKglu_1}=1000$ GeV (lower right panel). The CP-odd scalar mass is set to $750$ GeV and for certain values of $c_q$ its Yukawa coupling to top quarks can be highly suppressed for typical top-quark values of the $c_i$'s. } \label{fig:yPiplot} \vspace{-.2cm} \end{figure} We first analyze the very important Yukawa coupling between $\Pi_0(x)$ and top quarks, as this coupling might dominate the decays of the CP-odd scalar. This coupling comes essentially from the entry $(y^0_{\pi qu})_{ 33}$ (before rotation to the physical basis) although it receives small corrections after going to the physical basis. We focus on $(y^0_{\pi qu})_{ 33}$ which comes from the overlap integral \bea\label{YPtt} (y^{0}_{\pi qu})_{33} &=& i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} q^{0}_{t}(y)u^{0}_{t}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \eea In the RS limit, the warp factor is $\sigma(y)=ky$, and the top profiles are $q^{0}_{t}(y)= f(c_q) e^{(2-c_q)ky}$ and $u^{0}_{t}(y)=f(-c_u) e^{(2+c_u)ky} $, where $f(x)$ is a normalization factor. We also have $M_z(y)=\frac{g_5}{2c_W} v_0 e^{(a-1)ky}$, with $v_0$ a constant factor, so that the previous overlap integral in this limit reads \bea (y^{0}_{\pi qu})_{33} = i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \frac{2 c_W}{g_5}\frac{f(c_q) f(-c_u)}{v_0 m^2_{\pi_0}}\int_0^{y_1} dy\ e^{(2-a-c_q+c_u)k y} X_\pi'(y) \eea We integrate this by parts to find \bea \!\!\!(y^{0}_{\pi qu})_{33} &=& -i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \frac{2 c_W}{g_5}\frac{f(c_q) f(-c_u)}{v_0 m^2_{\pi_0}}\int_0^{y_1} dy\ (2-a-c_q+c_u) e^{(2-a-c_q+c_u)k y} X_\pi(y) + BT\ \ \ \ \ \ \eea where $BT= i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \frac{2 c_W}{g_5}\frac{f(c_q) f(-c_u)}{v_0 m^2_{\pi_0}}\ e^{(2-a-c_q+c_u)k y} X_\pi(y)\Big|_0^{y_1} $ is a boundary term. Note that the profile $X_\pi(y)$ has vanishing boundary conditions in the absence of Higgs localized brane kinetic terms. In that limit we can see that the coupling of the CP-odd scalar can actually vanish, when $(2-a-c_q+c_u)=0$ \cite{Archer:2012qa}. Note also that the Higgs localizer parameter $a$ is, in this RS limit, $a \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2$ and the bulk parameters $c_q$ and $c_u$ are defined such that, for example, charm or bottom quarks are assigned values more or less $c_q \in (0.45,0.55)$ and $c_u\in (-0.5,-0.6) $, whereas for top we have $c_{q_3}\sim 0.45$ and $c_{u_3}> -0.45$. This means that we expect the term $(2-a-c_q+c_u)$ to vanish, in the limit of $a\sim 2$, when $c_q-c_u \sim 0$, so that the suppression in this case seems only possible for the top quark, where both $c_q$ and $c_u$ could be small. Of course when the metric background is modified, and the boundary conditions include the brane kinetic terms there will be deviations from the previous values. Nevertheless it is clear that the Yukawa coupling of the CP-odd scalar field to top quarks can have highly suppressed values. Another way to see this is to consider the overlap integral in Eq.~(\ref{YPtt}). Because the profile $X_\pi(y)$ vanishes at the boundaries (or $almost$ vanishes, for small brane kinetic terms), then its derivative $X'_\pi(y)$ will have a node in the bulk, and therefore will change sign. That means that there can be some parameter choice for which it is possible for the overlap integral to vanish, since the fermion zero mode profiles have no nodes in the bulk. This feature is clearly seen in Fig. \ref{fig:yPiplot}, where we plot the absolute value of the Yukawa couplings between zero-mode fermions and both the Higgs and the CP-odd scalar $\Pi_0(x)$.\footnote{We are actually plotting the values defined in Eqs. (\ref{YYPI}) and (\ref{YYY1KK}), i.e. the zero mode Yukawa couplings before going to the fermion mass basis. In that basis, the couplings will inherit a small correction due to mixing with heavy KK fermions \cite{Azatov:2010pf}, so that the exact cancellation of the coupling will be replaced by a strong suppression.} The couplings shown are relative to the 5D bulk Higgs Yukawa coupling $Y_5$ and are plotted as functions of the fermion bulk mass parameter $c_q$ and $c_u$ (for the case where we take $c_q=-c_u$, for simplicity), for different overall KK scales. We observe that the CP-odd Yukawa couplings are fairly similar to the Higgs Yukawa couplings (i.e. exponentially sensitive to UV localization and then top-like when the zero mode is IR localized) except that there is a range of parameters where the coupling vanishes. Interestingly enough, this suppression happens for preferred values of the top quark bulk mass parameters. This means that the existence of suppressed couplings to top quarks of the CP-odd $\Pi_0$ is a natural possibility in this scenario. \subsection{Radiative couplings to photons and gluons} \label{subsec:gggamgam} Just like in the Higgs boson case, the radiative couplings of $\Pi_0(x)$ to gluons and photons will depend on the physical Yukawa couplings $y_{nn}$ between $\Pi_0$ and the fermions (zero modes and KK modes) running in the loop, as well as on the fermion masses $m_n$ (the eigenvalues of the mass matrix in Eq.(\ref{mumatrix})). The real and imaginary parts of the couplings are associated with different loop functions, $A^S_{1/2}$ and $A^P_{1/2}$, as they generate the two operators $\Pi_0G_{\mu\nu}G^{\mu\nu}$ and $\Pi_0 G_{\mu\nu}\tilde{G}^{\mu\nu}$.\footnote{The Yukawa couplings of $\Pi_0$ are mostly imaginary and thus the dominant contribution will come, as expected, from the operator $\Pi_0 G_{\mu\nu}\tilde{G}^{\mu\nu}$. Still, small real Yukawa coupling components are generated when going to the fermion mass basis, and so we keep the general formalism in our formulas.} The production cross section through gluon fusion is \bea \sigma_{gg\rightarrow \Pi_0} = {\alpha_s^2 m_{\Pi_0}^2\over 576 \pi} \left[\Big|\sum_{quarks}c^S_n\Big|^2 + \Big|\sum_{quarks} c^P_n\Big|^2\right]\ \eea and the decay widths to gluons and photons are \bea \Gamma_{\Pi_0\to gg} = \frac{\alpha_s^2 m^3_{\Pi_0}}{54\pi^2}\frac{1}{v^2}\left[\Big|\sum_{{quarks }}c^S_n\Big|^2 + \Big|\sum_{{quarks}} c^P_n\Big|^2\right]\ \eea \bea \Gamma_{\Pi_0\to \gamma\gamma} = \frac{\alpha^2 m^3_{\Pi_0}}{192\pi^3}\frac{1}{v^2} \left[\Big|\sum_{{quarks \atop leptons}} N_c Q_n^2 c^S_n\Big|^2 + \Big|\sum_{{quarks \atop leptons}} N_c Q_n^2 c^P_n\Big|^2\right]\ \eea where $\alpha_s$ and $\alpha$ are the strong and weak coupling constants, $N_c$ is the number of colors and $Q_n$ is the charge of the fermion, and where \bea c^S_n =\re\left({\frac{y_{nn}}{m_{n}}}\right) A^S_{1/2}(\tau_n) \hspace{.6cm} {\rm and} \hspace{.6cm} c^P_n = {\rm Im}\left(\frac{y_{nn}}{m_{n}}\right) A^P_{1/2}(\tau_n) \eea with $\ \tau_n = m^2_{\Pi_0}/4m^2_n\ $ and with the loop functions defined as \cite{Gunion:1989we} \bea\label{eq:loopfunction} A^S_{1/2}(\tau) &=& {3\over2}\left[\tau + (\tau -1)f(\tau)\right]\tau^{-2}, \\ A^P_{1/2}(\tau) &=&- {3\over2} f(\tau)/\tau , \eea and with \bea f(\tau) = \left \{ \begin{tabular}{cc} $\left[\arcsin\sqrt{\tau}\right]^2$ & $(\tau\le 1)$\\ $ -{1\over4}\left[\ln\left({1+ \sqrt{1-\tau^{-1}}\over 1- \sqrt{1-\tau^{-1}}}\right)-i\pi\right]^2$& $(\tau> 1)$ . \end{tabular} \right. \eea For heavy KK quarks with masses $m_n$ much greater than the CP-odd mass $m_{\Pi_0}$ (i.e. when $\tau$ is very small) the loop functions are essentially constant, as they behave asymptotically as $\displaystyle \lim_{\tau \to 0} A^S_{1/2} =1 \hspace{.2cm} \mbox{and} \hspace{.2cm} \lim_{\tau \to 0} A^P_{1/2} =3/2.$ On the other hand, for light quarks (all the SM quarks except top and bottom), the loop functions essentially vanish asymptotically as $\displaystyle \lim_{\tau \to \infty} A^S_{1/2} = \lim_{\tau \to \infty} A^P_{1/2} = 0.$ Moreover, we investigate a parameter region where the couplings of $\Pi_0$ to top quarks are highly suppressed. This means that the production mechanism must rely exclusively on the heavy KK fermions running in the loop and as we have seen, this coupling depends on the ratio $\frac{y_{nn}}{m_n}$ between the physical Yukawa coupling and the mass of the fermion running in the loop. To have an idea of the relative contribution of each of these KK fermions in the loop, in Fig. \ref{fig:yPiKKplot} we plot the mass normalized Yukawa couplings of Standard Model Higgs with top quarks, Higgs with first KK fermion and of $\Pi_0$ to first KK fermion, for different values of the KK scale. As expected we see that the $c_q$ dependence is mild (i.e. all KK fermions of any flavor will couple with similar strength) and also, as expected, we observe that the mass normalized couplings are quite suppressed with respect to the SM top quark case. Still the multiplicity of KK fermions is high, since there are 6 families of quarks and 3 families of charged leptons (the latter run in the diphoton loop), and for each family there are a few KK levels that give important contributions to the rate. \begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{yPiKKPoint6.pdf}\ \ \includegraphics[height=7cm]{yPiKKPoint5.pdf} \includegraphics[height=7cm]{yPiKKPoint4.pdf}\ \ \includegraphics[height=7cm]{yPiKKPoint2.pdf} \end{center} \vspace{-.2cm} \caption{Yukawa couplings between the lightest KK fermion and the $125$ GeV Higgs (middle curves) and the $750$ GeV CP-odd $\Pi_0$ (lower curves), divided by the KK fermion mass, for different values of the lightest KK gluon mass $M_{KKglu_1}$, as indicated on the panels. This mass normalized Yukawa coupling gives an estimate of the relative contribution of the respective KK fermion to the radiative coupling of the scalar to gluons and photons, to be compared with the mass normalized SM coupling of Higgs to top quarks (shown as a dashed line).} \label{fig:yPiKKplot} \vspace{-.3 cm} \end{figure} A numerical scan of the couplings, including all families and 3 full KK levels is computationally too intensive, so in order to produce the couplings plotted in Fig. \ref{fig:yPiKKplot} we performed an approximation, sufficient for the purposes of the graph. The KK fermion Yukawa couplings plotted neglect mixings between different KK levels and different flavors, and with the zero mode fermions. They are obtained as follows. Consider the $2\times2$ KK mass matrix \bea \left(\begin{matrix} Q_L(x)\ U_L(x) \end{matrix}\right)\ M_u\ \left(\begin{matrix} Q_R(x)\\ U_R(x) \end{matrix}\right) , \eea where $Q_L(x)$ and $Q_R(x)$ represent here a single flavor and a single KK level of the vector-like KK up-type doublets, and similarly for $U_L(x)$ and $U_R(x)$, vector-like KK up-type singlets. The mass matrix is thus \bea M_u= \left(\begin{matrix} m_Q & Y_1 \\ Y_2 & m_U \end{matrix}\right), \label{QUmatrix} \eea where the diagonal entries are the KK masses (large) whereas the off-diagonal entries are coming from Yukawa couplings and are therefore smaller. In order to give a simple estimate, we take for simplicity the fermion bulk mass parameters as $c_q=-c_u$, and the bulk Higgs Yukawa $Y^{5D}$ to be real, which leads to $Y_1=Y_2$ and $m_Q=m_U=m_{KK}$, with the masses and profiles obtained by solving Eq. (\ref{fermionprofile}). With the KK fermion profiles one obtains the off-diagonal entries \bea Y_1= \frac{(Y^{5D}_u)}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{ v(y)}{\sqrt{2}} Q_{L}(y)U_{R}(y) . \eea The matrix that diagonalizes (\ref{QUmatrix}) in this simple limit ($c_q=-c_u$) is $\left(\begin{matrix} -\frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} \\ \frac{1}{\sqrt{2} }& \frac{1}{\sqrt{2}} \end{matrix}\right)$ and the eigenvalues are $m_1=m_{KK}-Y_1$ and $m_2=m_{KK}+Y_1$. Now we apply this rotation to the CP-odd Yukawa coupling matrix \bea Y_\Pi= \left(\begin{matrix} {\cal O}(g) & Y_\pi \\ Y_\pi & {\cal O}(g) \end{matrix}\right), \eea where \bea Y_\pi& =& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q_{L}(y)U_{R}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)}, \eea and where for simplicity we have neglected gauge couplings terms compared to IR Yukawa terms (safe assumption when $Y^{5D}$ is large). After diagonalization, we obtain the two physical couplings between $\Pi_0$ and the KK fermions. When we normalize the couplings by the two eigenmasses and add the two contributions,\footnote{One needs to add the two contributions since there is a cancellation happening level by level.} we obtain \bea \sum_{i=1}^2 \frac{y_i}{m_i} =\frac{y_1^{\pi}}{m_1}+\frac{y_2^{\pi}}{m_2} = -2 \frac{Y_1 Y_\pi}{m_{KK}^2-Y_1^2} \simeq -2\frac{Y_1 Y_\pi}{m_{KK}^2}. \eea The last expression corresponds to the {\it mass normalized} Yukawa couplings of $\Pi_0$ plotted in Fig. \ref{fig:yPiKKplot}, and this describes very closely the behavior of the couplings obtained in the full flavor calculation. The parametric dependence of these couplings is $Y_{5D}^2 v/m_{KK}^2$, so that if $Y^{5D}\sim 3$ we expect mass normalized couplings of order $(10^{-3}-10^{-4})$ GeV$^{-1}$, if the overlap integral is of ${\cal O}(1)$. Since all the profiles of the integral are IR localized, one expects that integral to be ${\cal O}(1)$, although the precise numerical result varies between $0.5$ and $0.05$, depending on the values of the $c_q$ parameter, as shown in the plots. All in all it seems likely that after taking into consideration all the fermion flavors, and for a KK scale of order $1-2$ TeV, the overall KK fermion contribution to the radiative couplings of $\Pi_0(x)$ to photons and gluons can be close to the top quark contribution to the gluon and photon couplings of the Higgs in the SM model. \subsection{$\Pi_0Zh$ coupling} \label{subsec:zh} The coupling between the CP-odd scalar, the $Z$ boson and the Higgs will be extracted from the kinetic operator of the 5D Higgs, \bea \int d^4x dy\ e^{-2\sigma} D_\mu H^\dagger D^\mu H \left(1+ \delta(y-y_i) \frac{d_i}{k}\right). \eea Expanding the SM-like Higgs mode using Eq. (\ref{Hexpansion}) as well as the SM-like $Z_\mu$ and the $750$ GeV $\Pi_0$ using Eqs. (\ref{Aexpansion}) and (\ref{Piexpansion}), we can obtain the coefficient $g_{\Pi hZ}$ of the operator $Z^\mu(x)\Big(h(x)\partial_\mu\Pi_0(x)+\Pi_0(x)\partial_\mu h(x)\Big)$ \bea g_{\Pi hZ} = \frac{g_5^2}{4c_W^2} \int dy e^{-2\sigma} v(y)h(y) f_z(y) \frac{X'(y)}{M^2_z(y) m^2_{\pi}} \left(1+ \delta(y-y_i) \frac{d_i}{k}\right) \eea Now since $M_z(y)=\frac{g_5}{2c_W}e^{-\sigma} v(y)$, $h(y)\sim v(y)/v_4$, and $f_z\simeq 1/\sqrt{y_1}$ we can write \bea g_{\Pi hZ} \simeq \frac{1}{\sqrt{y_1}v_4 m^2_\pi} \big(X(y_1)+ X'(y_1)\frac{d_1}{k}\big)=0, \eea where we have used the boundary conditions for the profile $X(y)$ (see Eq.(\ref{Xbc})) and assumed no UV brane kinetic term ($d_0=0$). The coupling should thus vanish in the limit of flat $Z$ boson profile $f_z(y)$, and when the nontrivial Higgs VEV $v(y)$ is proportional to the Higgs scalar profile $h(y)$. Corrections to these limits scale as $v_4^2/m_{KK}^2$ and $m_h^2/m_{KK}^2$ in the RS case, and so we expect the overall coupling to be highly suppressed. The partial width for the decay $\Pi_0 \to h Z$ is \cite{djouadi1996} \bea \Gamma(\Pi_0 \to h Z)= \frac{g^2_{\Pi hZ}}{16\pi} \frac{m_Z^2}{m_{\pi}} \sqrt{\lambda(m^2_h,m^2_Z;m^2_\pi)}\lambda(m^2_h,m^2_\pi;m^2_Z) \eea where $m_Z$, $m_\pi$ and $m_h$ are the masses of the particles involved and where $\lambda(x,y;z) = (1 - x/z - y/z)^2 - 4 x y/z^2$. With the masses $m_Z=91$ GeV, $m_\pi=750$ GeV and $m_h=125$ GeV, the width becomes $\Gamma(\Pi_0 \to h Z) \sim (900\ g^2_{\Pi hZ})$ GeV. For example we compute numerically $g_{\Pi hZ}$ for three different values of $M_{KKglu_1}$, and with $m_{\Pi_0}=750$ GeV and find \begin{center} \begin{tabular}{|c||c|c|c|c|} \hline $M_{KKglu_1}$ &\ 1000 GeV\ &\ 1400 GeV\ &\ 2400 GeV\ \\ \hline $g_{\Pi hZ}$ & $3.8\times 10^{-4}$ & $3.1\times 10^{-3}$ &$1.1\times 10^{-2}$ \\ \hline $\Gamma(\Pi_0 \to hZ$) &$1.3\times 10^{-4}$ GeV &$8.8\times 10^{-3}$ GeV &$0.11$ GeV \\ \hline \end{tabular} \end{center} Note that the couplings and widths are small but we observe that the partial width becomes larger as the KK mass scale is increased. Since we need the partial width $\Gamma(\Pi_0 \to \gamma \gamma)$ to be similar to that of a $750$ GeV Higgs, i.e. $\Gamma(h_{750GeV} \to \gamma \gamma) \sim 10^{-5}$ GeV, we expect that at $M_{KKglu_1} =1400$ GeV the $Zh$ signal should start putting too much pressure on the allowed parameter space. This is confirmed in the full numerical analysis presented in the next section. \subsection{Estimates and numerical results} \label{subsec:numerical} With all the previous ingredients one can estimate the viability of this scenario in terms of the possible diphoton excess. Let's choose the KK scale such that $M_{KKglu_1}=1400$ GeV; when the bulk mass parameters of the top quark are around $|c_{u_3}|\sim 0.35$ we know that the top quark will have highly suppressed couplings to $\Pi_0$, as shown in the third panel of Fig. \ref{fig:yPiplot} . At the same time, the couplings of the KK tops (as well as all other KK quarks) will have relatively strong Yukawa couplings to $\Pi_0$ (third panel of Fig. \ref{fig:yPiKKplot}), so that the contribution of each of them to the radiative coupling of $\Pi_0$ to gluons is about an order of magnitude smaller than the top contribution to the $h-glu-glu$ coupling of the SM. Thus we could estimate that the overall contribution of all flavors and KK excitations can make up for the suppressed coupling, so that the production cross section of $\Pi_0$ is similar to that of a $750$ GeV SM-like Higgs. The production cross section of a $750$ GeV Higgs through gluon fusion, at the LHC running at 13 TeV is $497$ fb \cite{Dittmaier:2011ti}, so, roughly, this could be assumed for the $\Pi_0$ production cross section. Since the $\Pi_0$ decays to top quarks and $hZ$ are suppressed in this parameters space point, and its decays to $WW$ and $ZZ$ can only be radiative via the CP-odd gauge boson kinetic operator, the main decay channel is into gluons so that the branching of the diphoton channel should be very roughly \bea Br(\Pi_0 \to \gamma\gamma) \sim \frac{\alpha^2_{em}}{8 \alpha^2_s} \frac{N_\gamma}{N_{glu}} \eea where $N_\gamma$ and $N_{glu}$ are the multiplicities of states running in the $(\Pi_0 \gamma \gamma)$ loop and in the $(\Pi_0\ glu\ glu)$ loop respectively. In the diphoton loop there are $3$ extra families of charged lepton KK excitations making the multiplicity of states greater. If their multiplicity and their Yukawa couplings can partially make up for the color factor of 8, then the diphoton cross section might become of ${\cal O}$ (1 fb), as hinted by the December 2015 LHC data. To complete the analysis, we perform a full numerical computation of production and branching ratios in a setup where we consider an effective 4D scenario including three full KK levels for all fields, i.e., we consider $21 \times 21$ fermion mass matrices, which we diagonalize in order to obtain the physical Yukawa couplings. We choose a set of $c$-parameters and 5D Yukawa entries such that the SM masses and mixings are reproduced; the specific flavor choice for these parameters should not affect much the overall results since these depend on overlap integrals between IR localized fields, with very loose $c$-dependence. We choose the background metric parameters so that precision electroweak bounds are kept at bay, i.e. $\nu=0.5$ and $y_s=1.04 y_1$. Two average 5D Yukawa scales are considered, $Y_{5D} \simeq 3$ and $Y_{5D}\simeq 2$, to show the dependence on this parameter, and we also consider two different KK mass scales, $M_{KKglu_1}=1000$ GeV and $M_{KKglu_1}=1300$ GeV, which turn out to lead to successful signal generation. In order to see how tuned is the choice of top $c$-parameter, we plot the production cross section of the CP odd resonance, followed by decays into $\gamma \gamma$, $t\bar{t}$ and $Zh$, as functions of $c_{u_3}$ (the bulk mass parameter of the 5D singlet top quark), with the doublet bulk mass parameter fixed at $c_{q_3}=0.4$. (This value ensures typically suppressed bounds from $Zb_L{\bar b}_L$ bounds \cite{Cabrer}). \begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{prodplotPoint23}\ \ \includegraphics[height=7cm]{prodplotPoint22} \end{center} \vspace{-.2cm} \caption{Production cross sections at the LHC running at 13 TeV in the channels $\gamma\gamma$, $t\bar{t}$ and $hZ$. The lightest KK boson mass is $1000$ GeV and the 5D Yukawa couplings are $Y_{5D}\simeq 3$ in the left panel and $Y_{5D}\simeq 2$ in the right panel. The signal can always be produced at this KK scale but for larger 5D Yukawa couplings there are either too many top pairs or too many dijets produced. Overall we can conclude that for $M_{KKglu_1} = 1000$ GeV and when $1\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} Y_{5D}\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$ and $0.2\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} |c_{u_3}| \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.5$ the signal is easily reproduced.} \label{fig:prodplot1} \vspace{0. cm} \end{figure}\begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{prodplotPoint33}\ \ \includegraphics[height=7cm]{prodplotPoint32} \end{center} \vspace{-.2cm} \caption{Same as the previous figure but for a KK scale of $1300$ GeV. In this case the signal can only be achieved for larger Yukawa couplings, and at a more localized region near $c_{u_3}\sim 0.25$. The bound from $hZ$ now becomes the main constrain. For heavier KK mass scales, the signal can only be produced for much larger Yukawa couplings and the $hZ$ bound becomes too constraining.} \label{fig:prodplot2} \vspace{0. cm} \end{figure} The results are shown in Figs. \ref{fig:prodplot1} and \ref{fig:prodplot2}, where we have chosen two KK scales, $M_{KKglu_1}=1000$ GeV and $M_{KKglu_1}=1300$ GeV. In both cases we show results for $Y_{5D}\simeq 3$ and $Y_{5D}\simeq 2$ to illustrate the sensitivity on this bulk parameter, crucial for enhancing the radiative couplings of $\Pi_0$. When the KK scale is smaller and 5D Yukawa couplings are larger, the production of top pairs and gluon pairs can be too large. By reducing the 5D Yukawa couplings, enough signal can be generated with the dijets and top pairs under control, as well as the $Zh$ decay. The values of $|c_{u_3}|$ must be in the region $0.2\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} |c_{u_3}| \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.5$. For slightly smaller KK scales one expects a similar behavior, but such that 5D Yukawa couplings should be smaller in order to suppress overproduction of $\Pi_0$ particles. This leads to the question of how large can the KK scale be and still manage to produce enough signal. We observe that at $M_{KKglu_1}=1300$ GeV, larger 5D Yukawa couplings ($Y_{5d} \geq 3$) are required in order to have enough signal production. The branching fraction into $hZ$ starts to be problematic, and actually becomes worse for heavier KK scales. The signal production requires that the value of $|c_{u_3}|$ be located around the point where the top Yukawa couplings of $\Pi_0$ are suppressed, in this case around $|c_{u_3}|\sim 0.3$. We conclude therefore that in order to explain the $750$ GeV diphoton, the KK scale must be $M_{KKglu_1} \leq 1300$ GeV, with 5D Yukawa couplings $Y_{5D}\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$. For those values, the signal is quite generic (i.e. small, but typical, $c_{u_3}$ values are required), as long as the $\Pi_0$ mass is set to $750$ GeV with appropriate boundary kinetic terms. \section{Discussion} \label{sec:discussion} We performed an analysis of the scalar sector of warped space models to investigate whether the {\it minimal} model can accommodate a resonance at 750 GeV with the properties observed in the diphoton resonance at the CMS and ATLAS. We show that in the simplest extra dimensional extension of the SM, that is with a 5D Higgs doublet living in the bulk, the lowest pseudoscalar KK excitation can be responsible for the signal observed at 750 GeV. We emphasize that, unlike other explanations relying on scalar fields in warped models, ours does not introduce any new fields or representations, but relies on Higgs brane kinetic terms to lower the KK mass of the lightest CP odd Higgs resonance. This makes the model extremely constrained, with the only new parameter being the IR brane kinetic coefficient $d_{1}$. The lightest CP odd excitation, a mixture of the 5D Higgs field and $Z_5$, does not decay at tree level into $WW$ or $ZZ$, and, over a range of the parameter space, can have suppressed couplings to the top quark, thus a small decay width into $t {\bar t}$. The production through gluon fusion can be loop-enhanced through the effects of the usual KK fermion modes, and so can the diphoton decay. The coupling to $Zh$ is also suppressed although starts increasing dangerously for KK masses above $1500$ GeV. We also showed that in $AdS_5$ spaces (RS-type models) with a KK scale as low as $1500$ GeV, the lowest CP-odd scalar cannot have a mass at 750 GeV, but if the metric differs slightly from the $AdS_5$ metric (generalized warped metric spaces), the Higgs brane kinetic terms can produce a viable CP-odd scalar of mass 750 GeV. Within these modified metric scenarios, and for KK mass scales at around 1 TeV (consistent with precision electroweak bounds) this CP odd resonance obeys the (current) experimental constraints. We analyzed its production and decay for several values of the lowest KK gluon mass ($M_{KKglu_1}$), and show that consistency with the data requires $M_{KKglu_1} \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 1300$ GeV. Our analysis is quite general, and should be valid even in the absence of a signal at 750 GeV. The general conclusion to be taken from our analysis here is that warped space models, without any new particles, can explain a (relatively) light diphoton resonance at the LHC. Should a diphoton excess be found at higher values, even the RS model might accommodate such a state without the need modify the metric. Among the general features of the light CP-odd scalar resonance, resulting from the 5D Higgs doublet are the fact that its coupling to top pairs can be suppressed for appropriate top bulk mass parameters. Also its coupling to $Zh$ is generically suppressed due to the boundary conditions of the CP odd state. In addition, the model predicts that the spectrum for the CP odd and the charged scalars is essentially the same since their differential equations and boundary conditions are almost identical. This means that the lightest charged Higgs boson is expected to have a mass very close to the pseudoscalar mass, so about 750 GeV, in the scenario in which the latter is the diphoton resonance. If the diphoton resonance ends up being a real particle, and not a statistical fluctuation, the prediction from this scenario is that $ZZ$ and $WW$ should not be seen, whereas top pair production, dijet production and $Zh$ signal should be around the corner. Moreover a search for the charged scalars at around $750$ GeV might prove useful in order to disentangle the different models explaining the resonance. \section{Acknowledgments} MF and MT thank NSERC and FRQNT for partial financial support under grant numbers SAP105354 and PRCC-191578. \section{Appendix} In this section we consider the effect of brane localized kinetic terms associated with the 5D Higgs doublet and also with the gauge bosons. For simplicity, let's consider a 5D toy model with a Higgs scalar $H(x,y)$ charged under a local $U(1)$, defined by the following action: \bea S&=&\int d^4xdy \sqrt{g}\ \left(-\frac{1}{4} F^2_{MN} + |D^M H|^2 - V(H) \right) \\ &&\ \ \ + \sum_i \int d^4xdy \sqrt{g}\ \delta(y-y_i) \left(\frac{1}{4} r_i F^2_{MN} + d_i |D^M H|^2 - \lambda_i(H))\right) \label{5Daction} \eea where $F_{MN} = \partial _M A_N - \partial_N A_M$ and for simplicity we set gauge coupling constant to unity in the appropriate mass dimensions. The background spacetime metric is assumed to take the form \bea ds^2=e^{-2\sigma(y)}\eta_{\mu\nu} dx^\mu dx^\nu - dy^2 \eea where $\sigma(y)$ is the warp factor. We are interested in studying the effective 4D perturbative spectrum of the 5D Higgs field and the 5D gauge boson, around a nontrivial Higgs vacuum profile solution $<H>=v(y)$. \bea H(x,y)= \frac{1}{\sqrt{2}}\left(v(y) + h(x,y)\right) e^{i \pi(x,y)} \eea In particular we are interested in the CP-odd Higgs perturbations $\pi(x,y)$, whose equations of motion are coupled with the gauge boson perturbations. The equations read \bea && (1+r_i \delta_i) \partial_\mu \partial^\mu A_\mu - \left( (1+r_i \delta_i)e^{-2\sigma} A'_\mu\right)' + (1+d_i \delta_i) M_A^2 A_\mu +\non\\ &&\hspace{3cm}\partial_\mu \left((1+d_i \delta_i) M_A^2 \pi - (1+r_i \delta_i)\partial^\nu A_\nu - \left((1+r_i \delta_i)e^{-2\sigma} A_5\right)'\right) = 0\\ &&(1+r_i \delta_i) \partial_\mu \partial^\mu A_5- (1+r_i \delta_i) \partial^\nu A'_\nu+ (1+d_i \delta_i)M_A^2 (\pi'-A_5)=0\\ &&(1+d_i \delta_i) \partial_\mu \partial^\mu \pi - (1+d_i \delta_i) \partial^\nu A_\nu + M_A^{-2}\left((1+d_i \delta_i) M_A^2 e^{-2\sigma} (\pi'-A_5)\right)'= 0 \eea where $M_A= v(y) e^{-\sigma}$ and where $\displaystyle d_i \delta_i \equiv \sum_i d_i \delta(y-y_i) $ and $\displaystyle r_i \delta_i \equiv \sum_i r_i \delta(y-y_i) $. We fix partially the 5D gauge by imposing \bea (1+d_i \delta_i) M_A^2 \pi - (1+r_i \delta_i)\partial^\nu A_\nu - \left((1+r_i \delta_i)e^{-2\sigma} A_5\right)' = 0. \label{gaugefix} \eea The previous gauge fixing equation reads in the bulk: \bea M_A^2 \pi - \partial^\nu A_\nu - \left(e^{-2\sigma} A_5\right)' = 0. \label{bulkgaugefix} \eea Note that if we evaluate the bulk constrain Eq.~(\ref{bulkgaugefix}) at $y=y_1-\epsilon$ (i.e. right before the IR brane), we obtain: \bea \partial^\nu A_\nu \Big|_{y_1-\epsilon} = M_A^2 \pi - \left(e^{-2\sigma} A_5\right)'\Big|_{y_1-\epsilon}. \label{epsiloncondition} \eea On the other hand, the effect of the delta functions in Eq.(\ref{gaugefix}) is to produce a discontinuity in the 5D field $A_5$ at the brane location as \bea d_1 M_A^2 \pi -r_1 \partial^\nu A_\nu\Big|_{y_1-\epsilon} = \left[e^{-2\sigma} A_5\right]_{y_1-\epsilon}^{y_1}, \eea and similarly for the UV brane. We can thus multiply (\ref{epsiloncondition}) by $r_1$ and use it in the previous equation, and find the necessary boundary condition between $\pi$ and $A_5$, which ensures that $A_\mu$ is completely decoupled, even on the brane. We find \bea (d_1-r_1) M^2_A \pi + r_1 \left(e^{2\sigma} A_5\right)' \left. \vphantom{\int^A_A} \right|_{y_1-\epsilon} =- \left.e^{-2\sigma} A_5\right|_{y_1-\epsilon} \label{boundaryconditions} \eea where we have taken $A_5$ to vanish exactly on the brane, but it jumps right before the boundary. Inserting the gauge choice in the coupled equations of motion, one manages to decouple the gauge modes $A_\mu$ (in both the bulk and the branes) with a bulk equation \bea \partial_\mu \partial^\mu A_\mu - \left(e^{-2\sigma} A'_\mu\right)' + M_{A}^2 A_\mu =0 \eea and jump condition on $A'_\mu$ \bea r_1 \partial_\mu \partial^\mu A_\mu + d_1 M_{A}^2 A_\mu \Big|_{y_1-\epsilon}=- \left. e^{-2\sigma} A'_\mu\right|_{y_1-\epsilon} \eea where $A'_\mu$ again vanishes exactly on the brane, but has a jump right before it. We separate variables \bea A_\mu(x,y)= V^{4d}_\mu(x) V_y(y) \eea and find the separated equations for the gauge boson tower become \bea && \partial_\mu \partial^\mu V^{4d}_\mu(x)+m^2_A V^{4d}_\mu(x)=0\\ && \left( e^{-2\sigma} V'_y\right)' + ( m^2_A - M_A^2 ) V_y =0 \eea with jump conditions on $V'_y$ \bea (d_i M_{A}^2- r_i m^2_A) V_y \Big|_{y_1-\epsilon}= -\left. e^{-2\sigma} V'_y\right|_{y_1-\epsilon} \eea where the 4D effective mass $m^2_A$ is the constant of separation of variables. The remaining equations are, in the bulk, \bea && \partial_\mu \partial^\mu A_5+ M_A^2 (\pi'-A_5) - \left(M_A^2 \pi\right)' +\left( \left(e^{-2\sigma} A_5\right)' \right)' =0\\ && \partial_\mu \partial^\mu \pi + M_A^{-2} \left( M_A^2 e^{-2\sigma} (\pi'-A_5) \right)' - M_A^2 \pi + \left(e^{-2\sigma} A_5\right)'= 0\label{a5pimixedequation} \eea and the fields must verify the boundary conditions of Eq.(\ref{boundaryconditions}). We now perform a mixed separation of variables: \bea A_5(x,y)&=& G(x) g(y) + \pi_x(x) \eta(y) \label{expansion1}\\ \pi(x,y)&=& G(x) h(y) + \pi_x(x) \xi(y)\label{expansion2} \eea which is to say that both $A_5(x,y)$ and $\pi(x,y)$ contain each some Goldstone and CP-odd degree of freedom. The profiles $g(y)$, $\eta(y)$, $h(y)$ and $\xi(y)$ quantify how much of each they contain. Of course the functions $g$ and $h$ are related to each other, as well as $\eta$ and $\xi$. The relationships are such that $G(x)$ and $\pi_x(x)$ decouple. With the choice \bea h(y)&=& \frac{K(y)}{m^2_G}\\ g(y)&=&\frac{K'(y)}{m^2_G}\\ \eta(y)&=&\frac{e^{2\sigma}}{m^2_\pi}X(y)\\ \xi(y)&=&\frac{1}{m^2_\pi M^2_A}X'(y) \eea and using the mixed separation of variables in (\ref{expansion1}) and (\ref{expansion2}), the mixed equations of motion in (\ref{a5pimixedequation}) decouple and we obtain \bea && \frac{e^{2\sigma}}{M_A^2} X(y) \partial_\mu \partial^\mu \pi_x(x) +\pi_x(x) e^{2\sigma} X(y) - \pi_x(x) \left[M_A^{-2} X'(y)\right]'= 0\\ &&K(y) \partial_\mu\partial^\mu G(x)+ G(x) \left[ \left(K'e^{-2\sigma}\right)' + M_A^2 K(y)\right] = 0 . \eea Once separated, we obtain, for the CP odd physical scalars \bea &&\partial_\mu \partial^\mu \pi_x(x) + m_\pi^2 \pi_x(x) =0\\ &&\left(M_A^{-2} X'\right)'+ e^{2\sigma} \left(\frac{m_\pi^2}{M_A^2}-1\right) X=0 , \eea with boundary conditions \bea d_i X' = -X \eea and for the Goldstone modes, \bea &&\partial_\mu\partial^\mu G(x)+ m_G^2 G(x)=0\\ &&\left(K'e^{-2\sigma}\right)'+ (m_G^2-M_A^2) K=0, \eea with boundary conditions \bea (d_i M_A^2-r_i m^2_G ) K = -e^{-2\sigma} K' . \eea Note that both the equations and boundary conditions for the Goldstone bosons are identical to the ones for the gauge boson tower, as should be, so that they can then be gauged away level by level with the remaining gauge fixing freedom. \section{Introduction} The original motivation for warped extra-dimensions was to address the hierarchy problem, so that the fundamental scale of gravity is exponentially reduced along the extra dimension, from the Planck mass scale to the TeV scale. Thus, the TeV scale becomes the natural scale of the Higgs sector if this one is localized near the TeV boundary of the extra dimension, as first introduced by Randall and Sundrum (RS) \cite{RS1}. If SM fields are allowed to propagate in the extra dimension \cite{Davoudiasl:1999tf}, the scenario can also address the flavor puzzle of the SM, explaining fermion masses and mixings from the geographical location of fields along the extra dimension. However, processes mediated by the heavy resonances of the 5D bulk fields, Kaluza-Klein (KK) modes, generate dangerous contributions to electroweak and flavor observables (including dangerous deviations to the $Zb\bar{b}$ coupling) \cite{Burdman,RSeff,Agashe}, pushing the KK mass scale to $5-10$ TeV \cite{Weiler}. A popular mechanism to lower the KK scale involves using a custodial gauge $SU(2)_R$ symmetry \cite{custodial}, which ensures a small contribution to electroweak precision parameters, lowering the KK scale bound to about 3 TeV. Alternatively, one can study scenarios in which the metric is slightly modified from the RS metric background ($AdS_5$). This can be achieved quite naturally from the backreaction on the metric caused by a 5D scalar field stabilizing the original $AdS_5$ warped background \cite{Goldberger:1999uk}. When the 5D Higgs is sufficiently leaking into the bulk and when the metric background is modified near the TeV boundary, the scenario allows for KK scales as low as 1-2 TeV, with precision electroweak and flavor constraints under control \cite{Cabrer}. An inconvenience is that these scenarios are typically hard to probe experimentally as the couplings of all particles are very suppressed \cite{Falkowski:2008fz,deBlas:2012qf}. Still, it has been shown that it can still lead to interesting deviations in Higgs phenomenology, as the Higgs couplings can receive sufficient radiative corrections from the many KK fermions of the model \cite{Frank:2015zwd}. Run 2 LHC data at $\sqrt{s}=13$ TeV shows signals of a new resonance in the diphoton distribution at an invariant mass of 750 GeV with a 3.9 significance at ATLAS \cite{ATLAS750}, with 3.2 fb$^{-1}$ and 3.4 combined significance at CMS \cite{CMS:2016owr}, with 2.6 fb$^{-1}$(combining run 1 and run 2 results). ATLAS reports 14 events and CMS, 10. The experimental data is summarized in Table \ref{tab:siglims}. \begin{table} \begin{center} \begin{tabular}{|c|c|c|} \hline\hline Channel & $8~{\rm TeV}, \sigma_{max}$ & $13~{\rm TeV}, \sigma_{max}$\\ \hline\hline $\gamma\gamma$ &$.21\pm.22~{\rm fb}$ \cite{ATLAS750} & $5.5\pm 1.5~{\rm fb}$ \cite{ATLAS750} \\ & $.63\pm .31~{\rm fb}$ \cite{CMS:2016owr} &$4.8\pm 2.1~{\rm fb}$ \cite{CMS:2016owr} \\ \hline \hline \hline $t {\overline t}$ & $700~{\rm fb}$ \cite{Aad:2015fna} & $\sim 2000~{\rm fb}$ \\ \hline $jj$ & $2.1~{\rm pb}$ \cite{Khachatryan:2016ecr}& $\sim 10~{\rm pb}$ \\ \hline $hZ$ & $19~{\rm fb}$ \cite{Aad:2015wra} & $\sim 100~{\rm fb}$ \\ \hline\hline \end{tabular} \caption{\label{tab:siglims} LHC diphoton cross sections at ATLAS and CMS and relevant signal limits for the $750$ GeV CP odd scalar resonance considered here. At 13 TeV, these are the rough implied bounds from the 8 TeV limits.} \end{center} \end{table} In light of all this, we propose here a simple and economic explanation within warped extra-dimensional models. It would require the presence of a 5D bulk Higgs, and because the mass of the new resonance is 750 GeV, the Higgs should be as much delocalized as possible from the TeV brane (but still close enough to address the hierarchy problem). The reason is that the masses of the Higgs KK excitations will increase as the Higgs is pushed towards the brane, getting infinitely heavy in the limit of a brane Higgs. Out of these excitations some are CP odd scalars, making them a natural candidate for the signal since they do not couple at tree-level to $ZZ$ or $WW$. We will show that if the typical mass of the KK gluon (typically the lightest and most visible KK particle) is around 1-2 TeV, it is very simple to obtain a 750 GeV CP odd Higgs with the help of small (and positive) brane localized kinetic terms of the 5D Higgs. Since the CP odd scalars do not couple at tree-level to $ZZ$ or $WW$, the largest coupling is going to be to pairs of tops. As will be shown, this coupling can be naturally small in wide regions of the allowed parameter space. This way, the radiative coupling to gluons, large enough for producing CP odd scalars, can also dominate the decays and the (also) radiative decay into photons can then receive enough branching fraction. Explanations of the 750 GeV signal within warped scenarios have been put forward previously, with the resonance interpreted as a radion \cite{Ahmed:2015uqt}, (and/or dilaton \cite{Cao:2016udb}), as a KK graviton \cite{Falkowski:2016glr,Hewett:2016omf,Carmona:2016jhr}, a 5D field-related axion \cite{Chakrabarty:2016hxi} or as an additional 5D singlet scalar added to the model \cite{Bauer:2016lbe}. The explanation proposed here, while preserving minimality and agreement with the diphoton excess, is also satisfied naturally in a significant region of the parameter space. We proceed as follows. In Sec. \ref{sec:background} we describe briefly the warped scenario, followed by its Higgs and gauge sector in Sec. \ref{sec:higgsgauge}, and of the CP-odd sector in more detail in Sec \ref{sec:cpodd}. Within that section we look at the fermion couplings in \ref{subsec:fermion}, the $\gamma \gamma$ and $glu-glu$ couplings in \ref{subsec:gggamgam} and to $Zh$ couplings in \ref{subsec:zh}. Our numerical estimates are presented in \ref{subsec:numerical} and we conclude in Sec. \ref{sec:discussion}. We leave some of the details for the Appendix. \section{The background metric} \label{sec:background} The (stable) static spacetime background is: \bea ds^2 = e^{-2\sigma(y)}\eta_{\mu\nu} dx^\mu dx^\nu - dy^2, \label{RS} \eea where the extra coordinate $y$ ranges between the two boundaries at $y=0$ and $y=y_{1}$, and where $\sigma(y)$ is the warp factor responsible for exponentially suppressing mass scales at different slices of the extra dimension. In the original RS scenario, $\sigma(y)=ky$, with $k$ the curvature scale of the $AdS_5$ interval that we take of the same order as $M_{Pl}$. Nevertheless this configuration is not stable as it contains a massless radion, a result of having the length of the interval not fixed. In more general warped scenarios with stabilization mechanism, $\sigma(y)$ is a more general (growing) function of $y$. We consider here the specific case where a 5D bulk stabilizer field backreacts on the $AdS_5$ metric producing the warp factor \cite{Cabrer,Carmona}. \bea\label{genericmetric} \sigma(y) = ky - {1\over \nu^2} \text{log}\left(1-{y\over y_s}\right), \eea where $y = y_s$ is the position of a metric singularity, which stays beyond the physical interval considered here, i.e. $y_s> y_{1}$. In these modified metric scenarios, the Planck-TeV hierarchy is reproduced with a shorter extra-dimensional length due to a stronger warping near the TeV boundary, so that whereas in RS we have $k y_{1}\simeq 35$, in the modified scenarios we can have $k y_{1}\simeq 20-30$. The appeal of this particular modification lies on the possibility of allowing for light KK particles ($\sim 1$ TeV), while keeping flavor and precision electroweak bounds at bay. This happens when the Higgs profile leaks sufficiently out of the TeV brane so that all of its couplings to KK particles are suppressed compared to the usual RS scenario \cite{Cabrer, Falkowski:2008fz,Carmona}. We thus fix the Higgs localization to a point where it is maximally pushed away from the IR brane, while still solving the hierarchy problem (i.e. making sure that we are not reintroducing a new fine-tuning of parameters within the Higgs potential parameters \cite{Cabrer,Quiros:2013yaa}.) \section{Gauge and Higgs sector} \label{sec:higgsgauge} The matter content of the model is that of a minimal 5D extension of the Standard Model, so that we assume the usual strong and electroweak gauge groups $SU(3)_c\times SU(2)_L\times U(1)_Y$, with all fields propagating in the bulk. The fermions of the model are also bulk fields, with different 5D bulk masses, so that their zero mode wavefunctions are localized at different sides of the interval. This way the scenario also addresses the flavor puzzle of the SM, since hierarchical masses and small mixing angles for the SM fermions become a generic feature due to fermion localization and small wavefunction overlaps \cite{Frank:2015sua}. In the electroweak and Higgs sector we consider the following action \bea \hspace{-1cm}&&S=\int d^4xdy \sqrt{g}\ \left(-\frac{1}{4} F^2_{MN} + |D^M H|^2 - V(H) \right) \ \ \ \ \\ \hspace{-1cm} && + \sum_{i=1}^{2}\int d^4xdy \sqrt{g}\ \delta(y-y_i) \left(\frac{d_i}{k} |D^M H|^2 - \lambda_i(H)\right) \label{5Daction} \eea where the capital index $M$ will be used to denote the $5$ spacetime directions, while the Greek index $\mu$ will be used for the 4D directions. The coefficients $d_i$ (in units of $k$) are essentially free parameters encoding the importance of brane localized kinetic terms associated with the bulk Higgs field. These terms will allow for a slight modification of the spectrum of the KK Higgs excitations, particularly useful in reproducing a $750$ GeV CP-odd excitation. These brane kinetic terms can be thought of as exactly localized operators, or as bulk operators that happen to be dynamically localized due to couplings to some localizer VEV\footnote{In order to avoid tachyons and/or ghosts, the sign of the purely brane localized brane kinetic terms will be kept positive, i.e. $d_i>0$.}. The 5D Higgs doublet is expanded around a nontrivial VEV profile $v(y)$ as \bea \label{Hexpansion} H=\frac{1}{\sqrt{2}} e^{i g_5 \Pi}\left(\begin{matrix} 0\\ v(y) +h(x,y) \end{matrix} \right) \eea and the covariant derivative is $D_M= \partial_M + i g_5 A_M$ with \bea \label{Aexpansion} A_M=\left( \begin{matrix} s_W A_M^{em}+\frac{c_W^2-s_W^2}{2c_W}Z_M & \frac{1}{\sqrt{2}}W_M^+\\ \frac{1}{\sqrt{2}}W_M^-& -\frac{1}{2c_W} Z_M \end{matrix} \right) \eea and CP-odd and charged Higgs part is \bea \label{Piexpansion} \Pi=\left( \begin{matrix} \frac{c_W^2-s_W^2}{2c_W}\Pi_z & \frac{1}{\sqrt{2}} \Pi^+\\ \frac{1}{\sqrt{2}}\Pi^-& -\frac{1}{2c_W} \Pi_z \end{matrix} \right) \eea with the weak angle defined like in the SM, i.e. $s_W/c_W=g'_5/g_5$, where $g_5$ and $g_5'$ are the 5D coupling constants of $SU(2)_L$ and $U(1)_Y$. The extraction of degrees of freedom in this context has been performed in \cite{Quiros:2013yaa,Falkowski:2008fz,Archer:2012qa} and we outline here the main results. The effect of brane kinetic terms in the Higgs sector is new and its derivation is outlined in the Appendix. The 5D equations of motion for all these fields are coupled (except for the case of the real Higgs excitation $h(x,y)$) and in order to decouple them, one can partially fix the gauge, or add a gauge fixing term to the previous 5D action. For example, in the CP-odd case, the fields $Z_\mu(x,y)$, $Z_5(x,y)$ and $\Pi_z(x,y)$ must be unmixed. The partial gauge fixing constraint\footnote{There is still be some gauge freedom left, so that the towers of 4D Goldstone bosons that appear can be gauged away.} \bea \partial^\mu Z_\mu - M^2_z(y) \Pi_z +(e^{-2\sigma}Z_5)'=0 \eea manages to decouple the fields $Z_\mu$ from $Z_5$ and $\Pi_z$ in the bulk. We defined here $M_z(y)=\frac{g_5}{2c_W} v(y) e^{-\sigma(y)}$. However, the presence of the Higgs brane kinetic terms, proportional to $d_i$ in the action, forces us to extend the gauge choice on the branes, producing a lifting of $Z_5$ field so that the decoupling is maintained at the boundaries\footnote{In the absence of brane kinetic terms, $Z_5$ must have vanishing boundary conditions (Dirichlet) if $Z_\mu$ is to have Neumann conditions and thus develop a zero mode KK excitation in the effective 4D theory.}. The appropriate boundary condition at the IR brane is \bea Z_5(x,y_1) =- \frac{d_1}{k} M^2_z(y_1) e^{2\sigma(y_1)} \Pi_z(x,y_1) \eea where $y_1$ denotes the position of the boundary (note that if the brane kinetic term parameter $d_1$ tends to zero, the condition on $Z_5$ becomes Dirichlet, as expected). With this type of gauge choice, the 5D fields $Z_\mu$, $W_\mu$ and $A_\mu$ have independent 5D equations of motion. In order to extract the effective 4D degrees of freedom, we expand the gauge fields as $Z_\mu(x,y)=Z^n_\mu(x) f^n_z(y)$, $W_\mu(x,y)=W^n_\mu(x) f^n_w(y)$ and $A_\mu(x,y)=A^n_\mu(x) f^n_\gamma(y)$ (summation over $n$ is understood) and where $Z^0_\mu(x)$, $W^0_\mu(x)$ and $A^0_\mu(x)$ are the $Z$, $W$ and $\gamma$ gauge bosons of the SM. The extradimensional profiles $f^n_z(y)$, $f^n_w(y)$ and $f^n_\gamma(y)$ are solutions of \bea\label{gaugeeq} \left(e^{-2\sigma} f'_a\right)'+(m^2_{n} - M_a^2(y)) f_a =0 \eea where $a=z,w,\gamma$ and where $M_z(y)=\frac{g_5}{2c_W} v(y) e^{-\sigma(y)}$, as defined before, $M_w(y)=\frac{g_5}{2} v(y) e^{-\sigma(y)}$, and $M_\gamma=0$. The boundary conditions for these profiles are\footnote{We ignore here possible brane localized gauge kinetic terms and keep only the effects from Higgs brane kinetic terms. We include everything in the derivation outlined in the Appendix.} \bea\label{gaugebc} \frac{d_i}{k} M_a^2 f_a(y_i) = - e^{-2\sigma}f'_a(y_i). \eea The CP-even Higgs field is expanded as $h(x,y)=h^n(x) h^n_y(y)$ and the equations for the Higgs profiles are, with $h_y \equiv h_y^n(y)$: \bea \left(e^{-4\sigma} h'_y\right)'+(m^2_{h_n} - \mu^2_{bulk}) h_y =0, \eea where $\displaystyle \mu^2_{bulk} = \frac{\partial^2 V }{\partial H^2}\Big|_{H=v}$. The boundary conditions are \bea \left(\mu^2_{brane_i} - \frac{d_i}{k}\ m^2_{h_n} e^{2\sigma}\right) h_y = -e^{-4\sigma} h'_y, \eea with $\displaystyle \mu^2_{brane_i} = \frac{\partial^2 \lambda_i }{\partial H^2}\Big|_{H=v}$. Note that the mode $h^0(x)$ is interpreted as the SM Higgs boson. There are still some degrees of freedom left, and their 5D equations of motion still happen to be mixed. One of the coupled systems involves $Z_5$ and $\Pi_z$ and the other coupled system involves $\Pi^{\pm}$ and $W_5^\pm$. In order to disentangle these systems one must perform a mixed expansion, so that the decoupling of fields will happen KK level by KK level. The mixed expansions are, in the CP-odd sector, \bea\label{Z5decomp} \hspace{-.6cm}Z_5(x,y)&=& G^n(x)\frac{f'_{G_n}(y)}{m^2_{G_n}} + \Pi_n(x)\frac{e^{2\sigma}}{m^2_{\pi_n}}X_\pi(y) \\ \hspace{-.6cm}\Pi_z(x,y)&=& G^n(x)\frac{f_{G_n}(y)}{m^2_{G_n}} + \Pi_n(x) \frac{1}{m^2_{\pi_n}M^2_{z} }X_{\pi}'(y), \label{Pidecomp} \eea and in the charged scalar sector they are \bea \hspace{-.6cm}W_5^\pm(x,y) &=& G_n^\pm(x) \frac{f_{G_n^\pm}(y)}{m^2_{G_n^\pm}} + \Pi^\pm_n(x)\frac{e^{2\sigma}}{m^2_{\pi_n^\pm}}X_\pm(y)\ \ \\ \hspace{-.6cm}\Pi^\pm(x,y)&=& G_n^\pm(x) \frac{f_{G_n^\pm}(y)}{m^2_{G_n^\pm}} +\Pi^\pm_n(x) \frac{1}{m^2_{\pi_n^\pm}M^2_{w} }X_\pm'(y).\ \ \eea where $M_z(y)$ and $M_w(y)$ were defined below Eq~(\ref{gaugeeq}). The effective 4D physical fields are the tower of CP-odd neutral scalars $\Pi_n(x)$ and the tower of charged scalars $\Pi^\pm_n(x)$. Their associated extra-dimensional profiles $X_\pi(y)$ and $X_\pm(y)$ obey the equations \bea \left(\frac{1}{M_a^2(y)} X_a' \right)'+ \left(\frac{m^2_{\pi_a}}{M_a^2(y)}- 1 \right) e^{2\sigma} X_a=0 \label{Xeq} \eea where $M_a(y)=(M_z(y),M_{w}(y))$ and $X_a=(X_\pi,X_{\pm})$. The boundary conditions are \bea\label{Xbc} \frac{d_i}{k}\ X'_a= -X_a, \eea and note that vanishing Higgs brane kinetic terms implies Dirichlet boundary conditions for $X_a$. We checked that these bulk equations agree with \cite{Falkowski:2008fz,Quiros:2013yaa,Archer:2012qa}, the only new addition being the boundary conditions imposed by the presence of Higgs brane kinetic terms. In order for these 4D scalars to be canonically normalized, we require \bea \frac{1}{m^2_a}\int dy\ e^{2\sigma} \frac{X_a^2}{M_a^2}=1 \eea and this condition includes the effect of Higgs brane kinetic terms. The remaining 4D fields are $G_n(x)$ and $G_n^\pm(x)$, which are Goldstone bosons at each KK level. The profile wavefunctions $f_{G_a}(y)$ obey the same differential equations as the gauge profiles, Eq.(\ref{gaugeeq}), as well as the same boundary conditions, Eq.(\ref{gaugebc}). The spectrum is thus identical to the gauge bosons spectrum level by level. These fields appear in the effective 4D action coupled to ($\partial^\mu Z^n_\mu$) or ($\partial^\mu W^n_\mu$), and of course there is a leftover gauge freedom allowing us to gauge them away (i.e., they are pure gauge). We wish to identify the lightest CP-odd scalar $\Pi_0(x)$ with the observed diphoton peak at the LHC, so that we need to fix its mass $m_{\Pi_0}= 750$~GeV. In order to have an idea of the effects of the Higgs brane kinetic terms on the CP-odd scalar spectrum, we consider two different parameter points, one in which the RS background metric is recovered with $\nu=10$ and $y_s=4\times y_{1}$ ($\nu$ is the exponent appearing in the modified metric and if relatively large, the location of spurious singularity is sent away from the boundary, recovering essentially the $AdS_5$ metric). The other case is the situation where the metric modification allows for TeV size KK masses, which are safe from precision electroweak constrains. The parameters chosen there are $\nu=0.5$ and $y_s=1.04\times y_{1}$. In both parameter points, we fix the KK mass of the first gluon excitation to be $1500$ GeV\footnote{Of course, this RS point is presented for comparison only, since such light KK masses will produce too large deviations in the precision electroweak observables.}. \begin{figure}[t] \center \begin{center} \includegraphics[height=9.5cm]{XmassplotRS.pdf}\ \ \includegraphics[height=9.5cm]{Xmassplot.pdf} \end{center} \vspace{-.2cm} \caption{Mass spectrum of the first CP-odd Higgs scalars as a function of the brane kinetic term coefficient $d_{1}$ in the RS metric limit ($\nu=10$ and $y_s=4\ y_{1}$; left panel) and within a noticeably IR-modified $AdS_5$ metric ($\nu=0.5$ and $y_s=1.04\ y_{1}$; right panel). In both cases, the first KK gluon mass is fixed at 1500 GeV. One can see that in RS it is not possible to obtain a 750 GeV CP-odd scalar mass, whereas within the modified metric, a brane kinetic coefficient $d_{1}\simeq 0.51$ does produce a mass of 750 GeV, shown here as a dot. } \label{fig:Xmassplot} \vspace{-.2cm} \end{figure} In Fig. \ref{fig:Xmassplot} we show the spectrum of the first 4 KK levels of CP-odd Higgs bosons $\Pi_n(x)$, for $n=0,1,2,3$, as a function of the brane kinetic term $d_{1}$ in units of the curvature $k\sim M_{Pl}$. The effects of the UV localized brane kinetic term are warped suppressed and so we do not consider them here anymore. We can see that in the RS limit, it is not possible to reduce sufficiently the lightest CP-odd mass to 750 GeV, as its mass tends asymptotically from about $2500$ GeV without brane kinetic term to about $1750$ GeV for large brane kinetic terms. On the other hand, with the modified metric it becomes possible to reduce greatly the lightest CP-odd mass with relatively small brane kinetic term coefficients (which in this particular case tends asymthotically to about 400 GeV). Parameter points such that the metric modification lies between the two considered will actually have an intermediate behavior, with a lightest CP-odd mass having increasing asymptotic values as one recovers the RS background. Finally also note that the spectrum for the charged scalars is essentially the same since their differential equations and boundary conditions are identical except for the functions $M_z(y)$ and $M_w(y)$, which differ by about $10\%$. $M_w(y)$ produces a deviation from the CP-odd scalar spectrum of less than $5\%$. This means that the scenario under consideration should also contain a lightest charged Higgs scalar with a mass of about $750$ GeV also. The next question to ask is how big is the effect of the Higgs brane kinetic term on the gauge bosons, and in particular on the lowest ones, i.e. the SM $W$ and $Z$ bosons. These terms represent an additional (brane localized) contribution to the mass of the gauge bosons. In principle their mass is generated here from a bulk Higgs mechanism, unless the brane kinetic terms are overly important (not the limit we are working with here). We can quickly estimate its effect on the lowest lying gauge fields. These are essentially flat (like all gauge zero modes) and thus their wave function is $f^0_z\sim 1/\sqrt{y_{1}}$. The contribution of a brane localized mass squared term is $\delta m^2_z \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} d_{1}\ v^2/y_{1}\sim d_{1}\times 700$ GeV$^2 $. For IR brane kinetic term coefficient $d_{1}$ of ${\cal O}(1)$, this represents naively at most some $10\% $ contribution to the overall mass squared of either $W$ or $Z$. In the particular case of the modified metric with a brane coefficient $d_{1}\simeq 0.51$, and metric parameters $\nu =0.5$ and $y_s=1.04 \times y_1$ (which produces a light CP-odd scalar of $750$ GeV), the exact numerical effect on the zero mode gauge boson masses ($W$ and $Z$) is a shift of $3$ GeV with respect to the no-brane-kinetic-term limit. Of course, in the presence of brane kinetic terms, one redefines the VEV normalization constant, and the value of $g_5$, in order to correctly account for the SM gauge boson masses and electroweak couplings. \section{CP-odd Higgs couplings} \label{sec:cpodd} As a $750$ GeV CP-odd Higgs scalar $\Pi_0(x)$ is allowed in the spectrum, thanks to the effect of small brane localized Higgs kinetic terms, we now study its couplings to SM particles in order to see if the observed excess at the LHC can be associated with this excitation. Of course being a CP-odd scalar its tree-level couplings to $ZZ$ and $WW$ are zero, making it an ideal candidate for the observed exotic events. We thus need to focus on its tree-level couplings to fermions (and top quark in particular), to $Zh$ (where $h$ is the $125$ GeV Higgs) and to its radiative couplings to photons and gluons. We study these in the subsequent subsections. \subsection{Fermion couplings} \label{subsec:fermion} The couplings of $\Pi_0(x)$ to fermions arise from two sources in the action. First source comes from the 5D Higgs Yukawa couplings, and second, from the gauge fermion couplings. This is because the physical field $\Pi_0(x)$ contains some of CP-odd Higgs scalar, and some of $Z_5$ excitation, where $Z_5$ is the fifth component of the 5D vector boson $Z_M$. However the 5D Yukawa coupling allows for direct coupling of $\Pi_0(x)$ to two zero-mode fermions, whereas the gauge-fermion coupling allows only couplings between fermion zero-modes and higher KK fermion levels. As we will see, it is important to keep both couplings, since after electroweak symmetry breaking the physical SM fermions (top quarks in particular) are mostly zero-modes but also contain a small amount of higher KK excitations, and could thus inherit some of the original gauge-fermion coupling, especially if the tree-level Yukawa coupling between $\Pi_0(x)$ and zero-mode top quarks is suppressed (as it can be). The relevant terms in the action are the 5D Higgs Yukawa couplings and the fermion gauge interaction term, \bea S_{ff\Pi_0} \subset \int d^4xdy \sqrt{g}\left[ Y_u HQU + Y_d HQD + h.c.+ \bar{Q} {\cal D}\hspace{-3mm}/\ Q + \bar{U} {\cal D}\hspace{-3mm}/\ U + \bar{D} {\cal D}\hspace{-3mm}/\ D \right] \eea where $Q,U,D$ represent the 5D fermion $SU(2)_L$ doublets, up-type and down-type singlets (with generation indices and isospin indices suppressed). The kinetic terms contain the 5D covariant derivative and from them we extract the terms containing the CP-odd component $Z_5(x,y)$, and from the Higgs Yukawa couplings we extract the terms containing the CP-odd Higgs component $\Pi_z(x,y)$. We follow the approach of \cite{Frank:2015zwd,Diaz-Furlong:2016ril} and compute these couplings by considering only the effects of three full KK levels, i.e. computing $21 \times 21$ fermion Yukawa coupling matrices (with 3 $up$ and 3 $down$ families, each containing zero modes and 3 KK levels with an $SU(2)_L$ doublet and a singlet in each level, i.e., 3 zero modes plus $3\times 3 \times 2$ KK modes). Note that we are interested in the couplings of the $750$ GeV CP-odd scalar $\Pi_0(x)$ to SM fermions (top quarks primarily), but we also need its couplings with the rest of KK fermions, since these interactions will be crucial to generate large enough radiative couplings to photons and gluons. We first write the effective 4D up-type quark mass matrix as \bea \left(\begin{matrix} q^0_L(x)\ Q_L(x)\ U_L(x) \end{matrix}\right)\ M_u\ \left(\begin{matrix} u^0_R(x)\\ Q_R(x)\\ U_R(x) \end{matrix}\right) \eea in a basis where $q^0_L(x)$ and $u_R^0(x)$ represent three zero-mode flavors each (doublets and singlets of $SU(2)_L$), and $Q_L(x)$ and $Q_R(x)$ represent three flavors and three KK levels of the vector-like KK up-type doublets, and $U_L(x)$ and $U_R(x)$ represent three flavors and three KK levels of vector-like KK up-type singlets. The mass matrix is thus \bea M_u= \left(\begin{matrix} (y^{0}_{u})_{3\times3} & (0)_{3\times 9} & (Y^{qU})_{3\times 9}\\ (Y^{Qu})_{9 \times3} & (M_Q)_{9\times 9} & (Y_1)_{9 \times 9}\\ (0)_{9 \times3} & (Y_2)_{9\times 9} & (M_U)_{9\times 9} \end{matrix}\right) \label{mumatrix} \eea with the down sector mass matrix ${M}_d$ computed in the same way. The submatrices are obtained by evaluating the overlap integrals \bea\label{YYY1KK} y^0_u= \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{v(y)}{\sqrt{2}} q^{0,i}_{L}(y)u^{0,j}_{R}(y) \\ Y^{qU}= \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{v(y)}{\sqrt{2}} q^{0,i}_{L}(y)U^{n,j}_{R}(y) \\ Y^{Qu} = \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{v(y)}{\sqrt{2}} Q^{m,i}_{L}(y)u^{0,j}_{R}(y)\\ Y_1 = \frac{(Y^{5D}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{ v(y)}{\sqrt{2}} Q^{m,i}_{L}(y)U^{n,j}_{R}(y) \\ Y_2 = \frac{(Y^{5D^*}_u)_{ij}}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{ v(y)}{\sqrt{2}} Q^{m,i}_{R}(y)U^{n,j}_{L}(y)\, , \eea where the indices $m$ and $n$ track the KK level and $i,j=1,2,3$ are 5D flavor indices. The diagonal matrices $(M_Q)_{9\times 9}$ and $(M_U)_{9\times 9}$ are constructed with the masses of all the KK quarks involved. The masses and the profiles of the KK fermions appearing in these overlap integrals ($Q_L(y)$, $Q_R(y)$, $U_L(y)$ and $U_R(y)$) are obtained by solving differential equations for the fermion profiles \bea \partial_y\left(e^{(2c-1)\sigma(y)}\partial_y\left( e^{-(c+2)\sigma(y)}\right)\right)f(y)+e^{(c-1)\sigma(y)} m_n^2 f(y) =0 \label{fermionprofile} \eea where $f(y)$ is the KK profile. The mass eigenvalues $m_n$ are found by imposing Dirichlet boundary conditions on the wrong chirality modes. As mentioned before, we have included 3 full KK levels so that the mass matrices in the gauge basis are $21 \times 21$ dimensional matrices, which are not diagonal. One needs to diagonalize them, and by doing so, move to the quark physical basis where all the fermions couplings can then be extracted. In the CP-odd scalar sector, we can write the effective 4D Yukawa-type couplings to fermions in the same gauge basis as before \bea \left(\begin{matrix} q^0_L(x)\ Q_L(x)\ U_L(x) \end{matrix}\right) {\bf Y}_\pi \left(\begin{matrix} u^0_R(x)\\ Q_R(x)\\ U_R(x) \end{matrix}\right) \ \Pi_0(x) \eea where now the $21 \times 21$ coupling matrix ${\bf Y}_{\pi}$ is given by \bea {\bf Y}_{\pi}= \left(\begin{matrix} (y^{0}_{\pi qu})_{3\times3} & (a_{\pi qQ})_{3\times 9} & (Y_{\pi qU})_{3\times 9}\\ (Y_{\pi Qu})_{9 \times3} & (a_{\pi QQ})_{9\times 9} & (Y_{1}^{\pi})_{9 \times 9}\\ (a_{\pi uU})_{9 \times3} & (Y_{2}^{\pi})_{9\times 9} & (a_{\pi UU})_{9\times 9} \end{matrix}\right) \label{Ypimatrix} \eea The submatrices are obtained by the overlap integrals \bea\label{YYPI} y^{0}_{\pi qu} &=& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} q^{0,i}_{L}(y)u^{0,j}_{R}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y_{\pi qU}&=& i\frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} q^{0,i}_{L}(y)U^{n,j}_{R}(y)\frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y_{\pi Qu}& =& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q^{m,i}_{L}(y)u^{0,j}_{R}(y)\frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y^\pi_1& =& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q^{m,i}_{L}(y)U^{n,j}_{R}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \\ Y^\pi_2& =& i \frac{(Y^{5D^*}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q^{m,i}_{R}(y)U^{n,j}_{L}(y)\frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \, , \eea and \bea a_{\pi qQ}&=& \frac{g^{5D}_{L}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} q^{0,i}_{L}(y)Q^{n,j}_{R}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \\ a_{\pi uU}&=&\frac{g^{5D}_{R}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} u^{0,i}_{R}(y)U^{n,j}_{L}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \\ a_{\pi QQ}&=&\frac{g^{5D}_{L}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} Q^{m,i}_{L}(y)Q^{n,j}_{R}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \\ a_{\pi UU}&=&\frac{g^{5D}_{R}}{\sqrt{k}} \int_0^{y_1} dy e^{-2\sigma(y)} U^{m,i}_{R}(y)U^{n,j}_{L}(y) \frac{X_\pi(y)}{m^2_{\pi_0}} \eea where the $g^{5D}_{L,R}$ coupling are given by \bea g^{5D}_{L} = {g^{5D}\over \cos \theta_W } \left(T_3 - Q_q \sin^2 \theta_W \right) \eea \bea g^{5D}_{R} = {g^{5D}\over \cos \theta_W} Q_q \sin^2 \theta_W , \eea with $Q_q$ the charge of the quark, (here ${2 \over 3}$), $\theta_W$ the weak angle and $T_3 = {1\over 2}$. Note that when the interaction originates in the 5D Yukawa couplings, the profile to use is the one coming from the CP-odd Higgs component, i.e proportional to $X_\pi'(y)$. When the interaction originates in the gauge fermion coupling and thus comes from the $Z_5$ component, the profile to use is proportional to $X_\pi(y)$, with $X_\pi$ being the solution of Eq.~(\ref{Xeq}), using the decompositions of Eqs.~(\ref{Z5decomp}) and (\ref{Pidecomp}). When the fermion matrix in (\ref{mumatrix}) is diagonalized, the coupling matrix of fermions with the CP-odd field $\Pi_0(x)$ in (\ref{Ypimatrix}) is rotated, and we can then extract all the physical Yukawa couplings. All these couplings are needed later in order to compute the radiative couplings of $\Pi_0(x)$ with gluons and photons. \begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{yPiPoint6.pdf}\ \ \includegraphics[height=7cm]{yPiPoint5.pdf} \includegraphics[height=7cm]{yPiPoint4.pdf}\ \ \includegraphics[height=7cm]{yPiPoint2.pdf} \end{center} \vspace{-.2cm} \caption{Yukawa couplings between zero-mode fermions and the two lightest neutral scalars of the scenario, the $125$ GeV Higgs and the $750$ GeV CP-odd $\Pi_0$. The couplings are evaluated relative to the 5D bulk Higgs Yukawa coupling $Y_5$ and are shown as a function of the fermion bulk mass parameter $c_q$ (in the case $c_q=-c_u$ for simplicity) for different overall KK scales, $M_{KKglu_1}=3900$ GeV (upper left panel), $M_{KKglu_1}=2400$ GeV (upper right panel), $M_{KKglu_1}=1400$ GeV (lower left panel) and $M_{KKglu_1}=1000$ GeV (lower right panel). The CP-odd scalar mass is set to $750$ GeV and for certain values of $c_q$ its Yukawa coupling to top quarks can be highly suppressed for typical top-quark values of the $c_i$'s. } \label{fig:yPiplot} \vspace{-.2cm} \end{figure} We first analyze the very important Yukawa coupling between $\Pi_0(x)$ and top quarks, as this coupling might dominate the decays of the CP-odd scalar. This coupling comes essentially from the entry $(y^0_{\pi qu})_{ 33}$ (before rotation to the physical basis) although it receives small corrections after going to the physical basis. We focus on $(y^0_{\pi qu})_{ 33}$ which comes from the overlap integral \bea\label{YPtt} (y^{0}_{\pi qu})_{33} &=& i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} q^{0}_{t}(y)u^{0}_{t}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)} \eea In the RS limit, the warp factor is $\sigma(y)=ky$, and the top profiles are $q^{0}_{t}(y)= f(c_q) e^{(2-c_q)ky}$ and $u^{0}_{t}(y)=f(-c_u) e^{(2+c_u)ky} $, where $f(x)$ is a normalization factor. We also have $M_z(y)=\frac{g_5}{2c_W} v_0 e^{(a-1)ky}$, with $v_0$ a constant factor, so that the previous overlap integral in this limit reads \bea (y^{0}_{\pi qu})_{33} = i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \frac{2 c_W}{g_5}\frac{f(c_q) f(-c_u)}{v_0 m^2_{\pi_0}}\int_0^{y_1} dy\ e^{(2-a-c_q+c_u)k y} X_\pi'(y) \eea We integrate this by parts to find \bea \!\!\!(y^{0}_{\pi qu})_{33} &=& -i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \frac{2 c_W}{g_5}\frac{f(c_q) f(-c_u)}{v_0 m^2_{\pi_0}}\int_0^{y_1} dy\ (2-a-c_q+c_u) e^{(2-a-c_q+c_u)k y} X_\pi(y) + BT\ \ \ \ \ \ \eea where $BT= i \frac{(Y^{5D}_u)_{33}}{\sqrt{2k}} \frac{2 c_W}{g_5}\frac{f(c_q) f(-c_u)}{v_0 m^2_{\pi_0}}\ e^{(2-a-c_q+c_u)k y} X_\pi(y)\Big|_0^{y_1} $ is a boundary term. Note that the profile $X_\pi(y)$ has vanishing boundary conditions in the absence of Higgs localized brane kinetic terms. In that limit we can see that the coupling of the CP-odd scalar can actually vanish, when $(2-a-c_q+c_u)=0$ \cite{Archer:2012qa}. Note also that the Higgs localizer parameter $a$ is, in this RS limit, $a \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2$ and the bulk parameters $c_q$ and $c_u$ are defined such that, for example, charm or bottom quarks are assigned values more or less $c_q \in (0.45,0.55)$ and $c_u\in (-0.5,-0.6) $, whereas for top we have $c_{q_3}\sim 0.45$ and $c_{u_3}> -0.45$. This means that we expect the term $(2-a-c_q+c_u)$ to vanish, in the limit of $a\sim 2$, when $c_q-c_u \sim 0$, so that the suppression in this case seems only possible for the top quark, where both $c_q$ and $c_u$ could be small. Of course when the metric background is modified, and the boundary conditions include the brane kinetic terms there will be deviations from the previous values. Nevertheless it is clear that the Yukawa coupling of the CP-odd scalar field to top quarks can have highly suppressed values. Another way to see this is to consider the overlap integral in Eq.~(\ref{YPtt}). Because the profile $X_\pi(y)$ vanishes at the boundaries (or $almost$ vanishes, for small brane kinetic terms), then its derivative $X'_\pi(y)$ will have a node in the bulk, and therefore will change sign. That means that there can be some parameter choice for which it is possible for the overlap integral to vanish, since the fermion zero mode profiles have no nodes in the bulk. This feature is clearly seen in Fig. \ref{fig:yPiplot}, where we plot the absolute value of the Yukawa couplings between zero-mode fermions and both the Higgs and the CP-odd scalar $\Pi_0(x)$.\footnote{We are actually plotting the values defined in Eqs. (\ref{YYPI}) and (\ref{YYY1KK}), i.e. the zero mode Yukawa couplings before going to the fermion mass basis. In that basis, the couplings will inherit a small correction due to mixing with heavy KK fermions \cite{Azatov:2010pf}, so that the exact cancellation of the coupling will be replaced by a strong suppression.} The couplings shown are relative to the 5D bulk Higgs Yukawa coupling $Y_5$ and are plotted as functions of the fermion bulk mass parameter $c_q$ and $c_u$ (for the case where we take $c_q=-c_u$, for simplicity), for different overall KK scales. We observe that the CP-odd Yukawa couplings are fairly similar to the Higgs Yukawa couplings (i.e. exponentially sensitive to UV localization and then top-like when the zero mode is IR localized) except that there is a range of parameters where the coupling vanishes. Interestingly enough, this suppression happens for preferred values of the top quark bulk mass parameters. This means that the existence of suppressed couplings to top quarks of the CP-odd $\Pi_0$ is a natural possibility in this scenario. \subsection{Radiative couplings to photons and gluons} \label{subsec:gggamgam} Just like in the Higgs boson case, the radiative couplings of $\Pi_0(x)$ to gluons and photons will depend on the physical Yukawa couplings $y_{nn}$ between $\Pi_0$ and the fermions (zero modes and KK modes) running in the loop, as well as on the fermion masses $m_n$ (the eigenvalues of the mass matrix in Eq.(\ref{mumatrix})). The real and imaginary parts of the couplings are associated with different loop functions, $A^S_{1/2}$ and $A^P_{1/2}$, as they generate the two operators $\Pi_0G_{\mu\nu}G^{\mu\nu}$ and $\Pi_0 G_{\mu\nu}\tilde{G}^{\mu\nu}$.\footnote{The Yukawa couplings of $\Pi_0$ are mostly imaginary and thus the dominant contribution will come, as expected, from the operator $\Pi_0 G_{\mu\nu}\tilde{G}^{\mu\nu}$. Still, small real Yukawa coupling components are generated when going to the fermion mass basis, and so we keep the general formalism in our formulas.} The production cross section through gluon fusion is \bea \sigma_{gg\rightarrow \Pi_0} = {\alpha_s^2 m_{\Pi_0}^2\over 576 \pi} \left[\Big|\sum_{quarks}c^S_n\Big|^2 + \Big|\sum_{quarks} c^P_n\Big|^2\right]\ \eea and the decay widths to gluons and photons are \bea \Gamma_{\Pi_0\to gg} = \frac{\alpha_s^2 m^3_{\Pi_0}}{54\pi^2}\frac{1}{v^2}\left[\Big|\sum_{{quarks }}c^S_n\Big|^2 + \Big|\sum_{{quarks}} c^P_n\Big|^2\right]\ \eea \bea \Gamma_{\Pi_0\to \gamma\gamma} = \frac{\alpha^2 m^3_{\Pi_0}}{192\pi^3}\frac{1}{v^2} \left[\Big|\sum_{{quarks \atop leptons}} N_c Q_n^2 c^S_n\Big|^2 + \Big|\sum_{{quarks \atop leptons}} N_c Q_n^2 c^P_n\Big|^2\right]\ \eea where $\alpha_s$ and $\alpha$ are the strong and weak coupling constants, $N_c$ is the number of colors and $Q_n$ is the charge of the fermion, and where \bea c^S_n =\re\left({\frac{y_{nn}}{m_{n}}}\right) A^S_{1/2}(\tau_n) \hspace{.6cm} {\rm and} \hspace{.6cm} c^P_n = {\rm Im}\left(\frac{y_{nn}}{m_{n}}\right) A^P_{1/2}(\tau_n) \eea with $\ \tau_n = m^2_{\Pi_0}/4m^2_n\ $ and with the loop functions defined as \cite{Gunion:1989we} \bea\label{eq:loopfunction} A^S_{1/2}(\tau) &=& {3\over2}\left[\tau + (\tau -1)f(\tau)\right]\tau^{-2}, \\ A^P_{1/2}(\tau) &=&- {3\over2} f(\tau)/\tau , \eea and with \bea f(\tau) = \left \{ \begin{tabular}{cc} $\left[\arcsin\sqrt{\tau}\right]^2$ & $(\tau\le 1)$\\ $ -{1\over4}\left[\ln\left({1+ \sqrt{1-\tau^{-1}}\over 1- \sqrt{1-\tau^{-1}}}\right)-i\pi\right]^2$& $(\tau> 1)$ . \end{tabular} \right. \eea For heavy KK quarks with masses $m_n$ much greater than the CP-odd mass $m_{\Pi_0}$ (i.e. when $\tau$ is very small) the loop functions are essentially constant, as they behave asymptotically as $\displaystyle \lim_{\tau \to 0} A^S_{1/2} =1 \hspace{.2cm} \mbox{and} \hspace{.2cm} \lim_{\tau \to 0} A^P_{1/2} =3/2.$ On the other hand, for light quarks (all the SM quarks except top and bottom), the loop functions essentially vanish asymptotically as $\displaystyle \lim_{\tau \to \infty} A^S_{1/2} = \lim_{\tau \to \infty} A^P_{1/2} = 0.$ Moreover, we investigate a parameter region where the couplings of $\Pi_0$ to top quarks are highly suppressed. This means that the production mechanism must rely exclusively on the heavy KK fermions running in the loop and as we have seen, this coupling depends on the ratio $\frac{y_{nn}}{m_n}$ between the physical Yukawa coupling and the mass of the fermion running in the loop. To have an idea of the relative contribution of each of these KK fermions in the loop, in Fig. \ref{fig:yPiKKplot} we plot the mass normalized Yukawa couplings of Standard Model Higgs with top quarks, Higgs with first KK fermion and of $\Pi_0$ to first KK fermion, for different values of the KK scale. As expected we see that the $c_q$ dependence is mild (i.e. all KK fermions of any flavor will couple with similar strength) and also, as expected, we observe that the mass normalized couplings are quite suppressed with respect to the SM top quark case. Still the multiplicity of KK fermions is high, since there are 6 families of quarks and 3 families of charged leptons (the latter run in the diphoton loop), and for each family there are a few KK levels that give important contributions to the rate. \begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{yPiKKPoint6.pdf}\ \ \includegraphics[height=7cm]{yPiKKPoint5.pdf} \includegraphics[height=7cm]{yPiKKPoint4.pdf}\ \ \includegraphics[height=7cm]{yPiKKPoint2.pdf} \end{center} \vspace{-.2cm} \caption{Yukawa couplings between the lightest KK fermion and the $125$ GeV Higgs (middle curves) and the $750$ GeV CP-odd $\Pi_0$ (lower curves), divided by the KK fermion mass, for different values of the lightest KK gluon mass $M_{KKglu_1}$, as indicated on the panels. This mass normalized Yukawa coupling gives an estimate of the relative contribution of the respective KK fermion to the radiative coupling of the scalar to gluons and photons, to be compared with the mass normalized SM coupling of Higgs to top quarks (shown as a dashed line).} \label{fig:yPiKKplot} \vspace{-.3 cm} \end{figure} A numerical scan of the couplings, including all families and 3 full KK levels is computationally too intensive, so in order to produce the couplings plotted in Fig. \ref{fig:yPiKKplot} we performed an approximation, sufficient for the purposes of the graph. The KK fermion Yukawa couplings plotted neglect mixings between different KK levels and different flavors, and with the zero mode fermions. They are obtained as follows. Consider the $2\times2$ KK mass matrix \bea \left(\begin{matrix} Q_L(x)\ U_L(x) \end{matrix}\right)\ M_u\ \left(\begin{matrix} Q_R(x)\\ U_R(x) \end{matrix}\right) , \eea where $Q_L(x)$ and $Q_R(x)$ represent here a single flavor and a single KK level of the vector-like KK up-type doublets, and similarly for $U_L(x)$ and $U_R(x)$, vector-like KK up-type singlets. The mass matrix is thus \bea M_u= \left(\begin{matrix} m_Q & Y_1 \\ Y_2 & m_U \end{matrix}\right), \label{QUmatrix} \eea where the diagonal entries are the KK masses (large) whereas the off-diagonal entries are coming from Yukawa couplings and are therefore smaller. In order to give a simple estimate, we take for simplicity the fermion bulk mass parameters as $c_q=-c_u$, and the bulk Higgs Yukawa $Y^{5D}$ to be real, which leads to $Y_1=Y_2$ and $m_Q=m_U=m_{KK}$, with the masses and profiles obtained by solving Eq. (\ref{fermionprofile}). With the KK fermion profiles one obtains the off-diagonal entries \bea Y_1= \frac{(Y^{5D}_u)}{\sqrt{k}} \int_0^{y_1} dy e^{-4\sigma(y)} \frac{ v(y)}{\sqrt{2}} Q_{L}(y)U_{R}(y) . \eea The matrix that diagonalizes (\ref{QUmatrix}) in this simple limit ($c_q=-c_u$) is $\left(\begin{matrix} -\frac{1}{\sqrt{2}} & \frac{1}{\sqrt{2}} \\ \frac{1}{\sqrt{2} }& \frac{1}{\sqrt{2}} \end{matrix}\right)$ and the eigenvalues are $m_1=m_{KK}-Y_1$ and $m_2=m_{KK}+Y_1$. Now we apply this rotation to the CP-odd Yukawa coupling matrix \bea Y_\Pi= \left(\begin{matrix} {\cal O}(g) & Y_\pi \\ Y_\pi & {\cal O}(g) \end{matrix}\right), \eea where \bea Y_\pi& =& i \frac{(Y^{5D}_u)_{ij}}{\sqrt{2k}} \int_0^{y_1} dy e^{-3\sigma(y)} Q_{L}(y)U_{R}(y) \frac{X_\pi'(y)}{m^2_{\pi_0} M_z(y)}, \eea and where for simplicity we have neglected gauge couplings terms compared to IR Yukawa terms (safe assumption when $Y^{5D}$ is large). After diagonalization, we obtain the two physical couplings between $\Pi_0$ and the KK fermions. When we normalize the couplings by the two eigenmasses and add the two contributions,\footnote{One needs to add the two contributions since there is a cancellation happening level by level.} we obtain \bea \sum_{i=1}^2 \frac{y_i}{m_i} =\frac{y_1^{\pi}}{m_1}+\frac{y_2^{\pi}}{m_2} = -2 \frac{Y_1 Y_\pi}{m_{KK}^2-Y_1^2} \simeq -2\frac{Y_1 Y_\pi}{m_{KK}^2}. \eea The last expression corresponds to the {\it mass normalized} Yukawa couplings of $\Pi_0$ plotted in Fig. \ref{fig:yPiKKplot}, and this describes very closely the behavior of the couplings obtained in the full flavor calculation. The parametric dependence of these couplings is $Y_{5D}^2 v/m_{KK}^2$, so that if $Y^{5D}\sim 3$ we expect mass normalized couplings of order $(10^{-3}-10^{-4})$ GeV$^{-1}$, if the overlap integral is of ${\cal O}(1)$. Since all the profiles of the integral are IR localized, one expects that integral to be ${\cal O}(1)$, although the precise numerical result varies between $0.5$ and $0.05$, depending on the values of the $c_q$ parameter, as shown in the plots. All in all it seems likely that after taking into consideration all the fermion flavors, and for a KK scale of order $1-2$ TeV, the overall KK fermion contribution to the radiative couplings of $\Pi_0(x)$ to photons and gluons can be close to the top quark contribution to the gluon and photon couplings of the Higgs in the SM model. \subsection{$\Pi_0Zh$ coupling} \label{subsec:zh} The coupling between the CP-odd scalar, the $Z$ boson and the Higgs will be extracted from the kinetic operator of the 5D Higgs, \bea \int d^4x dy\ e^{-2\sigma} D_\mu H^\dagger D^\mu H \left(1+ \delta(y-y_i) \frac{d_i}{k}\right). \eea Expanding the SM-like Higgs mode using Eq. (\ref{Hexpansion}) as well as the SM-like $Z_\mu$ and the $750$ GeV $\Pi_0$ using Eqs. (\ref{Aexpansion}) and (\ref{Piexpansion}), we can obtain the coefficient $g_{\Pi hZ}$ of the operator $Z^\mu(x)\Big(h(x)\partial_\mu\Pi_0(x)+\Pi_0(x)\partial_\mu h(x)\Big)$ \bea g_{\Pi hZ} = \frac{g_5^2}{4c_W^2} \int dy e^{-2\sigma} v(y)h(y) f_z(y) \frac{X'(y)}{M^2_z(y) m^2_{\pi}} \left(1+ \delta(y-y_i) \frac{d_i}{k}\right) \eea Now since $M_z(y)=\frac{g_5}{2c_W}e^{-\sigma} v(y)$, $h(y)\sim v(y)/v_4$, and $f_z\simeq 1/\sqrt{y_1}$ we can write \bea g_{\Pi hZ} \simeq \frac{1}{\sqrt{y_1}v_4 m^2_\pi} \big(X(y_1)+ X'(y_1)\frac{d_1}{k}\big)=0, \eea where we have used the boundary conditions for the profile $X(y)$ (see Eq.(\ref{Xbc})) and assumed no UV brane kinetic term ($d_0=0$). The coupling should thus vanish in the limit of flat $Z$ boson profile $f_z(y)$, and when the nontrivial Higgs VEV $v(y)$ is proportional to the Higgs scalar profile $h(y)$. Corrections to these limits scale as $v_4^2/m_{KK}^2$ and $m_h^2/m_{KK}^2$ in the RS case, and so we expect the overall coupling to be highly suppressed. The partial width for the decay $\Pi_0 \to h Z$ is \cite{djouadi1996} \bea \Gamma(\Pi_0 \to h Z)= \frac{g^2_{\Pi hZ}}{16\pi} \frac{m_Z^2}{m_{\pi}} \sqrt{\lambda(m^2_h,m^2_Z;m^2_\pi)}\lambda(m^2_h,m^2_\pi;m^2_Z) \eea where $m_Z$, $m_\pi$ and $m_h$ are the masses of the particles involved and where $\lambda(x,y;z) = (1 - x/z - y/z)^2 - 4 x y/z^2$. With the masses $m_Z=91$ GeV, $m_\pi=750$ GeV and $m_h=125$ GeV, the width becomes $\Gamma(\Pi_0 \to h Z) \sim (900\ g^2_{\Pi hZ})$ GeV. For example we compute numerically $g_{\Pi hZ}$ for three different values of $M_{KKglu_1}$, and with $m_{\Pi_0}=750$ GeV and find \begin{center} \begin{tabular}{|c||c|c|c|c|} \hline $M_{KKglu_1}$ &\ 1000 GeV\ &\ 1400 GeV\ &\ 2400 GeV\ \\ \hline $g_{\Pi hZ}$ & $3.8\times 10^{-4}$ & $3.1\times 10^{-3}$ &$1.1\times 10^{-2}$ \\ \hline $\Gamma(\Pi_0 \to hZ$) &$1.3\times 10^{-4}$ GeV &$8.8\times 10^{-3}$ GeV &$0.11$ GeV \\ \hline \end{tabular} \end{center} Note that the couplings and widths are small but we observe that the partial width becomes larger as the KK mass scale is increased. Since we need the partial width $\Gamma(\Pi_0 \to \gamma \gamma)$ to be similar to that of a $750$ GeV Higgs, i.e. $\Gamma(h_{750GeV} \to \gamma \gamma) \sim 10^{-5}$ GeV, we expect that at $M_{KKglu_1} =1400$ GeV the $Zh$ signal should start putting too much pressure on the allowed parameter space. This is confirmed in the full numerical analysis presented in the next section. \subsection{Estimates and numerical results} \label{subsec:numerical} With all the previous ingredients one can estimate the viability of this scenario in terms of the possible diphoton excess. Let's choose the KK scale such that $M_{KKglu_1}=1400$ GeV; when the bulk mass parameters of the top quark are around $|c_{u_3}|\sim 0.35$ we know that the top quark will have highly suppressed couplings to $\Pi_0$, as shown in the third panel of Fig. \ref{fig:yPiplot} . At the same time, the couplings of the KK tops (as well as all other KK quarks) will have relatively strong Yukawa couplings to $\Pi_0$ (third panel of Fig. \ref{fig:yPiKKplot}), so that the contribution of each of them to the radiative coupling of $\Pi_0$ to gluons is about an order of magnitude smaller than the top contribution to the $h-glu-glu$ coupling of the SM. Thus we could estimate that the overall contribution of all flavors and KK excitations can make up for the suppressed coupling, so that the production cross section of $\Pi_0$ is similar to that of a $750$ GeV SM-like Higgs. The production cross section of a $750$ GeV Higgs through gluon fusion, at the LHC running at 13 TeV is $497$ fb \cite{Dittmaier:2011ti}, so, roughly, this could be assumed for the $\Pi_0$ production cross section. Since the $\Pi_0$ decays to top quarks and $hZ$ are suppressed in this parameters space point, and its decays to $WW$ and $ZZ$ can only be radiative via the CP-odd gauge boson kinetic operator, the main decay channel is into gluons so that the branching of the diphoton channel should be very roughly \bea Br(\Pi_0 \to \gamma\gamma) \sim \frac{\alpha^2_{em}}{8 \alpha^2_s} \frac{N_\gamma}{N_{glu}} \eea where $N_\gamma$ and $N_{glu}$ are the multiplicities of states running in the $(\Pi_0 \gamma \gamma)$ loop and in the $(\Pi_0\ glu\ glu)$ loop respectively. In the diphoton loop there are $3$ extra families of charged lepton KK excitations making the multiplicity of states greater. If their multiplicity and their Yukawa couplings can partially make up for the color factor of 8, then the diphoton cross section might become of ${\cal O}$ (1 fb), as hinted by the December 2015 LHC data. To complete the analysis, we perform a full numerical computation of production and branching ratios in a setup where we consider an effective 4D scenario including three full KK levels for all fields, i.e., we consider $21 \times 21$ fermion mass matrices, which we diagonalize in order to obtain the physical Yukawa couplings. We choose a set of $c$-parameters and 5D Yukawa entries such that the SM masses and mixings are reproduced; the specific flavor choice for these parameters should not affect much the overall results since these depend on overlap integrals between IR localized fields, with very loose $c$-dependence. We choose the background metric parameters so that precision electroweak bounds are kept at bay, i.e. $\nu=0.5$ and $y_s=1.04 y_1$. Two average 5D Yukawa scales are considered, $Y_{5D} \simeq 3$ and $Y_{5D}\simeq 2$, to show the dependence on this parameter, and we also consider two different KK mass scales, $M_{KKglu_1}=1000$ GeV and $M_{KKglu_1}=1300$ GeV, which turn out to lead to successful signal generation. In order to see how tuned is the choice of top $c$-parameter, we plot the production cross section of the CP odd resonance, followed by decays into $\gamma \gamma$, $t\bar{t}$ and $Zh$, as functions of $c_{u_3}$ (the bulk mass parameter of the 5D singlet top quark), with the doublet bulk mass parameter fixed at $c_{q_3}=0.4$. (This value ensures typically suppressed bounds from $Zb_L{\bar b}_L$ bounds \cite{Cabrer}). \begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{prodplotPoint23}\ \ \includegraphics[height=7cm]{prodplotPoint22} \end{center} \vspace{-.2cm} \caption{Production cross sections at the LHC running at 13 TeV in the channels $\gamma\gamma$, $t\bar{t}$ and $hZ$. The lightest KK boson mass is $1000$ GeV and the 5D Yukawa couplings are $Y_{5D}\simeq 3$ in the left panel and $Y_{5D}\simeq 2$ in the right panel. The signal can always be produced at this KK scale but for larger 5D Yukawa couplings there are either too many top pairs or too many dijets produced. Overall we can conclude that for $M_{KKglu_1} = 1000$ GeV and when $1\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} Y_{5D}\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$ and $0.2\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} |c_{u_3}| \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.5$ the signal is easily reproduced.} \label{fig:prodplot1} \vspace{0. cm} \end{figure}\begin{figure}[t] \center \begin{center} \includegraphics[height=7cm]{prodplotPoint33}\ \ \includegraphics[height=7cm]{prodplotPoint32} \end{center} \vspace{-.2cm} \caption{Same as the previous figure but for a KK scale of $1300$ GeV. In this case the signal can only be achieved for larger Yukawa couplings, and at a more localized region near $c_{u_3}\sim 0.25$. The bound from $hZ$ now becomes the main constrain. For heavier KK mass scales, the signal can only be produced for much larger Yukawa couplings and the $hZ$ bound becomes too constraining.} \label{fig:prodplot2} \vspace{0. cm} \end{figure} The results are shown in Figs. \ref{fig:prodplot1} and \ref{fig:prodplot2}, where we have chosen two KK scales, $M_{KKglu_1}=1000$ GeV and $M_{KKglu_1}=1300$ GeV. In both cases we show results for $Y_{5D}\simeq 3$ and $Y_{5D}\simeq 2$ to illustrate the sensitivity on this bulk parameter, crucial for enhancing the radiative couplings of $\Pi_0$. When the KK scale is smaller and 5D Yukawa couplings are larger, the production of top pairs and gluon pairs can be too large. By reducing the 5D Yukawa couplings, enough signal can be generated with the dijets and top pairs under control, as well as the $Zh$ decay. The values of $|c_{u_3}|$ must be in the region $0.2\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} |c_{u_3}| \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.5$. For slightly smaller KK scales one expects a similar behavior, but such that 5D Yukawa couplings should be smaller in order to suppress overproduction of $\Pi_0$ particles. This leads to the question of how large can the KK scale be and still manage to produce enough signal. We observe that at $M_{KKglu_1}=1300$ GeV, larger 5D Yukawa couplings ($Y_{5d} \geq 3$) are required in order to have enough signal production. The branching fraction into $hZ$ starts to be problematic, and actually becomes worse for heavier KK scales. The signal production requires that the value of $|c_{u_3}|$ be located around the point where the top Yukawa couplings of $\Pi_0$ are suppressed, in this case around $|c_{u_3}|\sim 0.3$. We conclude therefore that in order to explain the $750$ GeV diphoton, the KK scale must be $M_{KKglu_1} \leq 1300$ GeV, with 5D Yukawa couplings $Y_{5D}\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$. For those values, the signal is quite generic (i.e. small, but typical, $c_{u_3}$ values are required), as long as the $\Pi_0$ mass is set to $750$ GeV with appropriate boundary kinetic terms. \section{Discussion} \label{sec:discussion} We performed an analysis of the scalar sector of warped space models to investigate whether the {\it minimal} model can accommodate a resonance at 750 GeV with the properties observed in the diphoton resonance at the CMS and ATLAS. We show that in the simplest extra dimensional extension of the SM, that is with a 5D Higgs doublet living in the bulk, the lowest pseudoscalar KK excitation can be responsible for the signal observed at 750 GeV. We emphasize that, unlike other explanations relying on scalar fields in warped models, ours does not introduce any new fields or representations, but relies on Higgs brane kinetic terms to lower the KK mass of the lightest CP odd Higgs resonance. This makes the model extremely constrained, with the only new parameter being the IR brane kinetic coefficient $d_{1}$. The lightest CP odd excitation, a mixture of the 5D Higgs field and $Z_5$, does not decay at tree level into $WW$ or $ZZ$, and, over a range of the parameter space, can have suppressed couplings to the top quark, thus a small decay width into $t {\bar t}$. The production through gluon fusion can be loop-enhanced through the effects of the usual KK fermion modes, and so can the diphoton decay. The coupling to $Zh$ is also suppressed although starts increasing dangerously for KK masses above $1500$ GeV. We also showed that in $AdS_5$ spaces (RS-type models) with a KK scale as low as $1500$ GeV, the lowest CP-odd scalar cannot have a mass at 750 GeV, but if the metric differs slightly from the $AdS_5$ metric (generalized warped metric spaces), the Higgs brane kinetic terms can produce a viable CP-odd scalar of mass 750 GeV. Within these modified metric scenarios, and for KK mass scales at around 1 TeV (consistent with precision electroweak bounds) this CP odd resonance obeys the (current) experimental constraints. We analyzed its production and decay for several values of the lowest KK gluon mass ($M_{KKglu_1}$), and show that consistency with the data requires $M_{KKglu_1} \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 1300$ GeV. Our analysis is quite general, and should be valid even in the absence of a signal at 750 GeV. The general conclusion to be taken from our analysis here is that warped space models, without any new particles, can explain a (relatively) light diphoton resonance at the LHC. Should a diphoton excess be found at higher values, even the RS model might accommodate such a state without the need modify the metric. Among the general features of the light CP-odd scalar resonance, resulting from the 5D Higgs doublet are the fact that its coupling to top pairs can be suppressed for appropriate top bulk mass parameters. Also its coupling to $Zh$ is generically suppressed due to the boundary conditions of the CP odd state. In addition, the model predicts that the spectrum for the CP odd and the charged scalars is essentially the same since their differential equations and boundary conditions are almost identical. This means that the lightest charged Higgs boson is expected to have a mass very close to the pseudoscalar mass, so about 750 GeV, in the scenario in which the latter is the diphoton resonance. If the diphoton resonance ends up being a real particle, and not a statistical fluctuation, the prediction from this scenario is that $ZZ$ and $WW$ should not be seen, whereas top pair production, dijet production and $Zh$ signal should be around the corner. Moreover a search for the charged scalars at around $750$ GeV might prove useful in order to disentangle the different models explaining the resonance. \section{Acknowledgments} MF and MT thank NSERC and FRQNT for partial financial support under grant numbers SAP105354 and PRCC-191578. \section{Appendix} In this section we consider the effect of brane localized kinetic terms associated with the 5D Higgs doublet and also with the gauge bosons. For simplicity, let's consider a 5D toy model with a Higgs scalar $H(x,y)$ charged under a local $U(1)$, defined by the following action: \bea S&=&\int d^4xdy \sqrt{g}\ \left(-\frac{1}{4} F^2_{MN} + |D^M H|^2 - V(H) \right) \\ &&\ \ \ + \sum_i \int d^4xdy \sqrt{g}\ \delta(y-y_i) \left(\frac{1}{4} r_i F^2_{MN} + d_i |D^M H|^2 - \lambda_i(H))\right) \label{5Daction} \eea where $F_{MN} = \partial _M A_N - \partial_N A_M$ and for simplicity we set gauge coupling constant to unity in the appropriate mass dimensions. The background spacetime metric is assumed to take the form \bea ds^2=e^{-2\sigma(y)}\eta_{\mu\nu} dx^\mu dx^\nu - dy^2 \eea where $\sigma(y)$ is the warp factor. We are interested in studying the effective 4D perturbative spectrum of the 5D Higgs field and the 5D gauge boson, around a nontrivial Higgs vacuum profile solution $<H>=v(y)$. \bea H(x,y)= \frac{1}{\sqrt{2}}\left(v(y) + h(x,y)\right) e^{i \pi(x,y)} \eea In particular we are interested in the CP-odd Higgs perturbations $\pi(x,y)$, whose equations of motion are coupled with the gauge boson perturbations. The equations read \bea && (1+r_i \delta_i) \partial_\mu \partial^\mu A_\mu - \left( (1+r_i \delta_i)e^{-2\sigma} A'_\mu\right)' + (1+d_i \delta_i) M_A^2 A_\mu +\non\\ &&\hspace{3cm}\partial_\mu \left((1+d_i \delta_i) M_A^2 \pi - (1+r_i \delta_i)\partial^\nu A_\nu - \left((1+r_i \delta_i)e^{-2\sigma} A_5\right)'\right) = 0\\ &&(1+r_i \delta_i) \partial_\mu \partial^\mu A_5- (1+r_i \delta_i) \partial^\nu A'_\nu+ (1+d_i \delta_i)M_A^2 (\pi'-A_5)=0\\ &&(1+d_i \delta_i) \partial_\mu \partial^\mu \pi - (1+d_i \delta_i) \partial^\nu A_\nu + M_A^{-2}\left((1+d_i \delta_i) M_A^2 e^{-2\sigma} (\pi'-A_5)\right)'= 0 \eea where $M_A= v(y) e^{-\sigma}$ and where $\displaystyle d_i \delta_i \equiv \sum_i d_i \delta(y-y_i) $ and $\displaystyle r_i \delta_i \equiv \sum_i r_i \delta(y-y_i) $. We fix partially the 5D gauge by imposing \bea (1+d_i \delta_i) M_A^2 \pi - (1+r_i \delta_i)\partial^\nu A_\nu - \left((1+r_i \delta_i)e^{-2\sigma} A_5\right)' = 0. \label{gaugefix} \eea The previous gauge fixing equation reads in the bulk: \bea M_A^2 \pi - \partial^\nu A_\nu - \left(e^{-2\sigma} A_5\right)' = 0. \label{bulkgaugefix} \eea Note that if we evaluate the bulk constrain Eq.~(\ref{bulkgaugefix}) at $y=y_1-\epsilon$ (i.e. right before the IR brane), we obtain: \bea \partial^\nu A_\nu \Big|_{y_1-\epsilon} = M_A^2 \pi - \left(e^{-2\sigma} A_5\right)'\Big|_{y_1-\epsilon}. \label{epsiloncondition} \eea On the other hand, the effect of the delta functions in Eq.(\ref{gaugefix}) is to produce a discontinuity in the 5D field $A_5$ at the brane location as \bea d_1 M_A^2 \pi -r_1 \partial^\nu A_\nu\Big|_{y_1-\epsilon} = \left[e^{-2\sigma} A_5\right]_{y_1-\epsilon}^{y_1}, \eea and similarly for the UV brane. We can thus multiply (\ref{epsiloncondition}) by $r_1$ and use it in the previous equation, and find the necessary boundary condition between $\pi$ and $A_5$, which ensures that $A_\mu$ is completely decoupled, even on the brane. We find \bea (d_1-r_1) M^2_A \pi + r_1 \left(e^{2\sigma} A_5\right)' \left. \vphantom{\int^A_A} \right|_{y_1-\epsilon} =- \left.e^{-2\sigma} A_5\right|_{y_1-\epsilon} \label{boundaryconditions} \eea where we have taken $A_5$ to vanish exactly on the brane, but it jumps right before the boundary. Inserting the gauge choice in the coupled equations of motion, one manages to decouple the gauge modes $A_\mu$ (in both the bulk and the branes) with a bulk equation \bea \partial_\mu \partial^\mu A_\mu - \left(e^{-2\sigma} A'_\mu\right)' + M_{A}^2 A_\mu =0 \eea and jump condition on $A'_\mu$ \bea r_1 \partial_\mu \partial^\mu A_\mu + d_1 M_{A}^2 A_\mu \Big|_{y_1-\epsilon}=- \left. e^{-2\sigma} A'_\mu\right|_{y_1-\epsilon} \eea where $A'_\mu$ again vanishes exactly on the brane, but has a jump right before it. We separate variables \bea A_\mu(x,y)= V^{4d}_\mu(x) V_y(y) \eea and find the separated equations for the gauge boson tower become \bea && \partial_\mu \partial^\mu V^{4d}_\mu(x)+m^2_A V^{4d}_\mu(x)=0\\ && \left( e^{-2\sigma} V'_y\right)' + ( m^2_A - M_A^2 ) V_y =0 \eea with jump conditions on $V'_y$ \bea (d_i M_{A}^2- r_i m^2_A) V_y \Big|_{y_1-\epsilon}= -\left. e^{-2\sigma} V'_y\right|_{y_1-\epsilon} \eea where the 4D effective mass $m^2_A$ is the constant of separation of variables. The remaining equations are, in the bulk, \bea && \partial_\mu \partial^\mu A_5+ M_A^2 (\pi'-A_5) - \left(M_A^2 \pi\right)' +\left( \left(e^{-2\sigma} A_5\right)' \right)' =0\\ && \partial_\mu \partial^\mu \pi + M_A^{-2} \left( M_A^2 e^{-2\sigma} (\pi'-A_5) \right)' - M_A^2 \pi + \left(e^{-2\sigma} A_5\right)'= 0\label{a5pimixedequation} \eea and the fields must verify the boundary conditions of Eq.(\ref{boundaryconditions}). We now perform a mixed separation of variables: \bea A_5(x,y)&=& G(x) g(y) + \pi_x(x) \eta(y) \label{expansion1}\\ \pi(x,y)&=& G(x) h(y) + \pi_x(x) \xi(y)\label{expansion2} \eea which is to say that both $A_5(x,y)$ and $\pi(x,y)$ contain each some Goldstone and CP-odd degree of freedom. The profiles $g(y)$, $\eta(y)$, $h(y)$ and $\xi(y)$ quantify how much of each they contain. Of course the functions $g$ and $h$ are related to each other, as well as $\eta$ and $\xi$. The relationships are such that $G(x)$ and $\pi_x(x)$ decouple. With the choice \bea h(y)&=& \frac{K(y)}{m^2_G}\\ g(y)&=&\frac{K'(y)}{m^2_G}\\ \eta(y)&=&\frac{e^{2\sigma}}{m^2_\pi}X(y)\\ \xi(y)&=&\frac{1}{m^2_\pi M^2_A}X'(y) \eea and using the mixed separation of variables in (\ref{expansion1}) and (\ref{expansion2}), the mixed equations of motion in (\ref{a5pimixedequation}) decouple and we obtain \bea && \frac{e^{2\sigma}}{M_A^2} X(y) \partial_\mu \partial^\mu \pi_x(x) +\pi_x(x) e^{2\sigma} X(y) - \pi_x(x) \left[M_A^{-2} X'(y)\right]'= 0\\ &&K(y) \partial_\mu\partial^\mu G(x)+ G(x) \left[ \left(K'e^{-2\sigma}\right)' + M_A^2 K(y)\right] = 0 . \eea Once separated, we obtain, for the CP odd physical scalars \bea &&\partial_\mu \partial^\mu \pi_x(x) + m_\pi^2 \pi_x(x) =0\\ &&\left(M_A^{-2} X'\right)'+ e^{2\sigma} \left(\frac{m_\pi^2}{M_A^2}-1\right) X=0 , \eea with boundary conditions \bea d_i X' = -X \eea and for the Goldstone modes, \bea &&\partial_\mu\partial^\mu G(x)+ m_G^2 G(x)=0\\ &&\left(K'e^{-2\sigma}\right)'+ (m_G^2-M_A^2) K=0, \eea with boundary conditions \bea (d_i M_A^2-r_i m^2_G ) K = -e^{-2\sigma} K' . \eea Note that both the equations and boundary conditions for the Goldstone bosons are identical to the ones for the gauge boson tower, as should be, so that they can then be gauged away level by level with the remaining gauge fixing freedom.
7b429a8e453dc225a2ad4a516910f10a69b36da2
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section*{Acknowledgments} \noindent This research was supervised by Ken Ono and David Zureick-Brown at the Emory University Mathematics REU and was supported by the National Science Foundation (grant number DMS-1557960). We would like to thank David Zureick-Brown for suggesting the problem that led to the present article and for offering us his invaluable advice and guidance. We would also like to thank David Zureick-Brown for providing the intuition behind the proof of Lemma~\ref{proposition:lifting-l}. We would like to thank Jackson Morrow, Ken Ono, and David Zureick-Brown for making several helpful comments regarding the composition of this article. We would like to acknowledge Michael Aschbacher and Nick Gill for their helpful advice. We used {\tt Magma} and \emph{Mathematica} for explicit calculations. \bibliographystyle{alpha} \section{Proof of Theorem~\ref{jamesdreamcometrue}}\label{doyoulift} In this section, we provide a complete proof of the main theorem of this paper, namely Theorem~\ref{jamesdreamcometrue}. The basic strategy has two steps: lift from $\ell^2$ to $\ell^\infty$ (see Section~\ref{allthewaytonight}), and lift from $\ell$ to $\ell^2$ (see Section~\ref{bataneye}). Considerable care must be taken in dealing with the cases where $\ell = 2, 3$, so we treat these situations separately (see Sections~\ref{booyah} and~\ref{thisishisbananas}). We execute this strategy as follows: \subsection{Lifting from $\ell^2$ to $\ell^\infty$ for $\ell \ge 3$, and from $8$ to $2^\infty$} \label{allthewaytonight} \begin{lemma} \label{proposition:lifting-l-squared} If $\ell \ge 3$, then $H(\ell^2) = \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z)$ implies $H_\ell = \on{Sp}_{2g}(\bz_\ell)$. If $\ell = 2$, then $H(8) = \on{Sp}_{2g}(\bz / 8 \mathbb Z)$ implies $H_2 = \on{Sp}_{2g}(\bz_2)$. \end{lemma} \begin{proof} This is done in the $g = 1$ case in in~\cite[Lemma 3, Section IV.3.4]{serre1989abelian}, which readily generalizes to the case of $g \geq 2$. \end{proof} \subsection{Lifting from $\ell$ to $\ell^2$ for $\ell \ge 5$}\label{bataneye} \begin{lemma} \label{proposition:lifting-l} Fix $g \ge 2$. If $\ell \ge 5$, then $H(\ell) = \on{Sp}_{2g}(\bz / \ell \mathbb Z)$ implies $H(\ell^2) \subset \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z)$. \end{lemma} \begin{proof} It suffices to show that $H(\ell^2)$ contains all of \[ \ker( \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) \twoheadrightarrow \on{Sp}_{2g}(\bz / \ell \mathbb Z)) = \mathrm{id}_{2g} + \ell \cdot \mathfrak{sp}_{2g}(\bz / \ell \mathbb Z). \] We prove this by taking $\ell$-th powers of specific matrices. We want to pick $M \in \mathfrak{sp}_{2g}(\bz / \ell \mathbb Z)$ such that $\mathrm{id}_{2g} + M$ lies in $\on{Sp}_{2g}(\bz / \ell \mathbb Z)$ and such that $M^2 = 0$. As it happens, these two conditions are equivalent: indeed, since $M^T \Omega_{2g} + \Omega_{2g}M = 0$, we have \begin{align*} (\mathrm{id}_{2g} + M)^T \Omega_{2g} (\mathrm{id}_{2g} + M) &= \Omega_{2g} + M^T \Omega_{2g} + \Omega_{2g}M + M^T \Omega_{2g} M \\ &= \Omega_{2g} + M^T \Omega_{2g} M \\ &= \Omega_{2g} - \Omega_{2g}M^2, \end{align*} so the condition that $\mathrm{id}_{2g} + M$ is symplectic is equivalent to the condition that $M^2 = 0$. Moreover, by expanding in terms of matrices, we see that $M^2 = 0$ if and only if \[ \left[\begin{array}{c|c} A^2 + BC & AB - BA^T \\ \hline CA - A^T C & CB + (A^T)^2 \end{array}\right] = 0. \] So, if $A^2 = 0$ and two of $A, B,$ and $C$ are zero, the matrix $\mathrm{id}_{2g} + M$ will be symplectic. For $M \in \mathfrak{sp}_{2g}(\mathbb Z/\ell \mathbb Z)$, choose an arbitrary lift of $M$ in $\operatorname{Mat}_{2g \times 2g}(\bz / \ell^2 \mathbb Z)$, and by abuse of notation, also denote it $M$. By assumption, $H(\ell^2)$ contains an element of the form $\mathrm{id}_{2g} + M + \ell V$ for some $V \in \operatorname{Mat}_{2g \times 2g}(\bz / \ell \mathbb Z)$. This means that $H(\ell^2)$ contains \begin{align*} (\mathrm{id}_{2g} + M + \ell V)^\ell &\equiv \mathrm{id}_{2g} + \ell(M + \ell V) + \binom{\ell}{2} (M + \ell V)^2 + \cdots + \binom{\ell}{\ell - 1} (M + \ell V)^{\ell - 1} + (M + \ell V)^\ell \\ &\equiv \mathrm{id}_{2g} + \ell M \pmod{\ell^2}, \end{align*} where the last step above relies crucially upon the assumption that $\ell \geq 5$. We conclude that $H(\ell^2)$ contains $\mathrm{id}_{2g} + \ell M$ for every $M$ satisfying the above conditions. Taking $A = C = 0$, we see that $H(\ell^2)$ contains \[ \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c} 0 & B \\ \hline 0 & 0 \end{array}\right] \] for any symmetric matrix $B$. Similarly, taking $A = B = 0$ shows that $H(\ell^2)$ contains \[ \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c} 0 & 0 \\ \hline C & 0 \end{array}\right] \] for any symmetric matrix $C$. Taking $B = C = 0$ shows that $H(\ell^2)$ contains \[ \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c} A & 0 \\ \hline 0 & -A^T \end{array}\right] \] for any matrix $A$ with $A^2 = 0$. It is a standard fact that the span of such matrices $A$ is the space of trace zero matrices. Observe that $\operatorname{tr} A = 0$ is a single linear condition on $\mathfrak{sp}_{2g}(\bz / \ell \mathbb Z)$ that singles out a codimension-one linear subspace $W \subset \mathfrak{sp}_{2g}(\bz / \ell \mathbb Z)$, so that \begin{equation}\label{searstowerprom} \mathrm{id}_{2g} + \ell W \subset H(\ell^2) \cap \ker ( \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) \twoheadrightarrow \on{Sp}_{2g}(\bz / \ell \mathbb Z)). \end{equation} Observe that if the inclusion in~\eqref{searstowerprom} were strict, then the right-hand side would be all of $\mathrm{id}_{2g} + \ell \cdot \mathfrak{sp}_{2g}(\bz / \ell \mathbb Z)$, and the desired result follows. Therefore, suppose the inclusion in~\eqref{searstowerprom} is an equality. Since $W$ has index $\ell$ in $\mathfrak{sp}_{2g}(\bz / \ell \mathbb Z)$, it follows that $H(\ell^2)$ has index $\ell$ in $\on{Sp}_{2g}(\bz / \ell^2 \mathbb Z)$. We obtain a surjection \[ \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) \twoheadrightarrow \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) / H(\ell^2) \cong \bz / \ell \mathbb Z, \] which contradicts Proposition~\ref{corollary:finite-commutator}. \end{proof} \subsection{Lifting from $4$ to $8$} \label{booyah} \begin{lemma} \label{proposition:lifting-4} Fix $g \ge 2$. Then $H(4) = \on{Sp}_{2g}(\bz / 4 \mathbb Z)$ implies $H(8) = \on{Sp}_{2g}(\bz / 8 \mathbb Z)$. \end{lemma} \begin{proof} We modify the proof of Lemma~\ref{proposition:lifting-l}. As in that proof, we consider a matrix \[ M = \left[\begin{array}{c|c} A & B \\ \hline C & -A^T \end{array}\right] \in \mathfrak{sp}_{2g}(\bz / 2 \mathbb Z) \] with the property that $\mathrm{id}_{2g} + 2M$ lies in $\on{Sp}_{2g}(\bz / 4 \mathbb Z)$ and $M^2 = 0$. This time, however, the first condition automatically holds because \begin{align*} (\mathrm{id}_{2g} + 2M)^T \Omega_{2g}(\mathrm{id}_{2g} + 2M) &\equiv \Omega_{2g} + 2(M^T \Omega_{2g} + \Omega_{2g}M) + 4M^T \Omega_{2g}M \\ &\equiv \Omega_{2g} \pmod{4}. \end{align*} Nevertheless, note that the second condition is again satisfied whenever $A^2 = 0$ and two of $A, B,$ and $C$ are zero. Choose an arbitrary lift of $M$ in $\operatorname{Mat}_{2g \times 2g}(\bz / 4 \mathbb Z)$, and by abuse of notation also refer to it as $M$. By assumption, $H(8)$ contains an element of the form $\mathrm{id}_{2g} + 2M + 4V$ for some $V \in \operatorname{Mat}_{2g \times 2g}(\bz / 2 \mathbb Z)$, which means that $H(8)$ contains \begin{align*} (\mathrm{id}_{2g} + 2M + 4V)^2 &\equiv \mathrm{id}_{2g} \mathbin{+} 4M \pmod{8}. \end{align*} Taking $W \subset \mathfrak{sp}_{2g}(\bz / 2 \mathbb Z)$ to be the trace-zero subspace as before, this implies that \[ \mathrm{id}_{2g} \mathbin{+} 4 \cdot W \subset H(8) \cap \ker (\on{Sp}_{2g}(\bz / 8 \mathbb Z) \twoheadrightarrow \on{Sp}_{2g}(\bz / 4 \mathbb Z) ). \] If the inclusion were strict, again the right-hand side would contain the kernel of reduction, so the desired result follows. Therefore, suppose the inclusion is an equality, so that $H(8)$ has index $2$ in $\on{Sp}_{2g}(\bz / 8 \mathbb Z)$. If $g \ge 3$, Proposition~\ref{corollary:finite-commutator} tells us that $\on{Sp}_{2g}(\bz / 8 \mathbb Z)^{\ab}$ is trivial, which is a contradiction. If $g = 2$, the same proposition tells us that $H(8) = [\on{Sp}_4(\mathbb Z/ 8 \mathbb Z),\on{Sp}_4(\mathbb Z/8\mathbb Z)]$, so since the image of this commutator under the abelianization map \[ \on{Sp}_{4}(\bz / 8 \mathbb Z) \twoheadrightarrow \on{Sp}_{4}(\bz / 2 \mathbb Z) \twoheadrightarrow \mathbb Z/2\mathbb Z \] is trivial, $H(8)$ cannot surject onto $\on{Sp}_{4}(\bz / 2\mathbb Z)$, which is again a contradiction. \end{proof} \subsection{Lifting from 2 to 4 and from 3 to 9}\label{thisishisbananas} \begin{proposition} \label{proposition:lifting-2} Fix $g \ge 2$. The following statements hold: \begin{enumerate} \item Take $\ell = 2$. Then $H(2) = \on{Sp}_{2g}(\bz / 2 \mathbb Z)$ implies that $H(4) = \on{Sp}_{2g}(\bz / 4\mathbb Z)$. \item Take $\ell = 3$. Then $H(3) = \on{Sp}_{2g}(\bz / 3 \mathbb Z)$ implies that $H(9) = \on{Sp}_{2g}(\bz / 9 \mathbb Z)$. \end{enumerate} \end{proposition} \subsubsection*{Idea of Proof} The argument will proceed by induction on $g$. We start by verifying the base case $g = 2$ in Lemma~\ref{lemma:base-case-low-lifting}. We then inductively assume this holds for $g-1$ and prove it for $g$. We use the inductive hypothesis to construct a particular element lying in $H(\ell^2)$ in Lemma~\ref{lemma:particular-matrices-in-preimage}. Then, we use Lemma~\ref{lemma:code} to show that conjugates of this particular element generate $\mathfrak{sp}_{4}(\mathbb Z/ \ell^2 \mathbb Z)$, embedded as in ~\eqref{abovesubspace}. We finally translate around this copy of $\mathfrak{sp}_{4}(\mathbb Z/ \ell^2 \mathbb Z)$ to obtain that $H(\ell^2) = \on{Sp}_{2g}(\mathbb Z/\ell^2 \mathbb Z)$. The following lemma deals with the base case: \begin{lemma} \label{lemma:base-case-low-lifting} Proposition~\ref{proposition:lifting-2} holds in the case that $g = 2$. \end{lemma} \begin{proof} Let $\ell \in \{2,3\}$. We proceed by induction on $g$. The following \texttt{Magma} code verifies the base cases for $g = 2$, that the only subgroup of $\on{Sp}_4(\mathbb Z/ \ell^2 \mathbb Z)$ surjecting onto $\on{Sp}_4(\mathbb Z/ \ell \mathbb Z)$ is all of $\on{Sp}_4(\mathbb Z/\ell^2 \mathbb Z)$. \vspace*{0.1in} \begin{adjustwidth}{0.5in}{0in} \begin{alltt} for l in [2,3] do Z := Integers(); G := GL(4,quo<Z|l*l>); A := elt<G | 1,0,0,0, 1,-1,0,0, 0,0,1,1, 0,0,0,-1>; B := elt<G| 0,0,-1,0, 0,0,0,-1, 1,0,1,0, 0,1,0,0>; H := sub<G|A,B>; maximals := SubgroupClasses(H: Al := "Maximal"); S := quo<Z|l>; grp, f := ChangeRing(G, S); for H in maximals do if #f(H{\`{}}subgroup) eq #Sp(4,l) then assert false; end if; end for; end for; \end{alltt} \end{adjustwidth} \vspace*{0.1in} This concludes the proof of Lemma~\ref{lemma:base-case-low-lifting}. \end{proof} In order to handle the inductive step of Proposition~\ref{proposition:lifting-2}, we first introduce some notation. Let \( \phi_{\ell,2g} \colon \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) \twoheadrightarrow \on{Sp}_{2g}(\bz / \ell \mathbb Z) \) denote the usual reduction map. Our next aim is to define the maps $\pi, \iota_\ell$ in the diagram \begin{equation}\label{roarrrr} \begin{tikzcd}[row sep=tiny] & \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz/ \ell \mathbb Z)) \ar[hookrightarrow]{r} \ar[twoheadrightarrow]{dd} \ar[twoheadrightarrow, swap]{ld}{\pi} & \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) \ar[twoheadrightarrow]{dd}{\phi_{\ell,2g}} \\ \on{Sp}_{2g-2}(\bz / \ell^2 \mathbb Z) \ar[twoheadrightarrow, swap]{rd}{\phi_{\ell,2g-2}} \\ & \on{Sp}_{2g-2}(\bz/\ell \mathbb Z) \ar[hookrightarrow, swap]{r}{\iota_\ell} & \on{Sp}_{2g}(\bz/\ell \mathbb Z) \end{tikzcd} \end{equation} The map $\iota_\ell$ will be defined in~\eqref{iota-definition} and the map $\pi$ will be defined in~\eqref{pi-definition} For the present purpose, it is convenient to use a different $\Omega$-matrix, which we shall denote by $J_{2g}$, in the definition of symplectic group $\on{Sp}_{2g}$ and its Lie algebra $\mathfrak{sp}_{2g}$. Inductively define \[ J_2 \defeq \left[\begin{array}{cc} 0 & 1 \\ -1 & 0 \end{array}\right] \quad \text{and} \quad J_{2g} \defeq \left[\begin{array}{c|c} J_{2g-2} & 0 \\ \hline 0 & J_2 \end{array}\right]. \] The matrix $J_{2g}$ is a block-diagonal matrix with each block being a copy of $J_2$. Now let $M \in \on{Sp}_{2g-2}(\bz / \ell^2 \mathbb Z)$. The map \[ M \mapsto \left[\begin{array}{c|c} M & 0 \\ \hline 0 & \mathrm{id}_2 \end{array}\right] \] gives an inclusion \[ \iota_{\ell^2} \colon \on{Sp}_{2g-2}(\bz /\ell^2 \mathbb Z) \hookrightarrow \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z) \] which, when taken modulo $\ell$, reduces to an inclusion \begin{align} \label{iota-definition} \iota_{\ell} \colon \on{Sp}_{2g-2}(\bz / \ell \mathbb Z) \hookrightarrow \on{Sp}_{2g}(\bz / \ell \mathbb Z). \end{align} Thus, the group $\phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z))$ (see the diagram in~\eqref{roarrrr}) consists of matrices satisfying \[ \left[\begin{array}{c|c} A_{(2g-2) \times (2g-2)} & B_{(2g-2) \times 2} \\ \hline C_{2\times(2g-2)} & D_{2 \times 2} \end{array}\right] \equiv \left[\begin{array}{c|c} M & 0 \\ \hline 0 & \mathrm{id}_{2} \end{array}\right] \pmod{\ell} \] where $M \in \on{Sp}_{2g-2}(\bz / \ell \mathbb Z)$ and the blocks $A,B,C,D$ have the indicated sizes. For two such matrices, we have \begin{align*} \left[\begin{array}{c|c} A_1 & B_1 \\ \hline C_1 & D_1 \end{array}\right] \cdot \left[\begin{array}{c|c} A_2 & B_2 \\ \hline C_2 & D_2 \end{array}\right] &\equiv \left[\begin{array}{c|c} A_1A_2 + B_1C_2 & A_1B_2 + B_1D_2 \\ \hline C_1A_2 + D_1C_2 & C_1B_2 + D_1D_2 \end{array}\right] \\ &\equiv \left[\begin{array}{c|c} A_1A_2 & A_1B_2 + B_1D_2 \\ \hline C_1A_2 + D_1C_2 & D_1D_2 \end{array}\right] \pmod{\ell^2}, \end{align*} where the last step follows because $B_1C_2 \equiv C_1B_2 \equiv 0 \pmod{\ell^2}$. Therefore, the map \begin{align} \label{pi-definition} \pi \colon \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z)) \twoheadrightarrow \on{Sp}_{2g-2}(\bz / \ell^2 \mathbb Z) \quad \text{sending} \quad \left[\begin{array}{c|c} A & B \\ \hline C & D \end{array}\right] \mapsto A \end{align} is a group homomorphism. This completes the definition of the maps $\iota_\ell$ and $\pi$, and it is apparent that the diagram in~\eqref{roarrrr} commutes. With this notation set, we now continue our proof of Proposition~\ref{proposition:lifting-2}. In Lemma~\ref{lemma:particular-matrices-in-preimage}, we show via explicit matrix multiplication that one of two particular matrices lies in $H(\ell^2)$. \begin{lemma} \label{lemma:particular-matrices-in-preimage} Suppose that \( \pi(H(\ell^2) \cap \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z))) = \on{Sp}_{2g-2}(\bz / \ell^2 \mathbb Z) \) and $H(\ell) = \on{Sp}_{2g}(\mathbb Z/\ell \mathbb Z)$. Then, defining \begin{align} \label{droptotheground} \matone&\defeq \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c} \begin{array}{cccc} -1 & -1 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \end{array} & 0_{4 \times (2g-4)} \\ \hline 0_{(2g-4) \times 4} & 0_{(2g-4) \times (2g-4)} \end{array}\right] \quad \text{and} \\ \label{pushofabutton} \mattwo& \defeq \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c} \begin{array}{cccc} -1 & -1 & 1 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ 0 & -1 & 0 & 0 \end{array} & 0_{4 \times (2g-4)} \\ \hline 0_{(2g-4) \times 4} & 0_{(2g-4) \times (2g-4)} \end{array}\right], \end{align} we have that either $\matone$ or $\mattwo$ lies in $H(\ell^2)$. \end{lemma} \begin{proof} Since we are assuming \( \pi(H(\ell^2) \cap \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z))) = \on{Sp}_{2g-2}(\bz / \ell^2 \mathbb Z) \), \mbox{it follows that} \[ \mathrm{id}_{2g-2} + \ell \cdot \left[\begin{array}{c|c} \begin{array}{cc} 0 & 0 \\ 1 & 0 \end{array} & 0_{2 \times (2g-4)} \\\hline 0_{(2g-4) \times 2} & 0_{(2g-4) \times (2g-4)} \end{array}\right] \] lies in $\pi(H(\ell^2) \cap \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z)))$. It follows that $H(\ell^2)$ contains an element of the form $M = \mathrm{id}_{2g} + \ell U$ for \[ U = \left[\begin{array}{c|c|c} \begin{array}{cc} 0 & 0 \\ 1 & 0 \end{array} & 0_{2 \times (2g-4)} & A_{2 \times 2} \\ \hline 0_{(2g-4) \times 2} & 0_{(2g-4) \times (2g-4)} & B_{(2g-4) \times 2} \\ \hline A'_{2 \times 2} & B'_{2 \times (2g-4)} & C_{2 \times 2} \end{array}\right], \] where there is a linear relation between $A_{2 \times 2}$ and $A'_{2 \times 2}$, as well as a linear relation between $B_{(2g-4) \times 2}$ and $B'_{2 \times (2g-4)}$, imposed by the symplectic constraint $M^T J_{2g} M = J_{2g}$. For any $M \in H(\ell^2)$, the group $H(\ell^2)$ also contains \[ M^{-1}(\mathrm{id}_{2g} + \ell U)M = \mathrm{id}_{2g} + \ell M^{-1} U M, \] where the right-hand-side only depends on the reduction of $M$ modulo $\ell$. Since $H(\ell^2)$ surjects onto $\on{Sp}_{2g}(\bz / \ell \mathbb Z)$, the matrix $M \pmod{\ell}$ ranges over all elements of $\on{Sp}_{2g}(\bz / \ell \mathbb Z)$. With this in mind, take a matrix $M$ given by \begin{align*} M &\defeq \left[\begin{array}{c|c|c} \begin{array}{cc} 1 & 1 \\ 0 & 1 \end{array} & 0_{2 \times (2g-4)} & 0_{2\times2} \\ \hline 0_{(2g-4) \times 2} & \mathrm{id}_{2g-4} & 0_{(2g-4) \times 2} \\ \hline 0_{2 \times 2} & 0_{2 \times (2g-4)} & \mathrm{id}_2 \end{array}\right] \pmod{\ell} \end{align*} so that $H(\ell^2)$ contains \begin{align} \label{conjugated-matrix-for-lifting} M^{-1}(\mathrm{id}_{2g} + \ell U) M = \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c|c} \begin{array}{cc} -1 & -1 \\ 0 & 1 \end{array} & 0_{2 \times (2g-4)} & \left[\begin{array}{cc} 0 & -1 \\ 0 & 0 \end{array}\right] \cdot A_{2 \times 2} \rule[-3.75ex]{0pt}{0pt} \\ \hline 0_{(2g-4) \times 2} & 0_{(2g-4) \times (2g-4)} & 0_{(2g-4) \times 2} \\ \hline \rule{0pt}{5ex} A'_{2 \times 2} \cdot \left[\begin{array}{cc} 0 & 1 \\ 0 & 0 \end{array}\right] & 0_{2 \times (2g-4)} & 0_{2 \times 2} \end{array}\right]. \end{align} Multiplying~\eqref{conjugated-matrix-for-lifting} on the left by \( (\mathrm{id}_{2g} + \ell U)^{-1} \equiv \mathrm{id}_{2g} - \ell U \pmod{\ell^2} \) shows that $H(\ell^2)$ contains \begin{align*} N &\defeq (\mathrm{id}_{2g} + \ell U)^{-1}M^{-1}(\mathrm{id}_{2g} + \ell U) M \\ &= \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c|c} \begin{array}{cc} -1 & -1 \\ 0 & 1 \end{array} & 0_{2 \times (2g-4)} & \left[\begin{array}{cc} 0 & -1 \\ 0 & 0 \end{array}\right] \cdot A_{2 \times 2} \rule[-3.75ex]{0pt}{0pt} \\ \hline 0_{(2g-4) \times 2} & 0_{(2g-4) \times (2g-4)} & 0_{(2g-4) \times 2} \\ \hline \rule{0pt}{5ex} A'_{2 \times 2} \cdot \left[\begin{array}{cc} 0 & 1 \\ 0 & 0 \end{array}\right] & 0_{2 \times (2g-4)} & 0_{2 \times 2} \end{array}\right]. \end{align*} Conjugating $N$ by any matrix of the form \[ P \defeq \left[\begin{array}{c|c|c} \mathrm{id}_2 & 0_{2 \times (2g-4)} & 0_{2 \times 2} \\ \hline 0_{(2g-4) \times 2} & \mathrm{id}_{2g-4} & 0_{(2g-4) \times 2} \\ \hline 0_{2 \times 2} & 0_{2 \times (2g-4)} & M_{2 \times 2} \end{array}\right] \pmod{\ell}, \] where $M_{2 \times 2} \in \on{Sp}_2(\mathbb Z/\ell^2 \mathbb Z)$, results in the matrix \begin{align*} N' &\defeq P^{-1}NP \\ &= \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c|c} \begin{array}{cc} -1 & -1 \\ 0 & 1 \end{array} & 0_{2 \times (2g-4)} & \left[\begin{array}{cc} 0 & -1 \\ 0 & 0 \end{array}\right] \cdot A_{2 \times 2} \cdot M_{2 \times 2} \rule[-3.75ex]{0pt}{0pt} \\ \hline 0_{(2g-4) \times 2} & 0_{(2g-4) \times (2g-4)} & 0_{(2g-4) \times 2} \\ \hline \rule{0pt}{5ex} M_{2 \times 2}^{-1} \cdot A'_{2 \times 2} \cdot \left[\begin{array}{cc} 0 & 1 \\ 0 & 0 \end{array}\right] & 0_{2 \times (2g-4)} & 0_{2 \times 2} \end{array}\right]. \end{align*} By choosing $M_{2 \times 2}$ judiciously, we may arrange that \[\left[\begin{array}{cc} 0 & -1 \\ 0 & 0 \end{array}\right] \cdot A_{2\times 2} \cdot M_{2 \times 2} \quad = \quad \left[\begin{array}{cc} 1 & 0 \\ 0 & 0 \end{array}\right] \quad \text{or} \quad \left[\begin{array}{cc} 0 & 0 \\ 0 & 0 \end{array}\right], \] depending on whether the bottom row of $A_{2 \times 2}$ is nonzero or zero, respectively. Upon conjugating by the matrix \[ \left[\begin{array}{c|c} \mathrm{id}_2 & 0_{2 \times (2g-2)} \\ \hline 0_{(2g-2) \times 2} & \begin{array}{c|c|c} 0_{2 \times 2} & 0_{2 \times (2g-6)} & \mathrm{id}_2 \\ \hline 0_{(2g-6) \times 2} & \mathrm{id}_{2g-6} & 0_{(2g-6) \times 2} \\ \hline \mathrm{id}_2 & 0_{2 \times (2g-6)} & 0_{2 \times 2} \end{array} \end{array}\right] \in \on{Sp}_{2g}(\bz / \ell \mathbb Z), \] we conclude that $H(\ell^2)$ contains either $\matone$ or $\mattwo$. \end{proof} Next, in Lemma~\ref{lemma:code}, we show that we can conjugate the matrices $\matone$ and $\mattwo$ from Lemma~\ref{lemma:particular-matrices-in-preimage} to obtain all of $\on{Sp}_4(\mathbb Z/\ell \mathbb Z)$. \begin{lemma}\label{lemma:code} Let $\ell = 2$ or $3$. Let $\overline{\matone}$ and $\overline{\mattwo}$ denote the upper-left $4 \times 4$ blocks of $\matone$ and $\mattwo$, respectively. Then the sets \begin{align*} \left\{ M^{-1} \overline{\matone} M : M \in \on{Sp}_{4}(\bz / \ell \mathbb Z) \right\} \quad \text{and}\quad \left\{ M^{-1} \overline{\mattwo} M : M \in \on{Sp}_{4}(\bz / \ell \mathbb Z) \right\} \end{align*} are both spanning sets for $1 + \ell \cdot \mathfrak{sp}_{4}(\bz / \ell \bz)$. \end{lemma} \begin{proof} The following {\tt Magma} code verifies that each of the sets defined in the lemma statement span $\mathfrak{sp}_{2g}(\bz / \ell \mathbb Z)$. \vspace*{0.1in} \begin{adjustwidth}{0.5in}{0in} \begin{alltt} for l in [2, 3] do Z := Integers(); G := GL(4,quo<Z|l*l>); A := elt<G| 1,0,0,0, 1,1,0,0, 0,0,1,0, 0,0,0,1>; B := elt<G| 1,1,0,0, 0,1,1,0, 1,1,1,1, 1,1,0,1>; H := sub<G|A,B>; grp, f := ChangeRing(G, quo<Z|l>); Lie := Kernel(f) meet H; M := elt<G| 1-l,-l,0,0, 0,1+l,0,0, 0,0,1,0, 0,0,0,1>; N := elt<H| 1-l,-l,l,0, 0,1 + l,0,0, 0,0,1,0, 0,-l,0,1>; #sub<H|Conjugates(H,M)> eq #Lie; #sub<H|Conjugates(H,N)> eq #Lie; end for; \end{alltt} \end{adjustwidth} \vspace*{0.1in} This concludes the proof of Lemma~\ref{lemma:code} \end{proof} We now have the tools to prove Proposition~\ref{proposition:lifting-2}. \begin{proof}[Proof of Proposition~\ref{proposition:lifting-2}] The base case $g = 2$ is the content of Lemma~\ref{lemma:base-case-low-lifting}. Now take $g \ge 3$, and suppose the result holds for $g - 1$. We shall now use the inductive hypothesis to show that $\ker \phi_{\ell,2g} \subset H(\ell^2)$, which would imply that $H(\ell^2) = \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z)$. Since $H(\ell^2)$ surjects onto $\on{Sp}_{2g}(\bz / \ell \mathbb Z)$, we have that the group $H(\ell^2) \cap \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z))$ surjects onto $\on{Sp}_{2g-2}(\bz / \ell \mathbb Z)$, and therefore so does the group \( \pi(H(\ell^2) \cap \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z))). \) By the inductive hypothesis, \( \pi(H(\ell^2) \cap \phi_{\ell,2g}^{-1}(\on{Sp}_{2g-2}(\bz / \ell \mathbb Z))) =\on{Sp}_{2g-2}(\bz / \ell^2 \mathbb Z), \) so upon applying Lemma~\ref{lemma:particular-matrices-in-preimage}, we deduce that either $\matone$ or $\mattwo$ lies in $H(\ell^2)$. Now, conjugating by matrices of the form \[ \left[\begin{array}{c|c}M_{4 \times 4} & 0_{4 \times (2g-4)} \\ \hline 0_{(2g-4) \times 4} & \mathrm{id}_{2g-4} \end{array}\right]\pmod{\ell} \] where $M_{4 \times 4} \in \on{Sp}_{4}(\bz / \ell \mathbb Z)$ serves to conjugate the upper-left $4 \times 4$ of $\matone$ and $\mattwo$ by $M_{4 \times 4}$. By Lemma~\ref{lemma:code}, we have that $\mathfrak{sp}_{4}(\mathbb Z/\ell^2 \mathbb Z) \subset H(\ell^2)$, embedded as \begin{equation}\label{abovesubspace} \mathrm{id}_{2g} + \ell \cdot \left[\begin{array}{c|c} \mathfrak{sp}_4(\bz / \ell \mathbb Z) & 0_{4 \times (2g-4)} \\ \hline 0_{(2g-4) \times 4} & 0_{(2g-4) \times (2g-4)} \end{array}\right]. \end{equation} Construct $Q \in \on{Sp}_{2g}(\bz / \ell \mathbb Z)$ by taking any $g \times g$ permutation matrix and replacing each $1$ with an $\mathrm{id}_2$-block. Conjugating the subspace in~\eqref{abovesubspace} by various such $Q$ shows that $H(\ell^2)$ contains $\ker \phi_{\ell,2g}$. This can be seen by a straightforward argument involving choosing a basis for $\mathfrak{sp}_{2g}(\mathbb Z/\ell \mathbb Z)$ whose elements are elementary matrices or sums of \mbox{two elementary matrices.} \end{proof} \subsection{Finishing the Proof} We are now in position to complete the proof of Theorem~\ref{jamesdreamcometrue} \begin{proof}[Proof of Theorem~\ref{jamesdreamcometrue}] We split into three cases: \begin{itemize} \item Suppose $\ell \ge 5$. Then Lemma~\ref{proposition:lifting-l} implies $H(\ell^2) = \on{Sp}_{2g}(\bz / \ell^2 \mathbb Z)$. \item Suppose $\ell = 3$. Then Lemma~\ref{proposition:lifting-2} implies $H(9) = \on{Sp}_{2g}(\bz / 9 \mathbb Z)$. \item Suppose $\ell = 2$. Then Lemma~\ref{proposition:lifting-2} implies $H(4) = \on{Sp}_{2g}(\bz / 4 \mathbb Z)$, and Lemma~\ref{proposition:lifting-4} implies $H(8) = \on{Sp}_{2g}(\bz / 8 \mathbb Z)$. \end{itemize} Combining the above results with Lemma~\ref{proposition:lifting-l-squared} gives the desired conclusion. \end{proof} \section{Introduction} Let $g$ be a positive integer, and for a ring $R$, denote by $\mathrm{Sp}_{2g}(R)$ the group of $2g \times 2g$ symplectic matrices over $R$. Let $\mathbb Z_\ell$ denote the ring of $\ell$-adic integers, and consider the natural projection map $\on{Sp}_{2g}(\mathbb Z_\ell) \to \on{Sp}_{2g}(\mathbb Z/\ell \mathbb Z)$. In this paper, we show that when $g > 1$, there are no proper closed subgroups of $\on{Sp}_{2g}(\mathbb Z_\ell)$ that surject via this projection map onto all of $\on{Sp}_{2g}(\mathbb Z/\ell \mathbb Z)$. The case $g = 1$, in which $\on{Sp}_2 = \SL_2$, is well-understood. Indeed, as proven in~\cite[Lemma 3, Section IV.3.4]{serre1989abelian}, if $\ell \geq 5$ and $H_\ell \subset \SL_2(\mathbb Z_\ell)$ is a closed subgroup that surjects onto $\SL_2(\mathbb Z/\ell \mathbb Z)$, then $H_\ell = \SL_2(\mathbb Z_\ell)$. The corresponding result for $\ell \in \{2,3\}$ simply does not hold: in each of these cases, there are nontrivial subgroups of $\on{Sp}_{2g}(\mathbb Z / \ell^2 \mathbb Z)$ that surject onto $\SL_2(\mathbb Z/\ell \mathbb Z)$. See~\cite[Section IV.3.4, Exercises 1 -- 3]{serre1989abelian} for exercises outlining a proof, and also see for more comprehensive descriptions~\cite{dontyouknowit} for the case $\ell = 2$ and~\cite{onething} for the case $\ell = 3$. The objective of the present article is to generalize~\cite[Lemma 3, Section IV.3.4]{serre1989abelian}, to hold for all $g \geq 2$. Our main theorem is \mbox{stated as follows:} \begin{theorem}\label{jamesdreamcometrue} Let $g \ge 2$, let $\ell$ be a prime number, and let $H_\ell \subset \on{Sp}_{2g}(\bz_\ell)$ be a closed subgroup. If the mod-$\ell$ reduction of $H_\ell$ equals all of $\on{Sp}_{2g}(\bz / \ell \mathbb Z)$, then $H_\ell = \on{Sp}_{2g}(\bz_\ell)$, and \mbox{in particular, the} mod-$\ell^k$ reduction of $H_\ell$ equals all of $\on{Sp}_{2g}(\mathbb Z/\ell^k \mathbb Z)$ for each \mbox{positive integer $k$.} \end{theorem} \begin{remark} A more general version of Theorem~\ref{jamesdreamcometrue} is proven for a large class of semisimple Lie groups $G$ in \cite[Theorem B]{weigel:on-the-profinite-completion-of-arithmetic-groups-of-split-type} (except for the case that $g = 3$ and $\ell = 2$) and also in ~\cite[Theorem 1.3]{vasiu2003surjectivity}. In the present article, we provide an elementary and self-contained proof for the special case $G = \on{Sp}_{2g}$. In particular, our inductive method circumvents the use of Lie theory, and is therefore suitable for somewhat more general groups (e.g.\ certain finite-index subgroups of matrix groups over $\bz_\ell$) which arise in the study of Galois representations associated to abelian varieties, cf.\ \cite{landesman-swaminathan-tao-xu:hyperelliptic-curves}. \end{remark} \begin{remark} To give a typical application, one can directly use Theorem~\ref{jamesdreamcometrue} to reduce the problem of checking that the $\ell$-adic Galois representation associated to an abelian variety has maximal image to the simpler problem of checking that the mod-$\ell$ reduction has maximal image. Indeed, the conclusion of Theorem~\ref{jamesdreamcometrue} has been applied many times in the study of Galois representations, such as in~\cite[Proof of Lemma 2.4]{zywina2015example}, ~\cite[Proof of Theorem 8]{Hui2016}, ~\cite[Proof of Corollary 3.5]{achter-pries:integral-monodromy-of-hyperelliptic-and-trielliptic-curves},~\cite[Proof of Lemma 5.1]{achter-pries:monodromy-of-p-rank-strat-of-the-moduli-space-of-curves},~\cite[p.~467]{scoopdedoo}, as well in the authors' own papers~\cite{landesman-swaminathan-tao-xu:rational-families} and~\cite{landesman-swaminathan-tao-xu:hyperelliptic-curves}. \end{remark} The rest of this paper is organized as follows. In Section~\ref{subsection:stimpy}, we introduce the basic definitions and properties of the symplectic group. Next, in Section~\ref{subsection:abelianizations}, we compute the commutator subgroups of $\on{Sp}_{2g}(\mathbb Z_\ell)$ and $\on{Sp}_{2g}(\mathbb Z/\ell^k \mathbb Z)$ for every prime number $\ell$ and positive integer $k$. Finally, in Section~\ref{doyoulift}, we prove Theorem~\ref{jamesdreamcometrue}. \section{Background on Symplectic Groups} In this section, we first detail the basic definitions and properties of symplectic groups, and then proceed to prove a number of lemmas that are used in our proof of Theorem~\ref{jamesdreamcometrue}. \subsection{Symplectic Groups}\label{subsection:stimpy} Fix a commutative ring $R$, let $\operatorname{Mat}_{2g \times 2g}(R)$ denote the space of $2g \times 2g$ matrices with entries in $R$, and let $\Omega_{2g} \in \operatorname{Mat}_{2g \times 2g}(R)$ be defined by $$\Omega_{2g} \defeq \left[\begin{array}{c|c} 0 & \mathrm{id}_g \\ \hline -\mathrm{id}_g & 0\end{array}\right],$$ where $\mathrm{id}_g$ denotes the $g \times g$ identity matrix. We define the symplectic group $\on{Sp}_{2g}(R)$ as the set of $M \in \SL_{2g}(R)$ so that $M^T \Omega_{2g} M = \Omega_{2g}$. In the proof of Theorem~\ref{jamesdreamcometrue}, we will make heavy use of the ``Lie algebra'' $\mathfrak{sp}_{2g}(R)$, which is defined by \begin{align*} \mathfrak{sp}_{2g}(R) &\defeq \{M \in \operatorname{Mat}_{2g \times 2g}(R) : M^T \Omega_{2g} + \Omega_{2g} M = 0 \}. \end{align*} It is easy to see that $M^T \Omega_{2g} + \Omega_{2g}M = 0$ is equivalent to $M$ being a block matrix with $g \times g$ blocks of the form \[ M = \left[\begin{array}{c|c} A & B \\ \hline C & -A^T \end{array}\right], \] where $B$ and $C$ are symmetric. In what follows, we specialize to studying symplectic groups over $R = \mathbb Z$, $R = \mathbb Z_{\ell}$, or $R = \mathbb Z / \ell^k \mathbb Z$ for $\ell$ a prime number and $k$ a positive integer. We will adhere to the following notational conventions: \begin{itemize} \item Let $H_{\ell} \subset \on{Sp}_{2g}(\mathbb Z_\ell)$ be a closed subgroup. \item Let $H(\ell^k) \subset \on{Sp}_{2g}(\mathbb Z/\ell^k \mathbb Z)$ be the mod-$\ell^k$ reduction of $H_\ell$. \item Notice that the map $S \mapsto \mathrm{id}_{2g} + \ell^k S$ gives an isomorphism of groups $$\mathfrak{sp}_{2g}(\mathbb Z/ \ell \mathbb Z) \simeq \ker(\on{Sp}_{2g}(\mathbb Z/\ell^{k+1} \mathbb Z) \to \on{Sp}_{2g}(\mathbb Z/ \ell^k \mathbb Z))$$ for every $k \geq 1$. We will use the Lie algebra notation $\mathfrak{sp}_{2g}(\mathbb Z/\ell\mathbb Z)$ to denote the above kernel when we want to think of its elements additively, and we will use the kernel notation when we want to view its elements multiplicatively. \item For any group $G$, let $[G,G]$ be its commutator subgroup, and let $G^{\ab} = G/[G,G]$ be its abelianization. \end{itemize} \subsection{Commutators} \label{subsection:abelianizations} We shall now compute the abelianizations of $\on{Sp}_{2g}(\bz_\ell)$ and $\on{Sp}_{2g}(\bz / \ell^k \mathbb Z)$ for every integer $g \geq 2$, prime number $\ell$, and positive integer $k$. It will first be convenient for us to compute the abelianization $\on{Sp}_{2g}(\mathbb Z)^{\ab}$. \begin{lemma} \label{theorem:commutator-sp-z} The group $\on{Sp}_{2g}(\bz)^{\ab}$ is trivial when $g \geq 3$ and is isomorphic to \mbox{$\mathbb Z/2 \mathbb Z$ when $g = 2$.} \end{lemma} \begin{proof} The case $g \geq 3$ follows from~\cite[Remark, p.\ 123]{bass1967solution}, so it only remains to deal with the case $g = 2$. By~\cite[Satz]{bender1980presentation}, $\on{Sp}_{4}(\mathbb Z)$ has two generators $K$ and $L$ that satisfy several relations, three of which are given as follows: \begin{align*} K^2 &= \mathrm{id}_{2g}, \\ L^{12} &= \mathrm{id}_{2g}, \\ (K\cdot L^5)^5 &= (L^6\cdot K\cdot L^5\cdot K\cdot L^7\cdot K)^2. \end{align*} By the universal property of the abelianization, we have that $\on{Sp}_4(\mathbb Z)^{\ab}$ is a quotient of the rank-$2$ free abelian group $K \mathbb Z \oplus L \mathbb Z$ generated by $K$ and $L$. Thus, from the aforementioned multiplicative relations between $K$ and $L$ in $\on{Sp}_4(\mathbb Z)$, we obtain the following additive relations in the abelianization \begin{align*} 2K &= 0, \\ 12L &= 0, \\ 5(K + 5L) &= 2\cdot (6L + K + 5L + K + 7L + K). \end{align*} Substituting the first two relations above into the third relation, we find $L = K$, which implies that $\on{Sp}_4(\mathbb Z)^{\ab}$ is a quotient of $(K \mathbb Z \oplus L \mathbb Z)/(2K, K - L) \simeq \mathbb Z / 2\bz$. It remains to show that $\on{Sp}_{4}(\mathbb Z)$ maps surjectively onto $\mathbb Z/2\bz$. Postcomposing the surjection $\on{Sp}_{4}(\mathbb Z) \twoheadrightarrow \on{Sp}_{4}(\mathbb Z/2 \mathbb Z)$ with the isomorphism $\on{Sp}_{4}(\mathbb Z/2 \mathbb Z) \simeq S_6$ by \cite[3.1.5]{omeara1978symplectic} and then applying the sign map $S_6 \twoheadrightarrow \mathbb Z/2 \mathbb Z$ \mbox{yields the desired result.} \end{proof} \begin{remark} Let $\widehat{\mathbb Z}$ denote the profinite completion of $\mathbb Z$. It follows immediately from Lemma~\ref{theorem:commutator-sp-z}, together with the fact if $G_1$ and $G_2$ are groups with $\widehat{G_1} \simeq \widehat{G_2}$ then $\widehat{G_1^{\ab}} \simeq \widehat{G_2^{\ab}}$, that the group $\on{Sp}_{2g}(\widehat {\mathbb Z})^{\ab}$ is trivial for $g \geq 3$ and is isomorphic to \mbox{$\mathbb Z / 2 \mathbb Z$ when $g = 2$. \end{remark} Using Lemma~\ref{theorem:commutator-sp-z}, we can now compute the abelianizations of all aforementioned groups. \begin{proposition} \label{corollary:finite-commutator} We have the following results: \begin{enumerate} \item The group $\on{Sp}_{2g}(\mathbb Z_\ell)^{\ab}$ is trivial except when $g = \ell = 2$, in which case it is isomorphic to $\mathbb Z / 2 \mathbb Z$. \item Let $k \ge 1$. The group $\on{Sp}_{2g}(\mathbb Z/\ell^k \mathbb Z)^{\ab}$ is trivial except when $g = \ell = 2$, in which case it is isomorphic to $\mathbb Z / 2 \mathbb Z$. \end{enumerate} \end{proposition} \begin{proof} We first verify Statement (a). For $g \ge 3$, since $\on{Sp}_{2g}(\widehat{\mathbb Z})^{\ab}$ is trivial, and since we have a surjection $\on{Sp}_{2g}(\wh{\bz}) \twoheadrightarrow \on{Sp}_{2g}(\mathbb Z_\ell)$, the result follows immediately. Now take $g = 2$. First, we have surjections $\on{Sp}_4(\widehat{\mathbb Z}) \twoheadrightarrow \on{Sp}_4(\mathbb Z_2) \twoheadrightarrow \on{Sp}_4(\mathbb Z/2 \mathbb Z) \cong S_6$. Since the former and the latter have abelianizations isomorphic to $\mathbb Z/2 \mathbb Z$, using Lemma~\ref{theorem:commutator-sp-z}, it follows that $\on{Sp}_4(\mathbb Z_2)^{\ab} \cong \mathbb Z/2 \mathbb Z$. Then, since we have \begin{align*} \mathbb Z/2 \mathbb Z \cong \on{Sp}_4(\widehat{\mathbb Z})^{\ab} \cong \on{Sp}_4(\mathbb Z_2)^{\ab} \times \prod_{\ell\neq 2} \on{Sp}_4(\mathbb Z_\ell)^{\ab} \cong \mathbb Z/2\mathbb Z \times \prod_{\ell\neq 2} \on{Sp}_4(\mathbb Z_\ell)^{\ab}, \end{align*} it follows that $\on{Sp}_4(\mathbb Z_\ell)^{\ab}$ is trivial for $\ell \neq 2$. Now, observe that Statement (b) follows from Statement (a): indeed, notice that the surjection $\on{Sp}_{2g}(\mathbb Z_\ell) \rightarrow \on{Sp}_{2g}(\mathbb Z/ \ell^k \mathbb Z)$ induces a surjection $\on{Sp}_{2g}(\mathbb Z_\ell)^{\ab} \rightarrow \on{Sp}_{2g}(\mathbb Z/ \ell^k \mathbb Z)^{\ab}$ and in the case of $\ell = g = 2$, $\on{Sp}_4(\mathbb Z/ 2^k \mathbb Z)^{\ab}$ is nontrivial as it surjects onto $\on{Sp}_4(\mathbb Z/ 2 \mathbb Z)^{\ab} \cong \mathbb Z/2 \mathbb Z$. \end{proof}
239e8d959d043fb03f69af06baec1abdb42be7f6
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In this paper, we study the fully nonlinear, possibly degenerate, elliptic partial differential equation (PDE) in a bounded domain \begin{equation}\label{DP}\tag{DP$_\lambda$} \begin{cases} \gl u(x)+F(x,Du(x),D^2u(x))=0 \ \ \text{ in }\ \gO,&\\[3pt] \BC \ \ \text{ on }\bry, \end{cases} \end{equation} where $\gl$ is a given positive constant which is often called a discount factor, and $F\mid \ol\gO\tim\R^n\tim\bS^n\to\R$ is a given continuous function. Here $\gO$ is a bounded domain (that is, open and connected set) in $\R^n$, $\bS^n$ denotes the space of $n\tim n$ real symmetric matrices, and $\BC$ represents a boundary condition (state constraint, Dirichlet, or Neumann boundary condition). The unknown here is a real-valued function $u$ on $\ol \gO$, and $Du$ and $D^2u$ denote the gradient and Hessian of $u$, respectively. We are always concerned with viscosity solutions of fully nonlinear, possibly degenerate, elliptic PDE, and the adjective ``viscosity'' is omitted henceforth. Associated with the vanishing discount problem for (DP$_\lambda$) is the following ergodic problem \begin{equation}\label{E}\tag{E} \begin{cases} F(x,Du(x),D^2u(x))=c \ \ \text{ in }\ \gO,&\\[3pt] \BC \ \ \text{ on }\bry. \end{cases} \end{equation} We refer the boundary-value problem (E), with a given constant $c$, as (E$_c$), while the unknown for the ergodic problem (E) is a pair of a function $u \in C(\ol \gO)$ and a constant $c \in \R$ such that $u$ is a solution of (E$_c$). When $(u,c) \in C(\ol \gO) \times \R$ is a solution of (E), we call $c$ a critical value (or an additive eigenvalue). Our main goal is to study the vanishing discount problem for (DP$_\lambda$), that is, for solutions $v^\lambda$ of (DP$_\lambda$) with $\lambda>0$, we investigate the asymptotic behavior of $\{v^\lambda\}$ as $\lambda \to 0$. The particular question we want to address here is whether the {\it whole family} $\{v^\gl\}_{\gl>0}$ (after normalization) converges or not to a function in $C(\tt)$ as $\gl \to 0$. As the limiting equation (\ref{E}$_c$) is not strictly monotone in $u$ and has many solutions in general, proving or disproving such convergence result is challenging. The convergence results of the whole family $\{v^\gl\}_{\gl>0}$ were established for convex Hamilton-Jacobi equations in \cite{DFIZ} (first-order case in a periodic setting), \cite{AlAlIsYo} (first-order case with Neumann-type boundary condition), \cite{MiTr} (degenerate viscous case in a periodic setting). Recently, the authors \cite{IsMtTr1} have developed a new variational approach for this vanishing discount problem for fully nonlinear, degenerate elliptic PDEs, and proved the convergence of the whole family $\{v^\gl\}_{\gl>0}$ in the periodic setting. We develop the variational method introduced in \cite{IsMtTr1} further here to handle boundary value problems. Our goal is twofold. Firstly, we establish new representation formulas for $v^\gl$ as well as the critical value $c$ in the settings of the state constraint, Dirichlet, and Neumann boundary conditions. Secondly, we apply these representation formulas to show that $\{v^\gl\}_{\gl>0}$ (after normalization) converges as $\gl \to 0$. Let us make it clear that all of the results in this paper and the aforementioned ones require the convexity of $F$ in the gradient and Hessian variables. We refer to a forthcoming paper \cite{GMT} for convergence results of vanishing discount problems for some nonconvex first-order Hamilton-Jacobi equations. The main results, which, as mentioned above, consist of representation formulas and the convergence of $\{v^\gl\}_{\gl>0}$, are stated in Sections \ref{sec-s}, \ref{sec-d}, and \ref{sec-n} for the state constraint, Dirichlet, and Neumann problems, respectively. \subsection{Setting and Assumptions} We describe the setting and state the main assumptions here. Let $\cA$ be a non-empty, $\gs$-compact and locally compact metric space and $F\mid \ol\gO\tim\R^n\tim\bS^n\to\R$ be given by \begin{equation}\tag{F1}\label{F1} F(x,p,X)=\sup_{\ga\in\cA}(-\tr a(x,\ga)X-b(x,\ga)\cdot p-L(x,\ga)), \end{equation} where $a\mid \ol\gO\tim\cA \to\bS_+^n$, $b\mid \ol\gO\tim\cA\to\R^n$ and $L\mid \ol\gO\tim\cA\to\R$ are continuous. Here, $\bS_+^n$ denotes the set of all non-negative definite matrices $A\in\bS^n$, $\tr A$ and $p\cdot q$ designate the trace of $n\tim n$ matrix $A$ and the Euclidean inner product of $p,q\in\R^n$, respectively. Assume further that \begin{equation}\tag{F2}\label{F2} F\in C(\ol\gO\tim\R^n\tim\bS^n). \end{equation} It is clear under \eqref{F1} and \eqref{F2} that $F$ is degenerate elliptic in the sense that for all $(x,p,X)\in \ \ol \gO \tim\R^n\tim\bS^n$, if $Y\in\bS_+^n$, then $F(x,p,X+Y)\leq F(x,p,X)$ and that, for each $x\in\ol\gO$, the function $(p,X)\mapsto F(x,p,X)$ is convex on $\R^n\tim\bS^n$. The equations \eqref{DP}, (\ref{E}$_c$), with $F$ of the form \eqref{F1}, are called Bellman equations, or Hamilton-Jacobi-Bellman equations in connection with the theory of stochastic optimal control. In this viewpoint, the set $\cA$ is often called a control set or region. With the function $L$ in the definition of $F$, we define \[ \Phi^+:=\left\{\phi\in C(\tt\tim\cA)\mid \phi(x,\ga)=tL(x,\ga)+\chi(x) \ \text{for some} \ t>0, \, \chi\in C(\tt)\right\}. \] It is clear that $\Phi^+$ is a convex cone in $C(\tt \tim \cA)$. For $\phi= tL+\chi \in \Phi^+$, we define \[ F_\phi(x,p,X) = \sup_{\ga \in \cA} \left(- \tr a(x,\ga)X - b(x,\ga)\cdot p - \phi(x,\ga) \right). \] The form of $\phi$ allows us to compute that \begin{align*} F_\phi(x,p,X) &= \sup_{\ga \in \cA} \left(- \tr a(x,\ga)X - b(x,\ga)\cdot p - \phi(x,\ga) \right)\\ &=t \sup_{\ga \in \cA} \left(- \tr a(x,\ga) t^{-1}X - b(x,\ga)\cdot t^{-1}p - L(x,\ga) \right) - \chi(x)\\ &=tF(x,t^{-1}p,t^{-1}X) - \chi(x), \end{align*} which yields that $F_\phi \in C(\tt \times \R^n \times \bS^n)$ if we assume (F2). We note here that, except when $L(x,\ga)$ is independent of $\ga$, $\phi\in\Phi^+$ is represented uniquely as $\phi=tL+\chi$ for some $t>0$ and $\chi\in C(\tt)$. We often write $F[u]$ and $F_\phi[u]$ to denote the functions $x\mapsto F(x,Du(x),D^2u(x))$ and $x\mapsto F_\phi(x,Du(x),D^2u(x))$, respectively. The following are some further assumptions we need in the paper, and we put labels on these for convenience later. \begin{equation}\tag{CP$_{\rm loc}$}\label{CP} \left\{\text{ \begin{minipage}{0.83\textwidth} For any $\gl>0$, $\phi \in \Phi^+$, and open subset $U$ of $\gO$, if $v,\,w\in C(U)$ are a subsolution and a supersolution of $\gl u+F_\phi[u]=0$ in $U$, respectively, and $v \leq w$ on $\partial U$, then $v\leq w$ in $U$. \end{minipage} }\right.\end{equation} \begin{equation} \tag{EC}\label{EC} \left\{\text{ \begin{minipage}{0.85\textwidth} For $\gl>0$, let $v^\gl \in C(\tt)$ be a solution of \eqref{DP}. The family $\{v^\gl\}_{\gl>0}$ is equi-continuous on $\tt$. \end{minipage} }\right. \end{equation} \begin{equation}\tag{L}\label{L}\left\{ \text{ \begin{minipage}{0.85\textwidth} $L=+\infty\,$ at infinity,\ that is, for any $M\in\R$, there exists a compact subset $K$ of $\cA$ such that $L\geq M$ in $\tt\tim(\cA\setminus K)$. \end{minipage}} \right. \end{equation} We say that $L$ is coercive if condition \eqref{L} is satisfied, and, when $\cA$ is compact, condition \eqref{L} always holds. Some more assumptions are needed and stated in the upcoming sections according to the type of boundary conditions. \subsection*{Outline of the paper} In Section \ref{sec-pre}, we introduce some notation and preliminaries. We first consider the state constraint problem in Section \ref{sec-s} as it is the simplest one. The Dirichlet problem and the Neumann problem are studied in Section \ref{sec-d} and Section \ref{sec-n}, respectively. Finally, some examples are given in Section \ref{sec-ex}. \section{Notation and preliminaries} \label{sec-pre} First we present our notation. Given a metric space $E$, $\Lip(E)$ denotes the space of Lipschitz continuous functions on $E$. Also, let $C_{\rm c}(E)$ denote the space of continuous functions on $E$ with compact support. Let $\cR_E$, $\cR_E^+$ and $\cP_E$ denote the spaces of all Radon measures, all nonnegative Radon measures and Radon probability measures on $E$, respectively. For any function $\phi$ on $E$ integrable with respect $\mu\in\cR_E$, we write \[ \lan\mu,\phi\ran=\int_E \phi(x)\mu(d x). \] With the function $L$ from the definition of $F$, we set \[\begin{aligned} \Phi^+&\,:=\left\{tL+\chi\mid t>0, \, \chi\in C(\tt)\right\},\qquad \Psi^+:=\Phi^+\tim C(\bry),\\ \Psi^+(M)&\,:= \{(tL+\chi,\psi)\in\Psi^+\mid \|\chi\|_{C(\tt)}< tM, \|\psi\|_{C(\bry)}< tM\} \ \ \text{ for }M>0, \end{aligned}\] where $(t,\chi)$ in the braces above ranges over $(0,\,\infty)\tim C(\tt)$, and \[ \cR_L:=\{\mu\in\cR_{\bb}\mid L \text{ is integrable with respect to }\mu\}. \] Next, we give three basic lemmas related to the weak convergence of measures. Let $\cX$ be a $\gs$-compact, locally compact metric space and $f\mid \cX\to\R$ be a continuous function. Henceforth, when $f$ is bounded from below on $\cX$ and $\mu$ is a nonnegative Radon measure on $\cX$, we write \[ \int_\cX f(x)\mu(dx)=+\infty, \] to indicate that $f$ is not integrable with respect to $\mu$. We denote by $\lan\mu,f\ran$, for simplicity, the integral \[ \int_\cX f(x)\mu(dx). \] Assume that $f$ is coercive in the sense that $f=+\infty$ at infinity. In particular, $f$ is bounded below on $\cX$. \begin{lem}\label{basic-lsc} The functional $\mu\mapsto \lan \mu, f\ran$ is lower semicontinuous on $\cR_\cX^+$ in the topology of weak convergence of measures. \end{lem} \begin{proof} Let $\{\mu_j\}_{j\in\N}\subset\cR_{\cX}^+$ be a sequence converging to $\mu\in\cR^+_{\cX}$ weakly in the sense of measures. Select a compact set $K\subset\cX$ so that $f(x)>0$ for all $x\in\cX\setminus K$, and then a sequence $\{\chi_k\}_{k\in\N}\subset C_{\rm c}(\cX)$ so that $0\leq \chi_k\leq \chi_{k+1}\leq 1$ on $\cX$, $\chi_k=1$ on $K$ for all $k\in\N$, and $\chi_k \to \mathbf{1}_{\cX}$ pointwise as $k \to \infty$. Since $\chi_k f\in C_{\rm c}(\cX)$ and $\chi_k f\leq f$ on $\cX$, we have \[ \lan\mu,\chi_k f\ran =\lim_{j\to\infty}\lan\mu_j,\chi_kf\ran \leq\liminf_{j\to\infty}\lan\mu_j,f\ran \ \ \ \text{ for all }\ k\in\N. \] Hence, using the monotone convergence theorem if \[ \liminf_{j\to\infty}\lan\mu_j,f\ran<+\infty, \] we conclude that \[ \lan\mu,f\ran \leq\liminf_{j\to\infty}\lan\mu_j,f\ran. \qedhere \] \end{proof} \begin{lem} \label{basic-cpt} Let $r\in\R$ and $s>0$, and define \[ \cR^+_{f,r,s}=\{\mu\in\cR^+_\cX\mid \lan \mu, f\ran\leq r,\ \mu(\cX)\leq s\} \] Then $\cR^+_{f,r,s}$ is compact with the topology of weak convergence of measures. \end{lem} \begin{proof} Let $\{\mu_j\}_{j\in\N}$ be a sequence of measures in $\cR^+_{f,r,s}$. We need to show that there is a subsequence $\{\mu_{j_k}\}_{k\in\N}$ of $\{\mu_j\}$ that converges to some $\mu\in\cR^+_{f,r,s}$ weakly in the sense of measures. Since $f$ is bounded from below, the sequence $\{\lan\mu_j,f\ran\}_{j\in\N}$ is bounded from below, and, hence, it is bounded. We may thus assume, by passing to a subsequence if needed, that the sequence $\{\lan\mu_j,f\ran\}_{j\in\N}$ is convergent. Since $f=+\infty$ at infinity, the boundedness of the sequence $\{\lan\mu_j,f\ran\}_{j\in\N}$ and Chebyshev's inequality imply that the family $\{\mu_j\}_{j\in\N}$ is tight. Prokhorov's theorem guarantees that there is a subsequence $\{\mu_{j_k}\}_{k\in\N}$ of $\{\mu_j\}$ that converges to some $\mu\in\cR^+_\cX$ weakly in the sense of measures. The weak convergence of $\{\mu_{j_k}\}_{k\in\N}$ readily yields $\,\mu(\cX)\leq s$. Lemma \ref{basic-lsc} ensures that \[\lan\mu,f\ran\leq \lim_{k\to\infty}\lan\mu_{j_k},f\ran\leq r, \] which then shows that $\mu\in\cR^+_{f,r,s}$. The proof is complete. \end{proof} In the next lemma, we consider the case where $\cX=\bb$. \begin{lem}\label{mod} Let $\mu\in\cR^+_{\bb}$, $m\in\R$ and $L\in C(\bb)$. Assume that $\cA$ is not compact, $\mu(\bb)>0$, $m>\lan \mu,L\ran$, and that $L$ is coercive. Then, there exists $\tilde\mu\in\cR^+_{\bb}$ such that \begin{align} &\tilde\mu(\bb)=\mu(\bb),\tag{i}\\ &\lan\tilde\mu,L\ran=m,\tag{ii} \\ &\lan\tilde\mu,\psi\ran=\lan\mu,\psi\ran \ \ \ \text{ for all }\ \psi\in C(\tt).\tag{iii} \end{align} \end{lem} \begin{proof} Put $\, m_0:=\lan\mu,L\ran$, and pick $\ga_1\in\cA$ so that \[ \min_{x\in\tt}L(x,\ga_1)\tim \mu(\bb)>m. \] Define the Radon measure $\nu$ on $\bb$, through the Riesz representation theorem, by requiring \[ \lan\nu,\phi\ran=\int_{\bb}\phi(x,\ga_1)\mu(dxd\ga) \ \ \ \text{ for }\ \phi\in C_{\rm c}(\bb), \] and note that \[ \nu(\bb)=\int_{\bb}\mu(dx d\ga)=\mu(\bb). \] We set \[ m_1:=\lan\nu,L\ran, \] observe that \[ m_1\geq \int_{\bb}\min_{\tt}L(\cdot,\ga_1)\mu(dxd\ga) =\min_{\tt}L(\cdot,\ga_1)\mu(\bb)>m>m_0, \] and put \[ t=\fr{m-m_0}{m_1-m_0}. \] It is clear that $t\in (0,\,1)$ and, if we set \[ \tilde\mu=(1-t)\mu+t\nu, \] then $\tilde\mu$ is a nonnegative Radon measure on $\bb$. We observe \[ \tilde\mu(\bb)=(1-t)\mu(\bb)+t\nu(\bb)=\mu(\bb), \] that \[ \lan\tilde\mu,L\ran =(1-t)\lan\mu,L\ran+t\lan\nu,L\ran =(1-t)m_0+tm_1=m, \] and that for any $\psi\in C(\tt)$, \[\begin{aligned} \lan\tilde\mu,\psi\ran &\,=(1-t)\lan\mu,\psi\ran +t\int_{\bb}\psi(x)\nu(dxd\ga) \\&\,=(1-t)\lan\mu,\psi\ran +t\int_{\bb}\psi(x)\mu(dxd\ga) =\lan\mu,\psi\ran.\end{aligned} \] The proof is complete. \end{proof} \section{State constraint problem} \label{sec-s} We consider the state constraint boundary problem. To avoid confusion, we rename discount equation (DP$_\lambda$) as (S$_\gl$), and ergodic equation (E) as (ES) in this section. Here the letter S represents ``state constraint". The two problems of interest are \[\tag{S$_\gl$}\label{S} \begin{cases} \gl u+F[u]\leq 0 \ \ \text{ in }\ \gO, &\\[3pt] \gl u+F[u]\geq 0 \ \ \text{ on }\ \lbar\gO, \end{cases} \] for $\gl>0$, and \[\tag{ES}\label{ES} \begin{cases} F[u]\leq c \ \ \text{ in }\ \gO, &\\[3pt] F[u]\geq c \ \ \text{ on }\ \lbar\gO. \end{cases} \] Given a constant $c$, we refer as (ES$_c$) the state constraint problem (ES). We assume in addition the following assumptions, which concern with the comparison principle and solvability for \eqref{S}. \[\tag{CPS}\label{CPS} \left\{\text{ \begin{minipage}{0.85\textwidth} The comparison principle holds for \eqref{S} for every $\gl>0$. More precisely, if $v\in C(\ol \gO)$ is a subsolution of $\gl u + F[u] = 0$ in $\gO$, and $w\in C(\ol \gO)$ is a supersolution of $\gl u+F[u] = 0$ on $\tt$, then $v\leq w$ on $\tt$. \end{minipage} }\right. \] \[\tag{SLS}\label{SLS} \text{ \begin{minipage}{0.85\textwidth} For any $\gl>0$, \eqref{S} admits a solution $v^\gl \in C(\tt)$. \end{minipage} }. \] \begin{prop} Assume \eqref{F1}, \eqref{F2}, \eqref{CPS}, \eqref{SLS} and \eqref{EC}. Then there exists a solution $(u,c)\in C(\tt)\times\R$ of \eqref{ES}. Moreover, the constant $c$ is uniquely determined by \begin{equation}\label{critical-value-s} c=\inf\left\{d\in\R\mid \text{there exists} \ v\in C(\tt) \ \text{such that} \ F[v]\le d \ \text{in} \ \gO\right\}. \end{equation} \end{prop} We denote the constant given by \eqref{critical-value-s} by $c_{\rS}$ and call it the critical value of \eqref{ES}. The proof of the proposition above is somehow standard, and we give only its outline. \begin{proof} For each $\gl>0$ let $v^\gl\in C(\tt)$ be a (unique) solution of \eqref{S} and set $u^\gl:=v^\gl-m_\gl$, where $m_\gl:=\min_{\tt}v^\gl$. By \eqref{EC}, the family $\{u^\gl\}_{\gl>0}$ is relatively compact in $C(\tt)$. Set $M=\|v^1\|_{C(\tt)}$, and observe that for any $\gl>0$, the functions $v^1+(1+\gl^{-1})M$ and $v^{1}-(1+\gl^{-1})M$ are a supersolution and a subsolution of \eqref{S}, respectively. By the comparison principle \eqref{CPS}, we find that for any $\gl>0$, \[ v^{1}-(1+\gl^{-1})M\leq v^\gl\leq v^1+(1+\gl^{-1})M \ \ \text{ on }\tt, \] which implies that the collection $\{\gl m_\gl\}_{\gl>0}\subset\R$ is bounded. We may now choose a sequence $\{\gl_j\}_{j\in\N}$ converging to zero such that $\{u^{\gl_j}\}_{j\in\N}$ converges to a function $u\in C(\tt)$ and $\{\gl_j m_{\gl_j}\}_{j\in\N}$ converges to a number $-c^*$. Since $u^{\gl_j}$ is a solution of $\gl_j(u^{\gl_j}+m_{\gl_j})+F[u^{\gl_j}]\geq 0$ on $\tt$ and $\gl_j(u^{\gl_j}+m_{\gl_j})+F[u^{\gl_j}]\leq 0$ in $\gO$, we conclude in the limit $j\to\infty$ that $(u,c^*)$ is a solution of \eqref{ES}. Next we prove formula \eqref{critical-value-s}. We write $d^*$ the right side of \eqref{critical-value-s}. Since $u$ is a solution of (\ref{ES}$_{c^*}$), we get $\,d^*\leq c^*$. To show that $d^*\geq c^*$, we suppose $d^*<c^*$, and select $(v,d)\in C(\tt)\tim\R$ so that $v$ is a subsolution of $F[v]=d$ in $\gO$. By adding $v$ a constant if necessary, we may assume that $v>u$ on $\tt$. We observe that if $\ep>0$ is sufficiently small, then $u$ and $v$ are a supersolution and a subsolution of \eqref{S}, with $\gl$ replaced by $\ep$ and $0$ on its right side replaced by $(c+d)/2$, which means that the functions $u-(c+d)/(2\ep)$ and $v-(c+d)/(2\ep)$ are a supersolution and a subsolution of \eqref{S}, with $\gl$ replaced by $\ep$. By the comparison principle, we get \[ v-\fr{c+d}{2\ep}\leq u-\fr{c+d}{2\ep} \ \ \ \text{ on }\ \tt, \] but this is a contradiction, which shows that $d^*\geq c^*$. \end{proof} By normalization (replacing $F$ and $L$ by $F-c_{\rS}$ and $L+c_{\rS}$, respectively), we often assume \begin{equation}\tag{Z}\label{Z} \text{ \begin{minipage}{0.85\textwidth} the critical value $c_{\rS}$ of \eqref{ES} is zero. \end{minipage} }\end{equation} \subsection{Representation formulas} We assume \eqref{Z} in this subsection. For $z\in \ol \gO$ and $\gl \geq 0$, we define the sets $\cF^{\rS}(\gl)\subset C(\tt\tim\cA)\times C(\tt)$ and $\cG^{\rS}(z,\gl)\subset C(\tt)$, respectively, as \begin{align*} &\cF^{\rS}(\gl):=\left\{(\phi,u)\in \Phi^+\tim C(\tt)\mid u \ \text{ is a subsolution of } \ \gl u + F_\phi[u] =0 \ \text{ in } \ \gO \right\}, \\ &\cG^{\rS}(z,\gl):=\left\{\phi-\gl u(z)\mid (\phi,u)\in \cF^{\rS}(\gl) \right\}. \end{align*} As $\cG^{\rS}(z,0)$ is independent of $z$, we also write $\cG^{\rS}(0)$ for $\cG^{\rS}(z,0)$ for simplicity. \begin{lem}\label{thm3-sec3}Assume \eqref{F1}, \eqref{F2} and \eqref{CP}. For any $(z,\gl)\in\tt\tim[0,\infty)$, the set $\cG^{S}(z,\gl)$ is a convex cone with vertex at the origin. \end{lem} The proof of this lemma parallels that of \cite[Lemma 2.8]{IsMtTr1} and hence is skipped. For $(z,\gl)\in \tt \tim[0,\,\infty)$, we define the set $\cG^{\rS}(z,\gl)'\subset\cR_{L}$ (denoted also by $\cG^{\rS}(0)'$ if $\gl=0$) by \[ \cG^{\rS}(z,\gl)':=\{\mu\in \cR_{L}\mid \lan\mu,f\ran\geq 0 \ \ \text{ for all }f\in\cG^{\rS}(z,\gl)\}. \] The set $\cG^{\rS}(z,\gl)'$ is indeed the dual cone of $\cG^{\rS}(z,\gl)$ in $\cR_L$. Set $\cP^\rS:=\cP_{\bb}$, and, for any compact subset $K$ of $\cA$, let $\cP_K^\rS$ denote the subset of all probability measures $\mu\in\cP^\rS$ that have support in $\t\tim K$, that is, \[ \cP^\rS_K:=\{\mu\in\cP^\rS\mid \mu(\tt\tim K)=1\}. \] \begin{thm} \label{thm1-sc} Assume \eqref{F1}, \eqref{F2}, \eqref{CP}, \eqref{L}, \eqref{CPS}, \eqref{SLS}, \eqref{EC}, and, if $\gl = 0$, \eqref{Z} as well. Let $(z,\gl)\in\tt\tim[0,\,\infty)$ and let $v^\gl\in C(\tt)$ be a solution of \eqref{S}. Then \begin{equation} \label{sc-min} \gl v^\gl(z)=\inf_{\mu\in \cP^\rS\cap \cG^{\rS}(z,\gl)'}\lan \mu,L\ran. \end{equation} \end{thm} The proof of this theorem is similar to that of \cite[Theorem 3.3]{IsMtTr1}. However, we present it here, because it is a key component of the main result. \begin{proof} We note that, thanks to \eqref{L}, the term \[ \max_{x\in\tt}\min_{\ga\in\cA} L(x,\ga) \] defines a real number. We choose first a constant $L_\gl\in\R$ so that \begin{equation}\label{thm1-sc-m2} L_\gl\geq \max_{x\in\tt}\{\gl v^\gl(x),\min_{\ga\in\cA}L(x,\ga)\}, \end{equation} and then, in view of \eqref{L}, a compact subset $K_\gl$ of $\cA$ so that \begin{equation}\label{thm1-sc-m1} L(x,\ga)\geq L_\gl \ \ \ \text{ for all }\ (x,\ga)\in\tt \tim(\cA\setminus K_\gl). \end{equation} We fix a compact set $K\subset\cA$ so that $K=\cA$ if $\cA$ is compact, or, otherwise, $K\supsetneq K_\gl$. According to the definition of $\cG^{\rS}(z,\gl)'$, since $(L,v^\gl)\in\cF^{\rS}(\gl)$, we have \[ 0\leq \lan \mu,L-\gl v^\gl(z)\ran= \lan \mu,L\ran-\gl v^\gl(z)\ \ \text{ for all }\mu\in\cP^\rS\cap \cG^\rS(z,\gl)', \] and, therefore, \begin{equation}\label{thm1-sc-1} \gl v^\gl(z)\leq \inf_{\mu\in\cP^\rS\cap \cG^\rS(z,\gl)'}\lan\mu,\,L\ran. \end{equation} Next, we show \begin{equation}\label{thm1-sc-1+} \gl v^\gl(z)\geq \inf_{\mu\in\cP_K\cap \cG^{\rS}(z,\gl)'}\lan\mu,\,L\ran, \end{equation} which is enough to prove \eqref{sc-min}. We put $K_1:=\cA$ if $K_\gl=\cA$, or else, pick $\ga_1 \in K \setminus K_\gl$ and put $K_1 := K_\gl \cup \{\ga_1\}$. Clearly, $K_1$ is compact and $K_1 \subset K$. To prove \eqref{thm1-sc-1+}, we need only to show \begin{equation}\label{thm1-sc-1++} \gl v^\gl(z)\geq \inf_{\mu\in\cP_{K_1}\cap \cG^{\rS}(z,\gl)'}\lan\mu,\,L\ran. \end{equation} Suppose by contradiction that \eqref{thm1-sc-1++} is false, which means that \[ \gl v^\gl(z)<\inf_{\mu\in\cP_{K_1}\cap \cG^{\rS}(z,\gl)'}\lan \mu, L\ran. \] Pick $\ep>0$ sufficiently small such that \begin{equation} \label{thm1-sc-2} \gl v^\gl(z) + \ep<\inf_{\mu\in\cP_{K_1}\cap \cG^{\rS}(z,\gl)'}\lan \mu, L\ran. \end{equation} Since $\cG^{\rS}(z,\gl)$ is a convex cone with vertex at the origin, we infer that \[ \inf_{f\in\cG^{\rS}(z,\gl)} \lan \mu,f\ran= \begin{cases} 0 \ \ &\text{ if }\ \mu\in\cP_{K_1} \cap\cG^{\rS}(z,\gl)', \\ -\infty &\text{ if }\ \mu\in \cP_{K_1} \setminus \cG^{\rS}(z,\gl)'. \end{cases} \] and, hence, the right hand side of \eqref{thm1-sc-2} can be rewritten as \begin{equation}\begin{aligned} \label{thm1-sc-2+} \inf_{\mu\in\cP_{K_1} \cap \cG^{\rS}(z,\gl)'}\lan \mu, L\ran &\,= \inf_{\mu\in\cP_{K_1}}\ \Big(\lan \mu, L\ran - \inf_{f\in\cG^{\rS}(z,\gl)}\lan\mu,f\ran \Big) =\inf_{\mu\in\cP_{K_1}}\ \sup_{f\in\cG^{\rS}(z,\gl)}\,\lan \mu, L-f\ran. \end{aligned} \end{equation} Since $\cP_{K_1}$ is a convex compact space, with the topology of weak convergence of measures, we apply Sion's minimax theorem, to get \[ \min_{\mu\in\cP_{K_1}} \ \sup_{f\in\cG^{\rS}(z,\gl)}\,\lan \mu, L-f\ran = \sup_{f\in\cG^{\rS}(z,\gl)}\ \min_{\mu\in\cP_{K_1}} \,\lan \mu, L-f\ran. \] In view of this and \eqref{thm1-sc-2}, we can pick $(\phi,u) \in \cF^{\rS}(\gl)$ such that \begin{equation}\label{thm1-sc-3} \gl v^\gl(z)+\ep < \lan \mu, L-\phi + \gl u(z) \ran \ \ \text{ for all } \mu \in \cP_{K_1}, \end{equation} and $\phi=tL+\chi$ for some $t>0$ and $\chi \in C(\ol \gO)$. We now prove that there exists $\gth>0$ such that $w:=\gth u$ is a subsolution of \begin{equation}\label{thm1-sc-4} \gl w+F[w]=-\gl (v^\gl-w)(z)-\gth\ep\ \ \text{ in }\gO. \end{equation} Once this is done, we immediately arrive at a contradiction. Indeed, if $\gl>0$, then $\gz:=w+{(v^\gl-w)(z)}+\gl^{-1}\gth\ep$ is a subsolution of $\gl \gz+F[\gz]=0$ in $\gO$, and comparison principle \eqref{CPS} yields $\gz\leq v^\gl$, which, after evaluation at $z$, gives $\,\gl^{-1}\gth\ep\leq 0$, a contradiction. On the other hand, if $\gl=0$, then we set $\gz:=w+C$, with a constant $C>\min_{\tt}(v^\gl-w)$, choose a constant $\gd>0$ sufficiently small so that \[ \gd\min_{\tt}v^\gl\geq -\gth\ep/2 \ \ \text{ and } \ \ \gd\max_{\tt} \gz\leq \gth\ep/2, \] observe that the functions $\gz$ and $v^\gl$ are a subsolution and a supersolution of $\gd u+F[u]=-\gth\ep/2$, respectively, in $\gO$ and on $\tt$, and, by \eqref{CPS}, get $\gz\le v^\gl$ on $\tt$, which is a contradiction. To show \eqref{thm1-sc-4}, we consider the two cases separately. The first case is when $K_1=\cA$, and then, the Dirac measure $\gd_{(x,\ga)}$ belongs to $\cP_{K_1}$ for any $(x,\ga) \in \ol \gO \times \cA$, which together with \eqref{thm1-sc-3} yields \[ \gl v^\gl(z) + \ep < L - \phi + \gl u(z) \quad \text{on } \ol \gO \times \cA. \] Now, we have \[ F(x,p,X) \leq F_\phi(x,p,X) + \gl(u-v^\gl)(z) - \ep \quad \text{for} \ (x,p,X) \in \tt \times \R^n \times \bS^n. \] We choose $\theta=1$ and observe that $w=u$ is a subsolution of \eqref{thm1-sc-4}. The other case is that when $K_1=K_\gl \cup\{\ga_1\}$, with $\ga_1 \in K \setminus K_\gl$. As $\gd_{(x,\ga)}\in\cP_{K_1}$ for all $(x,\ga)\in\tt\tim K_1$, we observe, in light of \eqref{thm1-sc-3}, that \begin{equation}\label{thm1-sc-6} \gl v^\gl(z)+\ep<(1-t)L(x,\ga)-\chi(x)+\gl u(z) \ \ \text{ for all }(x,\ga)\in\tt\tim K_1. \end{equation} We subdivide the argument into two cases. Consider first the case when $t \leq 1$. Minimize both sides of \eqref{thm1-sc-6} in $\ga \in K_\gl$ and note by \eqref{thm1-sc-m2} and \eqref{thm1-sc-m1} that $\min_{\ga\in K_\gl}L(x,\ga)\leq L_\gl$, to get \[\begin{aligned} \gl v^\gl(z)+\ep&\,<(1-t)\min_{\ga \in K_\gl} L(x,\ga)-\chi(x)+\gl u(z) \\&\,\leq (1-t)L_\gl-\chi(x)+\gl u(z) \ \ \text{ for all } x \in\tt. \end{aligned}\] We use this, \eqref{thm1-sc-m1}, and \eqref{thm1-sc-6}, to deduce that \[ \gl v^\gl(z)+\ep<(1-t)L(x,\ga)-\chi(x)+\gl u(z) \ \ \text{ for all }(x,\ga)\in\tt\tim \cA. \] From this, we observe \[ \phi =L+(t-1)L+\chi<L-\gl (v^\gl-u)(z)-\ep \ \text{ on }\bb, \] and, hence, that $u$ is a subsolution of \eqref{thm1-sc-4}, with $\gth=1$. Next, we consider the case when $t \geq 1$. By \eqref{thm1-sc-m1} and \eqref{thm1-sc-6}, we get \[L(x,\ga_1) \geq L_\gl \ \ \text{ and } \ \ (t-1)L(x,\ga_1)+\chi(x)<-\gl (v^\gl-u)(z)-\ep \ \ \text{ for all }x\in\tt, \] which together yield \[ (t-1) L_\gl + \chi(x) \leq -\gl(v^\gl-u)(z) - \ep \ \ \text{ for all } x \in \tt. \] We take advantage of \eqref{thm1-sc-m2} to get further that \[ (t-1) \gl v^\gl(z) + \chi(x) \leq -\gl(v^\gl-u)(z) - \ep \ \ \text{ for all } x \in\tt. \] Therefore, \[ \chi < - t \gl v^\gl(z) + \gl u(z) - \ep, \] and \[ \phi=tL+\chi < tL + t\gl (u/t-v^\gl)(z) - \ep \ \ \text{ in } \tt \times \cA. \] From this we deduce that $w:=u/t$ is a subsolution of \[ \gl w +F[w]=-\gl(v^\gl-w)(z)-\ep/t \ \ \ \text{ in }\gO. \] This completes the proof. \end{proof} We remark that, in the proof above, we have proven the identity \[ \gl v^\gl(z)=\inf_{\mu\in\cP_K\cap\cG^\rS(z,\gl)'}\lan\mu,L\ran \] for a compact set $K\subset\cA$, which is a stronger claim than \eqref{sc-min}. The minimization problem \eqref{sc-min}, in Theorem \ref{thm1-sc}, has minimizers as stated in the upcoming corollary, Corollary \ref{cor1-sc}. \begin{lem} \label{sc-cpt} Assume \eqref{F1}, \eqref{F2} and \eqref{L}. Fix a point $z\in\tt$ and a sequence $\{\gl_j\}_{j\in\N}\subset[0,\,\infty)$. Let $\{\mu_j\}_{j\in\N}$ be a sequence of measures such that $\mu_j\in\cP^\rS\cap \cG^\rS(z,\gl_j)'$ for all $j\in\N$. Assume that for some $\rho\in\R$ and $\gl\in[0,\,\infty)$, \[ \lim_{j\to\infty}\lan\mu_j,L\ran=\rho \ \ \text{ and } \ \ \lim_{j\to\infty}\gl_j=\gl. \] Then there exist a measure $\nu\in\cP^\rS\cap \cG^\rS(z,\gl)'$ and a subsequence $\{\mu_{j_k}\}_{k\in\N}$ of $\{\mu_j\}$ such that $\,\lan\nu,L\ran=\rho\,$ and \[ \lim_{k\to\infty}\lan\mu_{j_k},\psi\ran=\lan\nu,\psi\ran \ \ \text{ for all }\ \psi\in C(\tt). \] \end{lem} \begin{proof} Lemma \ref{basic-cpt} implies that $\{\mu_j\}_{j\in\N}$ has a subsequence $\{\mu_{j_k}\}_{k\in\N}$ convergent in the topology of weak convergence of measures. Let $\mu_0\in\cR^+_{\bb}$ denote the limit of $\{\mu_{j_k}\}_{k\in\N}$. It follows from the weak convergence of $\{\mu_{j_k}\}$ that $\mu_0$ is a probability measure on $\bb$. Hence, $\mu_0\in\cP^\rS$. By the lower semicontinuity of the functional $\mu\mapsto\lan\mu,L\ran$ on $\cR^+_{\bb}$, as claimed by Lemma \ref{basic-lsc}, we find that $\, \rho\geq \lan\mu_0,L\ran$. If $\rho>\lan\mu_0,L\ran$, then $\cA$ is not compact and Lemma \ref{mod} ensures that there is $\tilde\mu_0\in\cP^\rS$ such that $\lan\tilde\mu_0,L\ran=\rho$ and $\lan\tilde\mu_0,\psi\ran=\lan\mu_0,\psi\ran$ for all $\psi\in C(\tt)$. We define $\nu\in\cP^\rS$ by setting $\nu=\tilde\mu_0$ if $\rho>\lan\mu_0,L\ran$ and $\nu=\mu_0$ otherwise, that is, if $\rho=\lan\mu_0,L\ran$. The measure $\nu\in\cP^\rS$ verifies that $\lan\nu,L\ran=\rho$ and $\lan\nu,\psi\ran=\lan\mu_0,\psi\ran$ for all $\psi\in C(\tt)$. It follows from the last identity that \[ \lim_{k\to\infty}\lan\mu_{j_k},\psi\ran=\lan\nu,\psi\ran \ \ \ \text{ for all }\ \psi\in C(\tt). \] It remains to check that $\nu\in\cG^\rS(z,\gl)'$. Let $f\in\cG^\rS(z,\gl)$, and select $(\phi,u)\in\cF^\rS(\gl)$, $t>0$ and $\chi\in C(\tt)$ so that $f=\phi-\gl u(z)$ and $\phi=tL+\chi$. Fix $j\in\N$, note that $(\phi+(\gl_j-\gl)u,u) \in\cF^\rS(\gl_j)$ and get \[ 0\leq \lan\mu_j,\phi+(\gl_j-\gl)u-\gl_j u(z)\ran =t\lan\mu_j,L\ran+ \lan\mu_j,\chi+(\gl_j-\gl)u-\gl_j u(z)\ran, \] because $\mu_j\in\cG^\rS(z,\gl_j)'$. Sending $j\to\infty$ yields \[ 0\leq t\rho +\lan\mu_0,\chi-\gl u(z)\ran. \] Since $\rho=\lan\nu,L\ran$ and $\lan\mu_0,\chi\ran =\lan\nu,\chi\ran$, we find that \[ 0\leq t\lan\nu,L\ran+\lan\nu,\chi-\gl u(z)\ran =\lan\nu,\phi-\gl u(z)\ran, \] which shows that $\nu\in\cG^\rS(z,\gl)'$, completing the proof. \end{proof} \begin{cor}\label{cor1-sc} Under the hypotheses of Theorem \ref{thm1-sc}, we have \begin{equation}\label{cor1-sc-1} \gl v^\gl(z)=\min_{\mu\in\cP^\rS\cap\cG^\rS(z,\gl)'} \lan\mu,L\ran. \end{equation} \end{cor} \begin{proof} In view of Lemma \ref{sc-cpt}, we need only to show that there exists a sequence $\{\mu_j\}_{j\in\N}\subset \cP^\rS\cap\cG^\rS(z,\gl)'$ such that \[ \lim_{j\to\infty} \lan\mu_j,L\ran =\inf_{\mu\in\cP^\rS\cap\cG^\rS(z,\gl)'}\lan\mu,L\ran, \] but this is obviously true. \end{proof} \begin{remark} \label{rem1-s} In the generality that \eqref{Z} is not assumed in Corollary \ref{cor1-sc}, we have \[ -c_{\rS}=\inf_{\mu\in \cP^\rS\cap \cG^{\rS}(0)'}\lan \mu,L\ran. \] \end{remark} \begin{definition} We denote the set of minimizers of \eqref{cor1-sc-1} by $\cM^{\rS}(z,\gl)$ for $\gl>0$, and by $\cM^{\rS}(0)$ for $\gl=0$. We call any $\mu \in \cM^{\rS}(0)$ a viscosity Mather measure, and, if $\gl>0$, any $\gl^{-1}\mu$, with $\mu \in \cM^{\rS}(z,\gl)$ and $\gl>0$, a viscosity Green measure. \end{definition} \subsection{Convergence with vanishing discount} The following theorem is our main result on the vanishing discount problem for the state constraint problem. \begin{thm}\label{thm2-sc} Assume \eqref{F1}, \eqref{F2}, \eqref{CP}, \eqref{L}, \eqref{CPS}, \eqref{SLS}, and \eqref{EC}. For each $\gl>0$, let $v^\gl\in C(\tt)$ be the unique solution of \eqref{S}. Then, the family $\{v^\gl+\gl^{-1}c_{\rS}\}_{\gl>0}$ converges to a function $u$ in $ C(\tt)$ as $\gl\to 0$. Furthermore, $(u,c_{\rS})$ is a solution of \eqref{ES}. \end{thm} In order to prove this theorem, we need the next lemma. \begin{lem}\label{lem2-sc} Assume \eqref{F1}, \eqref{F2}, \eqref{CPS}, \eqref{SLS}, \eqref{EC} and \eqref{Z}. For each $\gl>0$, let $v^\gl \in C(\tt)$ be the unique solution of \eqref{S}. Then $\{v^\gl\}_{\gl>0}$ is uniformly bounded on $\tt$. \end{lem} \begin{proof} Let $u \in C(\tt)$ be a solution of (\ref{ES}$_0$). It is clear that $u+\|u\|_{C(\tt)}$ and $u-\|u\|_{C(\tt)}$ are a supersolution and a subsolution of \eqref{S} respectively for any $\gl>0$. By the comparison principle, we get \[ u-\|u\|_{C(\tt)} \leq v^\gl \leq u+\|u\|_{C(\tt)} \ \ \text{on } \tt, \] which yields $\|v^\gl\|_{C(\tt)} \leq 2 \|u\|_{C(\tt)}$. \end{proof} We are now ready to prove the main convergence result in this section. \begin{proof}[Proof of Theorem \ref{thm2-sc}] As always, we assume $c_{\rS}=0$. Let $\cU$ be the set of accumulation points in $C(\tt)$ of $\{v^\gl\}_{\gl>0}$ as $\gl \to 0$. By Lemma \ref{lem2-sc} and \eqref{EC}, $\{v^\gl\}_{\gl>0}$ is relatively compact in $C(\tt)$. Clearly, $\cU \neq \emptyset$ and any $u\in\cU$ is a solution of (\ref{ES}$_0$). Our goal is achieved when we prove that $\cU$ has a unique element, or equivalently, for any $v, w \in \cU$, \begin{equation} \label{thm2-sc-1} v \geq w \ \ \text{ on } \tt. \end{equation} Fix any $v,w \in \cU$. There exist two sequences $\{\gl_j\}$ and $\{\gd_j\}$ of positive numbers converging to $0$ such that $v^{\gl_j} \to v$, $v^{\gd_j} \to w$ in $C(\tt)$ as $j \to \infty$. Fix $z \in \tt$. By Corollary \ref{cor1-sc}, there exists a sequence $\{\mu_j\}_{j\in\N}$ of measures such that $\mu_j \in \cM^{\rS}(L,z,\gl_j)$ for every $j \in \N$. Since \[ \lan\mu_j,L\ran=\gl_j v^{\gl_j}(z) \to 0=c_\rS\ \ \text{ as }\ j\to\infty, \] Lemma \ref{sc-cpt} guarantees that there is $\mu\in\cP^\rS\cap\cG^\rS(0)'$ such that \[ \lan\mu,L\ran=0 \ \ \ \text{ and } \ \ \ \lim_{j\to\infty}\lan\mu_j,\psi\ran=\lan\mu,\psi\ran \ \ \ \text{ for all }\ \psi\in C(\tt). \] We note that $(L-\gd_j v^{\gd_j}, v^{\gd_j}) \in \cF^{\rS}(0)$ and $(L+\gl_j w, w) \in \cF^{\rS}(\gl_j)$, which implies that \[ 0 \leq \lan \mu, L-\gd_j v^{\gd_j} \ran = -\gd_j \lan \mu, v^{\gd_j} \ran, \] and \[ 0 \leq \lan \mu_j, L + \gl_j w - \gl_j w(z) \ran = \gl_j (v^{\gl_j} - w)(z) + \gl_j \lan \mu_j, w \ran. \] Dividing the above inequalities by $\gd_j$ and $\gl_j$, respectively, and letting $j \to \infty$ yield \[ \lan \mu, w \ran \leq 0 \quad \text{and} \quad 0 \leq (v-w)(z) + \lan \mu, w \ran, \] and thus, $v(z) \geq w(z)$. This completes the proof. \end{proof} \section{Dirichlet problem} \label{sec-d} We consider the Dirichlet problem in this section. We rename \eqref{DP} and \eqref{E} as \eqref{D} and \eqref{ED}, respectively, in which the letter D refers to ``Dirichlet". For a given $g\in C(\bry)$, the two problems of interest are \begin{equation}\tag{D$_\gl$}\label{D} \begin{cases} \gl u+F[u]= 0 \ \ \text{ in }\ \gO, &\\[3pt] u=g \ \ \text{ on }\ \bry, \end{cases} \end{equation} for $\gl>0$, and \[\tag{ED}\label{ED} \begin{cases} F[u]= c \ \ \text{ in }\ \gO, &\\[3pt] u=g \ \ \text{ on }\ \pl\gO. \end{cases} \] As usual, (ED$_c$) refers the Dirichlet problem (ED), with a given constant $c$. The function $g\in C(\tt)$ is fixed in the following argument, while we also consider the Dirichlet problem \[\tag{D$_{\gl,\phi,\psi}$}\label{D'} \begin{cases} \gl u+F_\phi[u]=0 \ \ \text{ in }\ \gO,\\[3pt] u=\psi \ \ \ \text{ on }\ \bry, \end{cases} \] where $\gl\geq 0$, $\phi\in C(\tt\tim\cA)$ and $\psi\in C(\tt)$ are all given. Throughout this paper, we understand that $u\in C(\tt)$ is a subsolution (resp., a supersolution) of \eqref{D'}, with $\gl\geq 0$, if it is a subsolution of $\gl u+F_\phi[u]=0$ in $\gO$ in the viscosity sense and verifies $u\leq \psi$ pointwise on $\pl\gO$ (resp., a supersolution of \eqref{D'} in the viscosity sense). As always, we call $u\in C(\tt)$ a solution of \eqref{D'} if it is a subsolution and supersolution of \eqref{D'}. We assume in addition the following conditions. \[\tag{CPD}\label{CPD} \left\{\text{ \begin{minipage}{0.85\textwidth} The comparison principle holds for \eqref{D'}, with $\phi=L+\chi$, for any $\gl>0$, $\chi\in C(\tt)$ and $\psi\in C(\bry)$. That is, for any subsolution $u\in C(\tt)$ and supersolution $v\in C(\tt)$ of \eqref{D'}, with $\phi=L+\chi$, the inequality $\,u\leq v\,$ holds on $\tt$. \end{minipage} }\right. \] We remark here that, if $\phi=L+\chi$, then the equation $\,\gl u+F_\phi[u]=0\,$ can be written as $\gl u+F[u]=\chi$. \[\tag{SLD}\label{SLD} \left. \text{ \begin{minipage}{0.85\textwidth} For every $\gl>0$, \eqref{D} admits a solution $v^\gl \in C(\tt)$. \end{minipage} }\right. \] In the vanishing discount problem for the Dirichlet problem, the state constrain problem comes into play as the following results indicate. \begin{prop}\label{prop1-d0} Assume \eqref{F1}, \eqref{F2}, \eqref{CPD} and \eqref{SLD}. For $\gl>0$, let $v^\gl\in C(\tt)$ be the solution of \eqref{D}. Then, \emph{(i)}\ if \emph{(ED$_0$)} has a solution in $C(\tt)$, then \[ \lim_{\gl\to 0+ }\gl v^\gl(x)=0 \ \ \ \text{ uniformly on }\ \tt, \] and \emph{(ii)}\ if \emph{(ES$_c$)}, with $c>0$, has a solution in $C(\tt)$, then \[ \lim_{\gl\to 0+}\gl v^\gl(x)=-c \ \ \ \text{ uniformly on }\ \tt. \] \end{prop} Notice that, in the proposition above, the assumption for the claim (i) and that for (ii) are mutually exclusive, since the conclusions are exclusive of one another. \begin{proof} Assume first that (ED$_0$) has a solution in $C(\tt)$, which denote by $u\in C(\tt)$. We select $M>0$ so that $\|u\|_{C(\tt)}\leq M$ and note that the function $u+M$ is nonnegative on $\tt$ and hence a supersolution of \eqref{D} for any $\gl>0$ and, similarly, that $u-M$ is a subsolution of \eqref{D} for any $\gl>0$. Thus, by \eqref{CPD}, we have $u-M\leq v^\gl\leq u+M$ on $\tt$, which readily yields \[ \lim_{\gl\to 0}\gl v^\gl(x)=0 \ \ \ \text{ uniformly on }\ \tt. \] Next, assume that (ES$_c$), with $c>0$, has a solution in $C(\tt)$, and let $u\in C(\tt)$ be such a solution. Choose $M>0$ large enough so that $\|u\|_{C(\tt)}+\|g\|_{C(\bry)}\leq M$. Observe that, for any $\gl>0$, $u+M-\gl^{-1}c$ is a supersolution of $\gl w+F[w]=0$ on $\tt$ and, therefore, is a supersolution of \eqref{D} and that $u-M-\gl^{-1}c$ is a subsolution of \eqref{D} for any $\gl>0$. Hence, we get $u-M-\gl^{-1}c\leq v^\gl\leq u+M-\gl^{-1}c$ on $\tt$ for all $\gl>0$. This shows that \[ \lim_{\gl\to 0}\gl v^\gl(x)=-c. \] The proof is now complete. \end{proof} \begin{prop}\label{prop1-d} Assume \eqref{F1}, \eqref{F2}, \eqref{CPD}, \eqref{SLD} and \eqref{EC}. Then, there is a dichotomy: either problem \emph{(ED$_0$)}, or \emph{(ES$_c$)}, with some $c>0$, has a solution in $C(\tt)$. \end{prop} Here is an illustrative, simple example regarding the solvability of \eqref{ED} or \eqref{ES}. Let $n=1$ and $m\in\R$, and consider the case $\gO=(-1,\,1)$, $F(x,p)=|p|+m$, and $g=0$. It is easily checked that conditions \eqref{F1}, \eqref{F2}, \eqref{CPD}, \eqref{SLD}, and \eqref{EC} are satisfied. The Dirichlet problem \begin{equation}\label{ex-d-1} |u'|+m=c \ \ \text{ in }(-1,\,1)\quad \text{ and }\quad u(-1)=u(1)=0, \end{equation} with $c\in\R$, has a solution $u(x):=(c-m)(1-|x|)$ if and only if $c\geq m$. Moreover, if $c=m$, then any function $u(x):=C$, with $C\leq 0$, is a solution of \eqref{ex-d-1} and any constant function $u$ is a solution of the state constraint problem \begin{equation}\label{ex-d-2} |u'|+m\geq c \ \ \text{ in }[-1,\,1] \quad\text{ and }\quad |u'|+m\leq c \ \ \text{ in }(-1,\,1). \end{equation} Thus, if $m\leq 0$, then problem \eqref{ex-d-1}, with $c=0$, has a solution in $C([-1,\,1])$, and if $m > 0$, then problem \eqref{ex-d-2}, with $c=m$, has a solution in $C([-1,\,1])$. \begin{proof} For $\gl>0$, let $v^\gl\in C(\tt)$ be the solution of \eqref{D}. Choose a constant $M>0$ so that $\|g\|_{C(\bry)}+\|F(\cdot,0,0)\|_{C(\tt)}\leq M$ and observe that, if $\gl>0$, then the constant functions $\gl^{-1}M$ and $-\gl^{-1}M$ are, respectively, a supersolution and a subsolution of \eqref{D}. By \eqref{CPD}, we have $-\gl^{-1}M\leq v^\gl\leq \gl^{-1}M$ on $\tt$, and consequently, the family $\{\gl v^\gl\}_{\gl>0}$ is uniformly bounded on $\tt$. By \eqref{EC}, the family $\{\gl v^\gl\}_{\gl>0}$ is equi-continuous on $\tt$. Setting \[ w^\gl:=v^\gl-\min_{\tt} v^\gl \ \ \ \text{ on }\ \tt \ \text{ for }\ \gl>0, \] we observe that $\{w^\gl\}_{\gl>0}$ is relatively compact in $C(\tt)$, and choose a sequence $\{\gl_j\}_{j\in\N}\subset (0,\,\infty)$, converging to zero, so that $\{w^{\gl_j}\}_{j\in\N}$ converges in $C(\tt)$ to some function $u\in C(\tt)$ as $j\to\infty$. The uniform boundedness of $\{\gl v^\gl\}_{\gl>0}$ allows us to assume, after taking a subsequence if necessary, that the limit \[ d:=\lim_{j\to\infty}\gl_j\min_{\tt}v^{\gl_j}\, \in\,\R \] exists. Then the equi-continuity of $\{v^\gl\}_{\gl>0}$ ensures that as $j\to\infty$, \[ \gl_j v^{\gl_j}(x) \to d \ \ \ \text{ uniformly on }\tt. \] Now, according to the boundary condition $v^\gl\leq g$, pointwise on $\bry$, for any $\gl>0$, we have \begin{equation}\label{prop-d1-1} \min_{\tt}v^\gl\leq \min_{\bry}g \ \ \ \text{ for all }\ \gl>0, \end{equation} and hence, we find that $d\leq 0$. Set \[ m_j:=\min_{\tt}v^{\gl_j} \ \ \ \text{ for }j\in\N, \] and consider the sequence $\{m_j\}_{j\in\N}$, which is bounded from above due to \eqref{prop-d1-1}. By passing to a subsequence if needed, we may assume that \[ m:=\lim_{j\to\infty}m_j\in[-\infty,\,\min_{\bry}g]. \] Observe that, for any $j\in\N$, $w^{\gl_j}$ is a solution of \[ \begin{cases} \gl_j(w^{\gl_j}+m_j)+F[w_j]=0 \ \ \text{ in }\gO,\\[3pt] w^{\gl_j}+m_j=g \ \ \ \text{ on }\ \pl\gO. \end{cases} \] Thus, in the limit $j\to\infty$, we find that if $m>-\infty$, then $w$ is a solution of (\ref{ED}$_0$), with $g$ replaced by $g-m$, which says that the function $u:=w+m$ is a solution of (ED$_0$). Notice that if $m>-\infty$, then $d=0$. On the other hand, if $m=-\infty$, then, for $j$ sufficiently large, we have $v^{\gl_j}<g$ on $\bry$, which implies that $v^{\gl_j}$ is a supersolution of $\gl_j v^{\gl_j}+F[v^{\gl_j}]=0$ on $\tt$. This can be stated that $w^{\gl_j}$ is a supersolution of $\gl_j(w^{\gl_j}+m_j)+F[w^{\gl_j}]=0$ on $\tt$. Sending $j\to\infty$, we deduce that $w$ is a solution of (ES$_{-d}$). Note that if $w\in C(\tt)$ is a solution of (ES$_0$) and if $C\in \R$ is large enough so that $w-C\leq g$ on $\bry$, then $u:=w-C$ is a solution of (ED$_0$). Thus, we conclude that either problem (ED$_0$) has a solution in $C(\tt)$, or else problem (ES$_c$), with some $c>0$, has a solution in $C(\tt)$. \end{proof} \subsection{Representation formulas in the case $\gl>0$} We first need to do some setup to take the Dirichlet boundary condition into account. For $M>0$ and $\gl\ge 0$, we define $\cF^{\rD}(\gl)$ (resp., $\cF^{\rD}(M,\gl)$) as the set of all $(\phi,\psi,u)\in \Psi^+\tim C(\tt)$ (resp., $(\phi,\psi,u)\in \Psi^+(M)\tim C(\tt)$) such that $u$ is a subsolution of \eqref{D'}. Fix $z \in \tt$ and define the sets $\cG^{\rD}(z,\gl),\,\cG^{\rD}(M,z,\gl)\subset C(\tt)$ by \[ \begin{aligned} \cG^{\rD}(z,\gl):&\,=\left\{\big(\phi-\gl u(z), \gl(\psi-u(z))\big)\mid (\phi,\psi,u)\in\cF^{\rD}(\gl)\right\},\\ \cG^{\rD}(M,z,\gl):&\,=\left\{\big(\phi-\gl u(z), \gl(\psi-u(z))\big)\mid (\phi,\psi,u)\in\cF^{\rD}(M,\gl)\right\}. \end{aligned}\] We also write $\cG^\rD(0)$ and $\cG^{\rD}(M,0)$, respectively, for $\cG^{\rD}(z,0)$ and $\cG^{\rD}(M,z,0)$, which are independent of $z$. Notice that \[\cG^{\rD}(0)=\left\{(\phi, 0)\mid (\phi,\psi,u)\in\cF^{\rD}(0)\right\}\ \text{ and }\ \cG^{\rD}(M,0)=\left\{(\phi, 0)\mid (\phi,\psi,u)\in\cF^{\rD}(M,0)\right\}, \] and that \[ \cF^\rD(z,\gl)=\bigcup_{M>0} \cF^{\rD}(M,z,\gl) \ \ \ \text{ and } \ \ \ \cG^{\rD}(z,\gl)= \bigcup_{M>0} \cG^{\rD}(M,z,\gl). \] \begin{lem}\label{lem:dirichlet-convex} Assume \eqref{F1}, \eqref{F2}, and \eqref{CP}. For any $(z,\gl)\in\tt\tim[0,\infty)$ and $M>0$, the sets $\cF^\rD_{z,\gl}$, $\cF^\rD(M,z,\gl)$, $\cG^\rD(z,\gl)$ and $\cG^{\rD}(M, z,\gl)$ are convex cones with vertex at the origin. \end{lem} \begin{proof} It is easily seen that $\Psi^+$ and $\Psi^+(M)$ are convex cones with vertex at the origin. For $i=1,2$, let $(\phi_i,\psi_i,u_i)\in\cF^{\rD}(\gl)$, fix $t,s\in(0,\,\infty)$ and set \[ u=tu_1+su_2, \quad \phi=t\phi_1+s\phi_2 \ \ \text{ and } \ \ \psi=t\psi_1+ s\psi_2. \] As in the proof of \cite[Lemma 2.8]{IsMtTr1}, we find that $u$ is a subsolution of $\gl u+F_\phi[u]=0$ in $\gO$. Since $u_i\leq \psi_i$ pointwise in $\bry$ for $i=1,2$, we get immediately $u\leq \psi$ pointwise in $\bry$. Hence, we see that $(\phi,\psi,u)\in\cF^\rD(\gl)$. Assume, in addition, that $(\phi_i,\psi_i,u_i)\in\cF^\rD(M,\gl)$ for $i=1,2$. Then $(\phi_i,\psi_i)\in\Psi^+(M)$ for $i=1,2$, and, hence, the cone property of $\Psi^+(M)$ implies that $(\phi,\psi)\in\Psi^+(M)$, which proves, together with the property of $(\phi,\psi,u)$ being in $\cF^\rD(\gl)$, that $(\phi,\psi,u)\in\cF^\rD(M,\gl)$. Thus, we see that $\cF^{\rD}(\gl)$ and $\cF^{\rD}(M,\gl)$ (and also $\cG^{\rD}(z,\gl)$ and $\cG^{\rD}(M,z,\gl)$) are convex cones with vertex at the origin. \end{proof} Henceforth, we write \[ \cR_{1}:=\cR_{\bb},\quad \cR_2:=\cR_{\bry},\quad \cR_L^+:=\cR_{\bb}^+\cap\cR_L \ \ \text{ and } \ \ \quad\cR_2^+:=\cR_{\bry}^+. \] Define \[ \cP^{\rD}:=\left\{(\mu_1,\mu_2)\in\cR_{L}^+\times\cR_{2}^{+}\mid \mu_1(\tt\times \cA)+\mu_2(\bry)=1\right\}, \] and, for any compact subset $K$ of $\cA$, \[ \cP^{\rD}_K:=\left\{(\mu_1,\mu_2)\in\cP^{\rD}\mid \mu_1(\tt\times K)+\mu_2(\bry)=1\right\}. \] We define the dual cones $\cG^{\rD}(z,\gl)',\,\cG^\rD(M,z,\gl)'$ of $\cG^{\rD}(z,\gl),\,\cG^{\rD}(M,z,\gl)$ in $\cR_{L}\times\cR_{2}$, respectively, by \[\begin{aligned} \cG^\rD(z,\gl)':&\,=\left\{(\mu_1,\mu_2)\in \cR_{L}\times\cR_{2} \mid \lan\mu_1,f_1\ran+\lan\mu_2,f_2\ran\ge0 \ \ \text{ for all } (f_1,f_2)\in\cG^{\rD}(z,\gl)\right\}, \\ \cG^\rD(M,z,\gl)':&\,=\left\{(\mu_1,\mu_2)\in \cR_{L}\times\cR_{2} \mid \lan\mu_1,f_1\ran+\lan\mu_2,f_2\ran\ge0 \ \ \text{ for all } (f_1,f_2)\in\cG^{\rD}(M,z,\gl)\right\}. \end{aligned} \] It is obvious that \[ \cG^\rD(z,\gl)'=\bigcap_{M>0}\cG^\rD(M,z,\gl)'. \] As usual, we write $\cG^\rD(0)'$ and $\cG^\rD(M)'$ for $\cG^\rD(z,0)'$ and $\cG^\rD(M,z,0)'$, respectively. The following proposition is a key step toward Theorems \ref{thm1-d0} and \ref{thm3-d}, two of the main results in this section. \begin{thm} \label{thm1-d} Assume \eqref{F1}, \eqref{F2}, \eqref{CP}, \eqref{L} and \eqref{CPD}. Let $(z,\gl)\in\tt\tim (0,\,\infty)$ and $M> \|g\|_{C(\bry)}$. If $v^\gl\in C(\tt)$ is a solution of \eqref{D}, then \begin{equation}\label{thm1-d-0} \gl v^\gl(z)=\inf_{(\mu_1,\mu_2)\in \cP^{\rD} \cap \cG^\rD(M,z,\gl)'}\,\left( \lan \mu_1,\,L\ran+\gl\lan\mu_2,\,g\ran\right). \end{equation} \end{thm} We need a lemma for the proof of the theorem above. \begin{lem}\label{lem1-d}Assume \eqref{F1}, \eqref{F2}, and \eqref{CPD}. Let $\gl>0$. If $(\phi,\psi,u)\in\cF^{\rD}(\gl)$ and $\phi=t(L+\chi)$, with $t>0$ and $\chi\in C(\tt)$, then \[ \gl u\leq \max\left\{t\max_{\tt}(\chi-F(\cdot,0,0)),\,\gl\max_{\bry}\psi \right\} \ \ \text{ on }\tt. \] \end{lem} \begin{proof} Let $(\phi,\psi,u)\in\cF^{\rD}(\gl)$, and assume that $\phi=t(L+\chi)$ for some $t>0$ and $\chi\in C(\tt)$. Recall that \[\begin{aligned} F_\phi(x,p,X)&\,=\sup_{\ga\in\cA}\left(-\tr a X-b\cdot p -t(L+\chi)\right) =t\sup_{\ga\in\cA}\left(-\tr a t^{-1}X-b\cdot t^{-1}p -L\right)-t\chi \\&\,=tF(x,t^{-1}p,t^{-1}X)-t\chi(x), \end{aligned} \] to find that $u$ is a subsolution of \[ \gl u+tF(x,t^{-1}Du,t^{-1}D^2u)=t\chi \ \ \text{ in }\gO. \] Hence, the function $v:=t^{-1}u$ is a subsolution of \[ \gl v+F[v]=\chi \ \ \text{ in }\gO \] and satisfies $v\leq t^{-1}\psi$ pointwise on $\bry$. We set \[ A=\max\left\{\gl^{-1}\max_{\tt}(\chi-F(\cdot,0,0)),\ t^{-1}\max_{\bry}\psi\right\}, \] and note that the constant function $w:=A$ is a supersolution of \[ \gl w+F[w]=\chi \ \ \text{ in }\gO \ \ \text{ and } \ \ w=t^{-1}\psi \ \ \text{ on }\bry. \] Hence, comparison principle \eqref{CPD} guarantees that $v\leq A$ on $\tt$, which yields \[ \gl u(x)\leq \gl t A = \max\left\{ t\max_{\tt}(\chi-F(\cdot,0,0)),\, \gl\max_{\bry}\psi\right\}\ \ \ \text{ for all }\ x\in\tt. \qedhere \] \end{proof} \begin{proof}[Proof of Theorem \ref{thm1-d}] Let $v^\gl\in C(\tt)$ be a solution of \eqref{D}. Since $\|g\|_{C(\bry)}<M$, we have $(L,g,v^\gl)\in\cF^{\rD}(M,\gl)$. Hence, owing to the definition of $\cG^{\rD}(M,z,\gl)'$, we get \[\begin{aligned} 0&\,\leq \lan \mu_1,\,L-\gl v^\gl(z)\ran+\gl \lan\mu_2,\,g-v^\gl(z)\ran \\&\,= \lan \mu_1,L\ran+\gl\lan \mu_2,g\ran-\gl v^\gl(z) \ \ \text{ for all } (\mu_1,\mu_2)\in\cP^{\rD} \cap \cG^{\rD}(M,z,\gl)', \end{aligned}\] and, therefore, \begin{equation}\label{thm1-d-1} \gl v^\gl(z)\leq \inf_{(\mu_1,\mu_2)\in \cP^{\rD} \cap \cG^{\rD}(M,z,\gl)'} \left(\lan\mu_1,\,L\ran+\gl\lan\mu_2,\,g\ran\right). \end{equation} Next, we show the reverse inequality: \begin{equation}\label{thm1-d-1+} \gl v^\gl(z)\geq \inf_{(\mu_1,\mu_2) \in\cP^{\rD} \cap \cG^{\rD}(M,z,\gl)'} \left(\lan\mu_1,\,L\ran+\gl\lan\mu_2,\,g\ran\right). \end{equation} For this, we suppose toward a contradiction that \begin{equation}\label{thm1-d-2} \gl v^\gl(z)+\ep<\inf_{(\mu_1,\mu_2)\in\cP^{\rD}\cap \cG^{\rD}(M,z,\gl)'} \left(\lan \mu_1, L\ran+\gl\lan\mu_2,\,g\ran\right), \end{equation} for a small $\ep>0$. Now, we fix a compact subset $K$ of $\cA$ as follows. If $\cA$ is compact, then we set $K:=\cA$. Otherwise, pick an $\ga_0\in\cA$ and choose a constant $L_0>0$ so that \begin{equation}\label{ga_0L_0} \max_{x\in\tt}L(x,\ga_0)\leq L_0 \ \ \text{ and } \ \ (\gl+1) |v^\gl(z)|\leq L_0. \end{equation} Then set \begin{equation} \label{C_1C_2} C_1:=M+\|F(\cdot,0,0)\|_{C(\tt)}, \ \ \ C_2:=C_1(1+\gl), \end{equation} \begin{equation}\label{gdL_1} \gd:=\min\Big\{\fr 12,\,\fr\ep{4\gl(M+L_0)}\Big\} \ \ \text{ and } \ \ L_1:=\fr{M+L_0+C_2}{\gd}. \end{equation} Owing to \eqref{L}, we may select a compact set $K_0\subset\cA$ so that \begin{equation}\label{defK_0} L(x,\ga)\geq \max\{L_0,\,L_1\}(\,=L_1\,) \ \ \ \text{ for all }(x,\ga)\in\tt\tim(\cA\setminus K_0). \end{equation} Finally, pick an $\ga_1\in\cA\setminus K_0$ and define the compact set $K\subset\cA$ by \begin{equation}\label{defK} K:=K_0\cup\{\ga_0,\ga_1\}(\,=K_0\cup\{\ga_1\}\,). \end{equation} It follows from \eqref{thm1-d-2} that \begin{equation}\label{thm1-d-2+} \gl v^\gl(z)+\ep<\inf_{(\mu_1,\mu_2)\in\cP^{\rD}_{K} \cap \cG^{\rD}(M,z,\gl)'} \left(\lan \mu_1, L\ran+\gl\lan\mu_2,\,g\ran\right). \end{equation} By Lemma \ref{lem:dirichlet-convex}, $\cG^{\rD}(M,z,\gl)$ is a convex cone with vertex at the origin, and hence, we get \[ \inf_{(f_1,f_2)\in\cG^{\rD}(M,z,\gl)} \left(\lan \mu_1,\,f_1\ran+\gl\lan\mu_2,f_2\ran\right)= \begin{cases} 0 \ \ &\text{ if }\ (\mu_1,\mu_2)\in\cP^{\rD}_{K}\cap\cG^{\rD}(M,z,\gl)', \\ -\infty &\text{ if }\ (\mu_1,\mu_2)\in \cP^{\rD}_{K}\setminus \cG^{\rD}(M,z,\gl)', \end{cases} \] and moreover, \begin{equation}\begin{aligned} \label{thm1-d-2++} \inf_{(\mu_1,\mu_2)\in\cP^{\rD}_{K}\cap \cG^{\rD}(M,z,\gl)'} &\left(\lan \mu_1, L\ran+\gl\lan \mu_2,g\ran\right) \\&\,\kern-30pt= \inf_{(\mu_1,\mu_2)\in\cP^{\rD}_{K}}\ \sup_{(f_1,f_2)\in\cG^{\rD}(M,z,\gl)}\left(\lan \mu_1, L\ran+\gl\lan\mu_2,g\ran -\lan\mu_1,f_1\ran-\gl\lan\mu_2,f_2\ran \right) \\&\,\kern-30pt=\inf_{(\mu_1,\mu_2)\in\cP^{\rD}_{K}}\ \sup_{(f_1,f_2)\in\cG^{\rD}(M,z,\gl)}\,\left(\lan \mu_1, L-f_1\ran+\gl\lan\mu_2,g-f_2\ran\right). \end{aligned} \end{equation} Since $K$ is compact, $\cP^{\rD}_{K}$ is a compact convex subset of $\cR_{L}\tim\cR_2$, with the topology of weak convergence of measures. Also, for any $(f_1,f_2)\in\cG^\rD(M,z,\gl)$, the functional \[ \cP_K^\rD\ni (\mu_1,\mu_2)\mapsto \lan \mu_1, L-f_1\ran+\gl\lan\mu_2,g-f_2\ran\in\R \] is continuous. Thus, Sion's minimax theorem implies \[\begin{aligned} \min_{(\mu_1,\mu_2)\in\cP^{\rD}_{K}} &\ \sup_{(f_1,f_2)\in\cG^{\rD}(M,z,\gl)}\,\left(\lan \mu_1, L-f_1\ran+\gl\lan\mu_2,g-f_2\ran\right) \\&= \sup_{(f_1,f_2)\in\cG^{\rD}(M,z,\gl)}\ \min_{(\mu_1,\mu_2)\in\cP^{\rD}_{K}} \,\left(\lan \mu_1, L-f_1\ran+\gl\lan\mu_2,g-f_2\ran\right), \end{aligned}\] which, together with \eqref{thm1-d-2+} and \eqref{thm1-d-2++}, yields \[ \gl v^\gl(z)+\ep < \ \sup_{(f_1,f_2)\in\cG^{\rD}(M,z,\gl)}\ \min_{(\mu_1,\mu_2) \in\cP^{\rD}_{K}} \,\left(\lan \mu_1, L-f_1\ran+\gl\lan\mu_2,g-f_2\ran\right). \] Thus, we may choose $(\phi,\psi,u)\in\cF^{\rD}(M,\gl)$ and $(t,\chi)\in (0,\,\infty)\tim C(\tt)$ so that $\phi=t(L+\chi)$, $\|\chi\|_{C(\tt)}< M$, and \begin{equation}\label{thm1-d-4} \gl v^\gl(z)+\ep < \min_{(\mu_1,\mu_2)\in\cP^{\rD}_{K}} \,\left(\lan \mu_1, L-\phi+\gl u(z)\ran+\gl\lan\mu_2,g-\psi+u(z)\ran\right). \end{equation} Note that, by the definition of $\cF^{\rD}(M,\gl)$, the inequality $\|\psi\|_{C(\bry)}< tM$ is valid. Since $(0,\gd_x)\in \cP^{\rD}_{K}$ for all $x\in\bry$, we get from \eqref{thm1-d-4} \[ \gl v^\gl(z)+\ep<\gl(g-\psi+u(z))\quad\text{on} \ \bry, \] which reads \begin{equation}\label{thm1-d-4+} \psi< g+(u-v^\gl)(z)-\gl^{-1}\ep \ \ \text{ on }\bry. \end{equation} Also, since $(\gd_{(x,\ga)},0)\in\cP^\rD_K$ for all $(x,\ga)\in \tt\tim K$, we get from \eqref{thm1-d-4} \begin{equation}\label{thm1-d-4++} (t-1)L(x,\ga)+t\chi(x)<\gl (u-v^\gl)(z)-\ep \ \ \ \text{ for all }\ (x,\ga)\in\tt \tim K. \end{equation} We show that there are a constant $\gth>0$ and a subsolution $w=\theta u\in C(\tt)$ to \begin{equation}\label{thm1-d-6} \gl w+F[w]=-\gl (v^\gl-w)(z)-2^{-1}\gth\ep\ \ \text{ in }\gO, \end{equation} and \begin{equation}\label{thm1-d-6+} w\leq g -(v^\gl-w)(z)-(2\gl)^{-1}\gth\ep\ \ \ \text{ on }\bry. \end{equation} Once this is complete, we get a contradiction right away. Indeed, the function $\gz:=w+(v^\gl-w)(z)+(2\gl)^{-1}\gth\ep$ is a subsolution of \eqref{D}, and comparison principle \eqref{CPS} yields $\gz\leq v^\gl$, which, after evaluation at $z$, gives $\,(2\gl)^{-1}\gth\ep\leq 0$. This is a contradiction. Assume that $\cA$ is compact. Then we have $K=\cA$, and therefore, we get from \eqref{thm1-d-4++} \[ \phi=L+(t-1)L+t\chi<-\gl(v^\gl-u)(z)-\ep \ \ \ \text{ on }\ \bb, \] which ensures that the pair of $w:=u$ and $\gth:=1$ satisfies \eqref{thm1-d-6}, while \eqref{thm1-d-6+} for this pair is an immediate consequence of \eqref{thm1-d-4+}. We assume henceforth that $\cA$ is not compact. We split our further argument into two cases. Consider first the case when $t\leq 1$. Recall that $K=K_0\cup\{\ga_0,\ga_1\}$. By \eqref{thm1-d-4++}, we have \begin{equation}\label{temp1} (t-1)L(x,\ga_0)+t\chi(x)<\gl(u-v^\gl)(z)-\ep \ \ \ \text{ for all }\ x\in\tt. \end{equation} By the choice of $\ga_0$ and $L_0$, we have \[ L(x,\ga)\geq L_0\geq L(x,\ga_0) \ \ \ \text{ for all }\ (x,\ga)\in\tt\tim(\cA\setminus K). \] Then we combine this with \eqref{temp1}, to get \[ (t-1)L(x,\ga)+t\chi(x)<\gl(u-v^\gl)(z)-\ep \ \ \ \text{ for all }\ (x,\ga)\in\tt\tim(\cA\setminus K), \] which, furthermore, yields together with \eqref{thm1-d-4++} \[ (t-1)L(x,\ga)+t\chi(x)<\gl(u-v^\gl)(z)-\ep \ \ \ \text{ for all }\ (x,\ga)\in\tt\tim\cA. \] From this, we observe \[ \phi=L+(t-1)L+t\chi<L-\gl (v^\gl-u)(z)-\ep \ \text{ on }\bb, \] which shows together with \eqref{thm1-d-4+} the validity of \eqref{thm1-d-6} and \eqref{thm1-d-6+}, with $w:=u$ and $\gth:=1$. Secondly, we consider the case when $t> 1$. Recall the choice of $L_0$ and $K_0$, to see \[ |\gl v^\gl(z)|\leq L_0\leq L(x,\ga) \ \ \ \text{ for all }\ (x,\ga) \in \tt\tim (\cA\setminus K_0). \] Since $\ga_1\in K\setminus K_0$, this and \eqref{thm1-d-4++} yield \begin{equation}\label{temp2} (t-1)\gl v^\gl(z)+t\chi(x)\leq (t-1)L(x,\ga_1) +t\chi(x)<\gl (u-v^\gl)(z)-\ep, \end{equation} which verifies $t\chi < -t\gl v^\gl(z) + \gl u(z) -\ep$ on $\tt$ and, furthermore, \[ \phi=t(L+\chi) < tL - t \gl v^\gl(z) + \gl u(z) - \ep \quad \text{on} \ \tt \times \cA. \] This shows formally that \[ \gl u+\cL_\ga u\leq \phi<t L-t\gl (v^\gl-t^{-1}u)(z) -\ep \ \ \text{ in }\tt\tim\cA,\] from which we deduce that if we set $w:=t^{-1}u$ and $\gth=t^{-1}$, then \eqref{thm1-d-6} holds. We continue with the case when $t>1$. By Lemma \ref{lem1-d}, in view of \eqref{C_1C_2}, we get \[ \gl u\leq \max\{t(M+\|F(\cdot,0,0)\|_{C(\tt)}),\gl tM\} \leq C_1(1+\gl)t= C_2t, \] and, accordingly, \begin{equation}\label{temp4} \gl (u-v^\gl)(z)\leq C_2t+L_0\leq (C_2+L_0)t. \end{equation} Combining this with the second inequality of \eqref{temp2} and \eqref{defK_0}, we get \[ (t-1) L_1 < Mt+(L_0+C_2)t=(M+L_0+C_2)t. \] Using this and \eqref{gdL_1}, we compute \[ t-1<\fr{(M+L_0+C_2)t}{L_1}=\gd t\leq \fr t2, \] which shows that $t<2$, and we get from the above \begin{equation}\label{temp3} t-1<\gd t\leq 2\gd\leq \fr{\ep}{2\gl(M+L_0)}. \end{equation} Hence, we obtain \[ (t-1)(g-v^\gl(z))\leq (t-1)(M+L_0)<\fr{\ep}{2\gl}, \] and moreover, by \eqref{thm1-d-4+}, \[\begin{aligned} t^{-1}\psi&\,<g+(t^{-1}u-v^\gl)(z)-(t\gl)^{-1}\ep +(t^{-1}-1)(g-v^\gl(z)) \\&\,<g+(t^{-1}u-v^\gl)(z)-(2t\gl)^{-1}\ep \ \ \ \text{ on }\pl\gO, \end{aligned} \] which shows that \eqref{thm1-d-6+} holds with $w:=t^{-1}u$ and $\gth:=t^{-1}$. Therefore, both \eqref{thm1-d-6} and \eqref{thm1-d-6+} hold with $w:=t^{-1}u$ and $\gth:=t^{-1}$. The proof is complete. \end{proof} \begin{thm} \label{thm1-d0} Assume \eqref{F1}, \eqref{F2}, \eqref{CP}, \eqref{L} and \eqref{CPD}. Let $(z,\gl)\in\tt\tim (0,\,\infty)$. If $v^\gl\in C(\tt)$ is a solution of \eqref{D}, then \begin{equation}\label{thm1-d0-0} \gl v^\gl(z)=\min_{(\mu_1,\mu_2)\in \cP^\rD\cap \cG^\rD(z,\gl)'}\,\left( \lan \mu_1,\,L\ran+\gl\lan\mu_2,\,g\ran\right). \end{equation} \end{thm} We introduce the set $\cG^\rD(z,\gl)^\dag$ for $(z,\gl)\in\tt\tim[0,\,\infty)$ (resp., $\cG^\rD(M,z,\gl)^\dag$ for $(M,z,\gl)\in(0,\,\infty)\tt\tim[0,\,\infty)$ ) as the set of triples $(\rho,\mu_1,\mu_2)\in [0,\,\infty)\tim\cP^\rD$ satisfying \[ 0\leq t\rho+\lan\mu_1,tL+\chi-\gl u(z)\ran+\gl\lan\mu_2,\psi-u(z)\ran \] for all $(tL+\chi,\psi,u)\in\cF^\rD(\gl)$ (resp., $(tL+\chi,\psi,u)\in\cF^\rD(M,\gl)$), with $t>0$ and $\chi\in C(\tt)$. Notice that it is required here for $(\rho,\mu_1,\mu_2)\in \cG^\rD(z,\gl)^\dag$ (resp., $(\rho,\mu_1,\mu_2)\in \cG^\rD(M,z,\gl)^\dag$) to fulfill the condition $(\mu_1,\mu_2)\in\cP^\rD$. We remark that the inclusion $(\mu_1,\mu_2)\in \cP^\rD\cap \cG^\rD(z,\gl)'$ (resp., $(\mu_1,\mu_2)\in \cP^\rD\cap \cG^\rD(M,z,\gl)'$) holds if and only if $(0,\mu_1,\mu_2)\in\cG^\rD(z,\gl)^\dag$ (resp., $(0,\mu_1,\mu_2)\in\cG^\rD(M,z,\gl)^\dag$), that $\bigcap_{M>0}\cG^\rD(M,z,\gl)^\dag=\cG^\rD(z,\gl)^\dag$, and that if $N>M$, then $\cG^\rD(N,z,\gl)^\dag\subset\cG^\rD(M,z,\gl)^\dag$. The following lemmas are useful for the proof of Theorem \ref{thm1-d0}. \begin{lem}\label{cpt-d-1} Assume \eqref{F1} and \eqref{L}. Fix $R\in\R$. Then, \emph{(i)}\ the functional \[ (\mu_1,\mu_2)\mapsto \lan\mu_1,L\ran \] is lower semicontinuous on $\cP^\rD$, with the topology of weak convergence of measures, and \emph{(ii)}\ the set \[ \cP^\rD_R:=\{(\mu_1,\mu_2)\in\cP^\rD \mid \lan \mu_1,L\ran \leq R\} \] is compact in the topology of weak convergence of measures. \end{lem} It should be remarked that, by definition, a sequence $\{\mu_1^k,\mu_2^k\}_{k\in\N}\subset \cR_1\tim\cR_2$ converges to a $(\mu_1,\mu_2)\in\cR_1\tim\cR_2$ weakly in the sense of measures if and only if \[ \lim_{k\to\infty}\left(\lan\mu_1^k,\phi\ran+ \lan\mu_2^k,\psi\ran\right) =\lan\mu_1,\phi\ran+ \lan\mu_2,\psi\ran \ \ \ \text{ for all }\ (\phi,\psi) \in C_c(\bb)\tim C(\tt). \] \begin{proof} Let $\cX$ denote the disjoint union of $\bb$ and $\bry$, and note that $\cX$ has a natural metric, for instance, the metric $d$ on $\cX$ given by the formula \[ d(\xi,\eta)= \begin{cases} d_1(\xi,\eta) \ \ &\text{ if }\ \xi,\eta\in \bb,\\ d_2(\xi,\eta) &\text{ if }\ \xi,\eta\in\bry,\\ 1 &\text{ otherwise}, \end{cases} \] where $d_1$ is the given metric on $\bb$ and $d_2$ is the Euclidean metric on $\R^n$. With this metric structure, $\cX$ is $\gs$-compact and locally compact. Note that $B\subset\cX$ is a Borel set if and only if $B\cap \bb$ and $B\cap\bry$ are Borel sets in $\bb$ and $\bry$, respectively. We define $f\mid \cX\to \R$ by \[ f(\xi)= \begin{cases} L(\xi) \ &\text{ if }\ \xi\in\bb,\\ 0 &\text{ if }\ \xi\in\bry, \end{cases}\] and set \[ \cP_{\cX,R}:=\{\mu\in\cP_{\cX}\mid \lan\mu, f\ran\leq R\}, \] where $\cP_\cX$ is defined as the set of all Radon probability measures on $\cX$, Note that $f=+\infty$ at infinity, and observe by Lemmas \ref{basic-lsc} and \ref{basic-cpt} that, in the topology of weak convergence of measures, the functional $\mu\mapsto \lan \mu,f\ran$ is lower semicontinuous on $\cP_{\cX}$, and $\cP_{\cX,R}$ is compact. For $(\mu_1,\mu_2)\in\cP^{\cD}$, if we put \[ \tilde\mu(B):=\mu_1(B\cap\bb)+\mu_2(B\cap\bry) \ \ \text{ for any Borel set }B\subset\cX, \] then $\tilde \mu$ defines a (unique) Radon probability measure on $\cX$. With this notation, it is easy to see that \[ \lan\tilde \mu, f\ran=\lan\mu_1,L\ran \ \ \ \text{ and } \ \ \ \cP_{\cX,R}=\{\tilde\mu\mid (\mu_1,\mu_2)\in\cP^\rD_R\}. \] Hence, we conclude that, in the topology of weak convergence of measures, the functional $(\mu_1,\mu_2)\mapsto \lan\mu_1,L\ran$ is lower semicontinuous on $\cP^\rD$ and the set $\cP_R^{\rD}$ is compact. \end{proof} \begin{lem} \label{d-cpt} Assume \eqref{F1}, \eqref{F2} and \eqref{L}. Let $(M,z,\gl)\in(0,\,\infty)\tim\tt\tim[0,\,\infty)$ and let $\{\gl_j\}_{j\in\N}\subset[0,\,\infty)$ be a sequence converging to $\gl$. Let $\{(\mu_1^j,\mu_2^j)\}_{j\in\N}$ be a sequence of measures such that $(\mu_1^j,\mu_2^j)\in\cP^\rD\cap \cG^\rD(M,z,\gl_j)'$ for all $j\in\N$. Assume that the sequence $\{\du{\mu_1^j, L}\}_{j\in\N}$ is convergent and that $\{(\mu_1^j,\mu_2^j)\}_{j\in\N}$ converges to some $(\mu_1^0,\mu_2^0)\in\cP^\rD$ weakly in the sense of measures. \emph{(i)}\ If we set \[\rho=\lim_{j\to\infty}\du{\mu_1^j,L}-\du{\mu_1^0,L},\] then $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(M,z,\gl)^\dag$. \emph{(ii)}\ Assume in addition that $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(z,\gl)^\dag$. If either $\cA$ is compact or $\mu_1^0\not=0$, then there exist a pair $(\nu_1,\nu_2)\in\cP^\rD\cap \cG^\rD(z,\gl)'$ of measures such that $\,\lan\nu_1,L\ran=\rho+\du{\mu_1^0,L}\,$ and, for all $\, (\psi,\eta)\in C(\tt)\tim C(\bry)$, \[ \lan\mu_1^{0},\psi\ran+\du{\mu_2^{0},\eta} =\lan\nu_1,\psi\ran+\du{\nu_2,\eta}. \] \end{lem} The proof of the lemma above is similar to that of Lemma \ref{sc-cpt}, but we give the proof for completeness. \begin{proof} Note first that the lower semicontinuity of the functional $(\mu_1,\mu_2)\mapsto\lan\mu_1,L\ran$ on $\cP^\rD$, as claimed by Lemma \ref{cpt-d-1}, implies that $\rho\geq 0$. To check the property that $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(M,z,\gl)^\dag$, let $f=(f_1,f_2)\in\cG^\rD(M,z,\gl)$, and select $(\phi,\psi,u)\in\cF^\rD(\gl)$, $t>0$ and $\chi\in C(\tt)$ so that $f=(\phi-\gl u(z),\gl(\psi-u(z)))$ and $\phi=tL+\chi$. Let $j\in\N$ and note that $(\phi+(\gl_j-\gl)u,\psi,u)\in \cF^\rD(\gl_j)$. Hence, if $j$ is large enough, then $(\phi+(\gl_j-\gl)u,\psi,u)\in \cF^\rD(M, \gl_j)$ and we get \[\begin{aligned} 0&\,\leq \lan\mu_1^j,\phi+(\gl_j-\gl)u-\gl_j u(z)\ran+ \du{\mu_2^j,\gl_j(\psi-u(z))} \\&\,=t\lan\mu_1^j,L\ran+ \lan\mu_1^j,\chi+(\gl_j-\gl)u-\gl_j u(z)\ran +\gl_j\du{\mu_2^j,\psi-u(z)}. \end{aligned}\] Sending $j\to\infty$ yields \[ 0\leq t(\rho+\du{\mu_1^0,L}) +\lan\mu_1^0,\chi-\gl u(z)\ran+\du{\mu_2^0,\gl(\psi-u(z))}, \] which ensures that $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(M,z,\gl)^\dag$, proving assertion (i). Next, we assume that $(\rho,\mu_1^0,\mu_2^0)\in\cP^\rD\cap\cG^\rD(z,\gl)'$ and show the existence of $(\nu_1,\nu_2)$ having the properties described in assertion (ii). Since $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(z,\gl)^\dag$, we have \begin{equation}\label{d-cpt-1} 0\leq t\rho +\lan\mu_1^0,tL+\chi-\gl u(z)\ran+\du{\mu_2^0,\gl(\psi-u(z))} \end{equation} for all $(\phi,\psi,u)\in\cF^\rD(\gl)$, where $\phi=tL+\chi$, $t>0$ and $\chi\in C(\tt)$. If $\cA$ is compact, the weak convergence of $\{(\mu_1^j,\mu_2^j)\}$ implies that $\rho=0$. Thus, in this case, the pair $(\nu_1,\nu_2):=(\mu_1^0,\mu_2^0)$ has all the required properties. Now, assume that $\cA$ is not compact and $\mu_1^0\not=0$. Lemma \ref{mod} ensures that there is $\tilde\mu_1^0\in\cR_L^+$ such that $\tilde\mu_1^0(\bb)=\mu_1^0(\bb)$, $\lan\tilde\mu_1^0,L\ran=\rho$ and $\lan\tilde\mu_1^0,\psi\ran=\lan\mu_1^0,\psi\ran$ for all $\psi\in C(\tt)$. We define $(\nu_1,\nu_2)\in\cR_L^+\tim\cR_2^+$ by $(\nu_1,\nu_2)=(\tilde\mu_1^0,\mu_2)$. It is obvious that $(\nu_1,\nu_2)\in \cP^\rD$, $\lan\nu_1,L\ran=\rho+\du{\mu_1^0,L}$ and $\lan\nu_1,\psi\ran+\du{\nu_1,\eta}=\lan\mu_1^0,\psi\ran+\du{\mu_2^0,\eta}$ for all $(\psi,\eta)\in C(\tt)\tim C(\bry)$. These properties and inequality \eqref{d-cpt-1} imply that $(\nu_1,\nu_2)\in \cG^\rD(z,\gl)'$, which completes the proof. \end{proof} \begin{proof}[Proof of Theorem \ref{thm1-d0}] Since $\cG^\rD(z,\gl)'\subset \cG^\rD(M,z,\gl)'$ for any $M>0$, Theorem \ref{thm1-d} yields \begin{equation} \label{thm1-dp-1} \gl v^\gl(z)\leq \inf_{(\mu_1,\mu_2)\in\cP^\rD\cap\cG^\rD(z,\gl)'} \left(\du{\mu_1,L}+\gl\du{\mu_2,g}\right). \end{equation} To prove the reverse inequality of \eqref{thm1-dp-1}, in view of Theorem \ref{thm1-d}, we may select a sequence $\{(\mu_1^k, \mu_2^k)\}_{k\in\N}$ of pairs of measures on $\bb$ and on $\bry$ so that for all $k\in\N$, \begin{equation}\label{thm1-d-00} \gl v^\gl(z)+\fr 1k>\lan\mu_1^k,L\ran+\gl\lan\mu_2^k,g\ran \ \ \ \text{ and } \ \ \ (\mu_1^k,\mu_2^k)\in\cP^\rD\cap\cG^\rD(k,z,\gl)'. \end{equation} Thanks to Lemma \ref{cpt-d-1}, there exists a subsequence $\{(\mu_1^{k_j},\mu_2^{k_j})\}_{j\in\N}$ of $\{(\mu_1^k,\mu_2^k)\}$ such that $\{(\mu_1^{k_j},\mu_2^{k_j})\}_{j\in\N}$ converges to some $(\mu_1^0,\mu_2^0)\in \cP^\rD$ weakly in the sense of measures. Since $\{\du{\mu_1^k,L}\}_{k\in\N}$ is bounded, we may assume that $\{\du{\mu_1^{k_j},L}\}_{j\in\N}$ is convergent. We set $\rho=\lim_{j\to\infty}\du{\mu_1^{k_j},L}-\du{\mu_1^0,L}$, and note that \begin{equation}\label{thm1-dp-2} \gl v^\gl(z)\geq \lim_{k\to\infty}\du{\mu_1^k,L}+\gl\du{\mu_2^0,g} =\rho+\du{\mu_1^0,L}+\gl\du{\mu_2^0,g}, \end{equation} and, by Lemma \ref{d-cpt}, that $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(m,z,\gl)^\dag$ for all $m\in\N$, which implies that $(\rho,\mu_1^0,\mu_2^0)\in\cG^\rD(z,\gl)^\dag$. If either $\cA$ is compact or $\mu_1^0\not=0$, then Lemma \ref{d-cpt} guarantees that there exists $(\nu_1,\nu_2)\in\cP^\rD\cap\cG^\rD(z,\gl)'$ such that $\du{\nu_1,L}=\rho+\du{\mu_1^0,L}$ and $\du{\nu_1,\psi}+\du{\nu_2,\eta}=\du{\mu_1^0,\psi}+\du{\mu_2^0,\eta}$ for all $(\psi,\eta)\in C(\tt)\tim C(\bry)$. These identities combined with \eqref{thm1-dp-2} yield \[ \gl v^\gl(z)\geq \du{\nu_1,L}+\gl\du{\nu_2,g}, \] which shows that \eqref{thm1-d0-0} holds, with $(\nu_1,\nu_2)$ being a minimizer of the right hand side of \eqref{thm1-d0-0}. It remains the case when $\cA$ is not compact and $\mu_1^0=0$. To treat this case, we observe first that $z\in\bry$ and $\mu_2^0=\gd_z$. Indeed, otherwise, there exists $\gz\in C^2(\tt)$ such that $\du{\mu_2^0,\gz-\gz(z)}\not=0$. By replacing $\gz$ by a constant multiple of $\gz$, we may assume that $\rho+\gl\du{\mu_2^0,\gz-\gz(z)}<0$. Noting that $(\gl\gz+L+F[\gz],\gz,\gz)\in\cF^\rD(\gl)$, we get \[ 0\leq \rho+\du{\mu_1^0,\gl\gz+L+F[\gz]-\gl\gz(z)}+\gl\du{\mu_2^0,\gz-\gz(z)}=\rho+\gl\du{\mu_2^0,\gz-\gz(z)}, \] which contradicts the choice of $\gz$. Thus, we have $z\in\bry$ and $\mu_2^0=\gd_z$. Since \[ \du{\mu_1^0,\phi-\gl u(z)}+\gl\du{\mu_2^0,\psi-u(z)} =\gl(\psi-u)(z)\geq 0 \] for any $(\phi,\psi,u)\in \cF^\rD(\gl)$, we have $(0,\mu_1^0,\mu_2^0)\in\cG^\rD(z,\gl)^\dag$, which implies that $(\mu_1^0,\mu_2^0)\in\cP^\rD\cap\cG^\rD(z,\gl)'$. Thus, we find from \eqref{thm1-dp-1} and \eqref{thm1-dp-2} that \eqref{thm1-d0-0} holds with $(\mu_1^0,\mu_2^0)$ as a minimizer of the right hand side of \eqref{thm1-d0-0}. The proof is complete. \end{proof} \subsection{Formula for the critical value} According to Propositions \ref{prop1-d0} and \ref{prop1-d}, under the hypotheses \eqref{F1}, \eqref{F2}, \eqref{CPD}, \eqref{SLD} and \eqref{EC}, there exists a unique number $c^*\in[0,\,\infty)$ such that, if $c^*=0$, then (ED$_0$) has a solution in $C(\tt)$ and, if $c^*>0$, then (ES$_{c^*}$) has a solution in $C(\tt)$. We call this number $c^*$ the critical value for the vanishing discount problem for \eqref{D} and denote it by $c_\rD$. Then, owing to Proposition \ref{prop1-d0}, we have \begin{equation}\label{lim-cD} \lim_{\gl\to 0+}\gl v^\gl(x)=-c_\rD \ \ \ \text{ uniformly on }\ \tt, \end{equation} where $v^\gl$ is the solution of \eqref{D}. The following proposition gives a formula for the critical value $c_\rD$ similar to \eqref{critical-value-s}. \begin{prop}\label{prop-d-cv} Assume \eqref{F1}, \eqref{F2}, \eqref{L}, \eqref{CPD}, \eqref{SLD}, and \eqref{EC}. Then, \begin{equation}\label{critical-value-d-1} c_\rD=\min\{c\geq 0\mid \emph{(ED$_c$)} \text{ has a solution in }C(\tt)\}. \end{equation} \end{prop} \begin{proof} We set \[ d:=\inf\{c\geq 0\mid \text{(ED$_c$)} \text{ has a solution in }C(\tt)\}. \] By the definition of $c_\rD$ above, which is based on Propositions \ref{prop1-d} and \ref{prop1-d0}, we have \begin{equation}\label{cD=lim} \lim_{\gl\to 0}(-\gl v^\gl(x))=c_\rD \ \ \ \text{ uniformly on }\ \tt. \end{equation} Moreover, Proposition \ref{prop1-d} ensures that, if $c_\rD=0$, then (ED$_{c_{\rD}}$) has a solution in $C(\tt)$ and, if $c_\rD>0$, then (ES$_{c_\rD}$) has a solution in $C(\tt)$. Let $u\in C(\tt)$ be a solution of (ED$_0$) and (ES$_{c_\rD}$), respectively, if $c_\rD=0$ and $c_\rD>0$. Note that, if $c_\rD>0$, then the function $u-C$, with $C>0$ chosen so large that $u-C\leq g$ on $\bry$, is a solution of (ED$_{c_\rD}$). It is now clear that $d\leq c_\rD$. It is enough to show that $d\geq c_\rD$. Suppose to the contrary that $d<c_\rD$, and get a contradiction. We choose $(v,c)\in C(\tt)\tim[d,c_\rD)$ so that $v$ is a subsolution of (ED$_c$). Fix any $C>0$ so that $u-C\leq g$ on $\bry$. Since $c_\rD>0$, $u$ is a solution of (ES$_{c_\rD}$), and the function $w:=u-C$ is a supersolution of (ED$_{c_\rD}$). Fix $c_0\in(c,\,c_\rD)$ and select $\gl>0$ sufficiently small so that $\gl v\leq c_0-c$ and $-\gl w\leq c_\rD-c_0$ on $\tt$, which means that $v$ and $w$ are a subsolution and a supersolution, respectively, of \eqref{D}, with $L$ replaced by $L+c_0$. By \eqref{CPD}, we get $v\leq w=u-C$ on $\tt$, but this gives a contradiction when $C$ is sufficiently large. \end{proof} Another formula for the critical value $c_\rD$ is stated in the next theorem. Henceforth, we write \[ \cP^\rD_1=\{\mu_1\mid (\mu_1,\mu_2)\in\cP^\rD \cap\cG^\rD(0)'\} \ \ \text{ and } \ \ \cP^\rD_{1,0}=\{\mu\in\cP^\rD_1\mid \mu(\tt\tim\cA)=1\}. \] A crucial remark here is that \begin{equation}\label{equiDS} \cP^\rD_{1,0}=\cP^\rS\cap\cG^\rS(0)'. \end{equation} Indeed, the argument below guarantees the validity of the identity above. It is obvious that, for $\phi\in C(\bb)$, inclusion $(\phi,\psi,u)\in \cF^\rD(0)$ holds for some $(\psi,u)\in C(\bb)\tim C(\bry)$ if and only if $(\phi,u,u)\in\cF^\rD(0)$. Hence, for $\phi\in C(\bb)$, we have $(\phi,0)\in\cG^\rD(0)$ if and only if $(\phi,u,u)\in\cF^\rD(0)$ for some $u\in C(\tt)$, which is if and only if $(\phi,u)\in\cF^\rS(0)$ for some $u\in C(\tt)$, and moreover, this is if and only if $\phi\in\cG^\rS(0)$. Using these observations, it is easy to see that, for any $\mu\in \cP_L$, $\mu\in\cP^\rS\cap\cG^\rS(0)'$ if and only if $\mu\in\cG^\rS(0)'$, which is if and only if $\lan\mu,\phi\ran\geq 0$ \ for all $\, (\phi,0)\in\cG^\rD(0)$. This is equivalent to the condition that $(\mu,0)\in\cP^\rD\cap\cG^\rD(0)'$, which verifies that \eqref{equiDS} is valid. \begin{thm} \label{thm3-d} Assume \eqref{F1}, \eqref{F2}, \eqref{CP}, \eqref{L}, \eqref{CPD}, \eqref{SLD}, and \eqref{EC}. Then, \begin{equation}\label{thm3-d-main} -c_\rD =\min_{\mu\in\cP^\rD_1}\lan\mu,L\ran. \end{equation} Furthermore, if $c_\rD>0$, then \begin{equation}\label{thm3-d-main'} -c_\rD=\min_{\mu\in\cP^\rS\cap\cG^\rS(0)'} \lan\mu,L\ran. \end{equation} \end{thm} \begin{proof} We show first that \eqref{thm3-d-main} is valid. Let $(\mu_1,\mu_2)\in\cP^\rD\cap\cG^\rD(0)'$ and $v\in C(\tt)$ be a solution of (ED$_{c_\rD}$). The existence of such a $v$ is guaranteed by Proposition \ref{prop-d-cv}. Noting that $(L+c_\rD,v,v)\in \cF^\rD(0)$, we get \[ 0\leq \lan\mu_1,L+c_\rD\ran\leq \lan\mu_1,L\ran+c_\rD, \] which yields \begin{equation}\label{thm3-d-mm} -c_\rD\leq \inf_{\mu\in\cP^\rD_1}\lan\mu,L\ran. \end{equation} Pick a point $z\in\tt$ and a sequence $\{\gl_j\}_{j\in\N} \subset(0,\,\infty)$ converging to zero. For $j\in\N$, owing to Theorem \ref{thm1-d0}, we select $(\mu_1^j,\mu_2^j)\in \cP^{\rD}\cap \cG^\rD(z,\gl_j)'$ so that \begin{equation}\label{thm3-d-0} \gl_j v^{\gl_j}(z)=\lan\mu_1^j,L\ran +\gl_j\lan\mu_2^j,g\ran. \end{equation} Note that the sequence $\{\lan\mu_1^j,L\ran\}_{j\in\N}$ is bounded. We apply Lemma \ref{cpt-d-1}, with $\gl=0$, to find that there is a subsequence of $\{(\gl_{j},\mu_1^j,\mu_2^j)\}$, which we denote again by the same symbol, such that the sequence $\{(\mu_1^j, \mu_2^j)\}_{j\in\N}$ converges weakly in the sense of measures to some $(\mu_1^0,\mu_2^0)\in\cP^\rD$, the sequence $\{\lan\mu_1^j,L\ran\}_{j\in\N}$ is convergent, and \[ \lim_{j\to\infty}\lan\mu_1^j,L\ran\geq \lan\mu_1^0,L\ran. \] Set \[ \rho^0=\lim_{j\to\infty} \du{\mu_1^j,L}-\lan\mu_1^0,L\ran, \] and note by Lemma \ref{d-cpt} that $(\rho^0,\mu_1^0,\mu_2^0)\in \cG^\rD(0)^\dag$. Note also that if $\cA$ is compact, then $\rho^0=0$. From \eqref{thm3-d-0} and \eqref{cD=lim}, we get \begin{equation}\label{thm3-d-00} -c_\rD=\rho^0+\lan\mu_1^0,L\ran. \end{equation} If $\rho^0=0$, then we have \begin{equation}\label{thm3-d-m4} -c_\rD=\lan\mu_1^0,L\ran, \end{equation} and $(\mu_1^0,\mu_2^0)\in\cP^\rD\cap\cG^\rD(0)'$. These observations combined with \eqref{thm3-d-mm} show that, if $\rho^0=0$, \eqref{thm3-d-main} holds. Assume instead that $\rho^0\not=0$. Then, $\rho^0>0$ and $\cA$ is not compact. Since $c_\rD\geq 0$ by Proposition \ref{prop1-d0}, we see from \eqref{thm3-d-00} that $\mu_1^0\not=0$. By Lemma \ref{d-cpt}, there exists $(\nu_1,\nu_2)\in\cP^\rD\cap\cG^\rD(0)'$ such that $\rho^0+\du{\mu_1^0,L}=\du{\nu_1,L}$. This together with \eqref{thm3-d-00} and \eqref{thm3-d-mm} proves that, if $\rho^0\not=0$, then \eqref{thm3-d-main} holds. Thus, we conclude that \eqref{thm3-d-main} is always valid. Finally, we consider the case $c_\rD>0$ and prove \eqref{thm3-d-main'}. We fix any minimizer $\mu\in\cP^\rD_1$ of the minimization problem \eqref{thm3-d-main}, and show that $\mu(\bb)=1$, which means $\mu\in\cP^\rD_{1,0}$ and completes the proof of \eqref{thm3-d-main'}. The condition for $\nu\in\cR_L^+$ to be in $\cP^\rD_1$ is described by the inequalities $\nu(\bb)\leq 1$ and \begin{equation}\label{thm3-d-m5} 0\leq t\lan\nu,L\ran+\lan\nu,\chi\ran \end{equation} for all $(tL+\chi,\psi,u)\in\cF^\rD(0)$, where $t>0$ and $\chi\in C(\tt)$. Recall that $\lan\mu,L\ran=-c_\rD<0$. Suppose for the moment that $\mu(\bb)<1$. By choosing $\gth>1$ so that $\gth\mu(\bb)\leq 1$, we get a new measure $\nu:=\gth\mu\in\cR_L^+$ that satisfies \eqref{thm3-d-m5}, which shows that $\nu\in\cP^\rD_1$, and $\lan\nu,L\ran=-\gth c_\rD<-c_\rD$. This is a contradiction, which proves that $\mu(\bb)=1$. \end{proof} \begin{definition}We call $\hat\mu\in\cP_1^\rD$ a \emph{viscosity Mather measure} if it satisfies $\,\du{\hat\mu,L}=\inf_{\mu\in\cP_1^\rD}\du{\mu,L}$, and denote by $\cM^\rD$ the set of all viscosity Mather measures $\mu \in \cP_1^\rD$. For $(z,\gl)\in\tt\tim(0,\,\infty)$, we denote by $\cM^{\rD}(z,\gl)$ the set of all measures $(\hat\mu_1,\hat\mu_2)\in\cP^\rD\cap\cG^{\rD}(z,\gl)'$ that satisfies \[ \lan \hat\mu_1,\,L\ran+\gl\lan\hat\mu_2,\,g\ran =\inf_{(\mu_1,\mu_2)\in\cP^\rD\cap\cG^{\rD}(z,\gl)'} \,\left(\lan \mu_1,\,L\ran+\gl\lan\mu_2,\,g\ran\right), \] and call $\gl^{-1}(\mu_1,\mu_2)$, with $(\mu_1,\mu_2)\in\cM^{\rD}(z,\gl)$, a \emph{viscosity Green measure} associated with \eqref{D}. \end{definition} \subsection{Convergence with vanishing discount} \begin{thm}\label{thm2-d} Assume \eqref{F1}, \eqref{F2}, \eqref{L}, \eqref{CP}, \eqref{CPD}, \eqref{SLD} and \eqref{EC}. For each $\gl>0$, let $v^\gl\in C(\tt)$ be the unique solution of \eqref{D}. Then, the family $\{v^\gl+\gl^{-1}c_{\rD}\}_{\gl>0}$ converges, as $\gl\to 0$, to a function $u$ in $C(\tt)$. Furthermore, the function $u$ is a solution of \emph{(ED$_{0}$)} or \emph{(ES$_{c_\rD}$)}, if $c_\rD=0$ or $c_\rD>0$, respectively. \end{thm} \begin{proof} By Propositions \ref{prop1-d0} and \ref{prop1-d}, we find that there is a solution $v\in C(\tt)$ of (ED$_0$) or (ES$_{c_\rD}$), respectively, if $c_\rD=0$ or $c_\rD>0$. Fix such a function $v$, and argue as in the proof of Proposition \ref{prop1-d}, to show that $\{v^\gl+\gl^{-1}c_\rD\}_{\gl>0}$ is uniformly bounded on $\tt$. Indeed, if $c_\rD=0$, then the functions $v+\|v\|_{C(\tt)}$ and $v-\|v\|_{C(\tt)}$ are a supersolution and a subsolution of \eqref{D} for any $\gl>0$ and \eqref{CPD}, which implies that \[ v-\|v\|_{C(\tt)}\leq v^\gl\leq v+\|v\|_{C(\tt)} \ \ \ \text{ on }\ \tt\ \text{ for any }\ \gl>0, \] and, hence, the uniform boundedness of $\{v^\gl\}_{\gl>0}$ on $\tt$ in the case $c_\rD=0$. If $c_\rD>0$ and if $\gl>0$ is sufficiently large, then we observe that the functions $v+\|v\|_{C(\tt)}-\gl^{-1}c_\rD$ and $v-\|v\|_{C(\tt)}-\gl^{-1}c_\rD$ are a supersolution and a subsolution of \eqref{D} and, therefore, that \[ v-\|v\|_{C(\tt)}-\gl^{-1}c_\rD \leq v^\gl\leq v+\|v\|_{C(\tt)}-\gl^{-1}c_\rD, \] which shows the uniform boundedness of $\{v^\gl+\gl^{-1}c_\rD\}_{\gl>0}$ on $\tt$. Thus, together with \eqref{EC}, the family $\{v^\gl+\gl^{-1}c_\rD\}_{\gl>0}$ is relatively compact in $C(\tt)$. We denote by $\cU$ the set of accumulation points in $C(\tt)$ of $\{v^\gl+\gl^{-1}c_\rD\}$ as $\gl\to 0+$. The relative compactness in $C(\tt)$ of the family $\{v^\gl+\gl^{-1}c_\rD\}_{\gl>0}$ ensures that $\cU\not=\emptyset$. In order to prove the convergence, as $\gl\to 0$, of the whole family $\{v^\gl+\gl^{-1}c_\rD\}_{\gl>0}$, it is enough to show that $\cU$ has a unique element. Let $v,w\in\cU$, and we demonstrate that $v=w$. For this, we select sequences $\{\gl_j\}_{j\in\N}$ and $\{\gd_j\}_{j\in\N}$ of positive numbers converging to zero such that \[ \lim_{j\to\infty}v^{\gl_j}+\gl_j^{-1}c_\rD=v\ \ \text{ and } \ \ \lim_{j\to\infty}v^{\gd_j}+\gd_j^{-1}c_\rD=w \ \ \ \text{ in }\ C(\tt). \] As in the last part of the proof of Proposition \ref{prop1-d}, we deduce that $v$ and $w$ are solutions of (ED$_0$) or (ES$_{c_\rD}$) if $c_\rD=0$ or $c_\rD>0$, respectively. Next, fix a point $z\in\tt$ and, owing to Theorem \ref{thm1-d0}, select $(\mu_1^j,\mu_2^j)\in\cM^\rD(z,\gl_j)$ for every $j\in\N$ so that \begin{equation}\label{thm3-d-1} \gl_j v^{\gl_j}(z)=\lan\mu_1^j,L\ran+\gl_j\lan\mu_2^j,g\ran \ \ \ \text{ for all }\ j\in\N. \end{equation} Arguing as in the proof of Theorem \ref{thm3-d}, we may assume, after passing to a subsequence if necessary, that there is $(\mu_1,\mu_2)\in\cP^\rD\cap \cG^\rD(0)'$ such that \begin{align}\label{thm3-d-2} &\lim_{j\to\infty}\lan\mu_1^j,L\ran= \lan\mu_1,L\ran=-c_\rD, \\&\lim_{j\to\infty}\lan\mu_1^j,\psi\ran=\lan\mu_1,\psi\ran \ \ \ \text{ for all }\ \psi\in C(\tt),\label{thm3-d-3} \\&\lim_{j\to\infty}\lan\mu_2^j,h\ran=\lan\mu_2,h\ran \ \ \ \text{ for all }\ h\in C(\bry).\label{thm3-d-4} \end{align} Now, $w$ is a solution of either (ED$_{c_\rD}$) or (ES$_{c_\rD}$), and thus, we have $(L+c_\rD+\gl_jw,w,w)\in \cF^\rD(\gl_j)$ for all $j\in\N$. Also, we have $(L-\gd_jv^{\gd_j},v^{\gd_j},v^{\gd_j})\in\cF^\rD(0)$ for all $j\in\N$. Hence, we get \begin{align} &0\leq \lan\mu_1^j,L+c_\rD+\gl_j w-\gl_jw(z)\ran +\gl_j\lan \mu_2^j,w-w(z)\ran,\label{thm3-d-5} \\&0\leq \lan\mu_1,L-\gd_j v^{\gd_j}\ran. \label{thm3-d-6} \end{align} Sending $j\to\infty$ in \eqref{thm3-d-5} yields together with \eqref{thm3-d-2} \[ 0\leq \lan\mu_1,L\ran+c_\rD\mu_1(\bb) =c_\rD(-1+\mu_1(\bb))\leq 0, \] which shows that \begin{equation}\label{thm3-d-7} c_\rD(-1+\mu_1(\bb))=0. \end{equation} Combining this, \eqref{thm3-d-2} and \eqref{thm3-d-6}, we compute that for any $j\in\N$, \[ 0\leq -c_\rD+\lan \mu_1^j,-\gd_jv^{\gd_j}\ran =-\lan\mu_1,c_\rD+\gd_j v^{\gd_j}\ran, \] which reads \[ \lan\mu_1,v^{\gd_j}+\gd_j^{-1}c_\rD\ran\leq 0. \] Hence, in the limit $j\to\infty$, \begin{equation}\label{thm3-d-8} \lan\mu_1,w\ran\leq 0. \end{equation} Moreover, combining \eqref{thm3-d-1}, \eqref{thm3-d-5} and \eqref{thm3-d-7} gives \[ 0\leq \gl_j\left(v^{\gl_j}+ \gl_j^{-1}c_\rD +\lan\mu_1^j,w\ran -w(z)+\lan\mu_2^j,w-g\ran\right), \] from which, after dividing by $\gl_j$ and taking the limit $j\to\infty$, we get \[ 0\leq v(z)+\lan\mu_1,w\ran-w(z). \] This and \eqref{thm3-d-8} show that $w(z)\leq v(z)$. Because $v,w\in\cU$ and $z\in\tt$ are arbitrary, we find that $\cU$ is a singleton. \end{proof} \section{Neumann problem} \label{sec-n} We consider the Neumann boundary problem in this section. As above, we relabel \eqref{DP} as \eqref{N}, and \eqref{E} as \eqref{EN}. The letter N in \eqref{N} and \eqref{EN} indicates ``Neumann". Let $\gamma$ be a continuous vector field on $\pl\gO$ pointing outward from $\gO$, and $g\in C(\pl\gO)$ be a given function. The Neumann problems of interest are \[\tag{N$_{\gl}$}\label{N} \begin{cases} \gl u+F[u]= 0 \ \ &\text{ in }\ \gO, \\[3pt] \gamma\cdot Du=g \ \ &\text{ on }\ \bry, \end{cases} \] for $\gl>0$, and \begin{equation} \tag{EN}\label{EN}\begin{cases} F[u]=c \ \ &\text{ in } \gO,\\ \gamma\cdot Du=g \ \ &\text{ on }\pl\gO, \end{cases}\end{equation} As be fore, we refer as (EN$_c$) this problem with a given constant $c$. The function $g\in C(\bry)$ is fixed throughout this section. We need to consider the Neumann boundary problem with general datum $(\gl,\phi,\psi) \in[0,\,\infty)\tim \Phi^+\tim C(\bry)$: \begin{equation}\tag{N$_{\gl,\phi,\psi}$}\label{N'} \begin{cases} \gl u+F_\phi[u]=0 \ \ \text{ in }\ \gO,&\\[3pt] \gamma\cdot Du=\psi \ \ \ \text{ on }\ \bry, & \end{cases} \end{equation} In addition to \eqref{F1}, \eqref{F2}, \eqref{L} and \eqref{EC}, we introduce a few assumptions proper to the Neumann boundary condition. \[\tag{OG}\label{OG} \left\{\text{ \begin{minipage}{0.85\textwidth} $\gO$ has $C^1$-boundary and $\gamma$ is oblique to $\pl\gO$ in the sense that $\gamma\cdot \mathbf{n}>0$ on $\pl\gO$, where $\mathbf{n}$ denotes the outward unit normal to $\gO$. \end{minipage} }\right. \] \[\tag{CPN}\label{CPN} \left\{\text{ \begin{minipage}{0.85\textwidth} For any $\gl>0$ and $(\phi,\psi)\in\Psi^+$, the comparison principle holds for \eqref{N'}, that is, if $v\in C(\tt)$ and $w\in C(\tt)$ are a subsolution and a supersolution of \eqref{N'}, respectively, then the inequality $\,v\leq w\,$ holds on $\tt$. \end{minipage} }\right. \] The following is a local version of the hypothesis above. \[\tag{CPN$_{\rm loc}$}\label{CPN'} \left\{\text{ \begin{minipage}{0.82\textwidth} For any $\gl>0$ and $(\phi,\psi)\in\Psi^+$, the localized comparison principle holds for \eqref{N'}, that is, for any relative open subset $V$ of $\tt$, if $v\in C(\lbar V)$ and $w\in C(\lbar V)$ are a subsolution and a supersolution, respectively, of \[ \gl u+F_\phi[u]=0 \ \ \text{ in }\ V\cap \gO \quad\text{ and }\quad \gamma\cdot Du=\psi \ \ \ \text{ on }\ V\cap\pl\gO, \] and if $\,v\leq w\,$ on $\gO\cap\pl V$, then the inequality $\,v\leq w\,$ holds on $\lbar V$. \end{minipage} }\right. \] \[\tag{SLN}\label{SLN} \text{ \begin{minipage}{0.85\textwidth} For any $\gl > 0$, \eqref{N} admits a solution $v^\gl \in C(\tt)$. \end{minipage} }. \] Condition \eqref{OG} guarantees that there exists a $C^\infty$-function $\gz$ on $\lbar\gO$ such that \begin{equation}\label{gz} \gamma\cdot D\gz\geq 1 \ \ \ \text{ on }\ \pl\gO, \end{equation} and that any classical subsolution (resp., supersolution) of \begin{equation}\label{eq-loc} \gl u+F_\phi[u]=0 \ \ \text{ in }\ V\cap \gO \quad\text{ and }\quad \gamma\cdot Du=\psi \ \ \ \text{ on }\ V\cap\pl\gO, \end{equation} where $V$ is an open subset of $\R^n$, is also a viscosity subsolution (resp., supersolution) of \eqref{eq-loc}. Note that \eqref{CPN'} implies \eqref{CPN}. To see this, one may select $V$ to be $\R^n$ in \eqref{CPN'}. As in the boundary conditions treated above, by a rescaling argument, one sees that the condition \eqref{CPN} (reps., \eqref{CPN'}), only with those $\phi=L+\chi$, where $\chi\in C(\tt)$ is arbitrary, implies the full condition \eqref{CPN} (resp., \eqref{CPN'}). \begin{prop} \label{prop1-n} Assume \eqref{F1}, \eqref{F2}, \eqref{OG}, \eqref{CPN}, \eqref{SLN} and \eqref{EC}. Then there exists a solution $(u,c)\in C(\tt)\times\R$ of \eqref{EN}, and such a constant $c\in \R$ is unique. Moreover, if $v^\gl\in C(\tt)$ is a solution of \eqref{N}, then \begin{equation}\label{c-n} c=-\lim_{\gl \to 0} \gl v^\gl(x) \quad \text{ uniformly on }\ \tt. \end{equation} \end{prop} We denote by $c_{\rN}$ the constant given by the proposition above and call it the critical value of \eqref{EN}. \begin{proof}[Outline of proof] As noted above, there exists a function $\gz\in C^2(\tt)$ satisfying \eqref{gz}. We may as well assume that $\gz\geq 0$ on $\tt$. Choose two positive constants $M_1$ and $M_2$ so that $|g|\leq M_1$ on $\bry$ and $|F[\pm M_1\gz]|\leq M_2$ on $\tt$, and observe that, for any $\gl>0$, the functions $M_1\gz+\gl^{-1}M_2$ and $-M_1\gz-\gl^{-1}M_2$ are a supersolution and a subsolution of \eqref{N}. By \eqref{CPN}, we get $\,|v^\gl|\leq M_1\gz+\gl^{-1}M_2$ on $\tt$. This shows that $\{\gl v^\gl\}_{\gl>0}$ is uniformly bounded on $\tt$. Using \eqref{EC}, we find that the family $\{v^\gl-m^\gl\}_{\gl>0}$, where $m^\gl:=\min_{\tt}v^\gl$, is relatively compact in $C(\tt)$, while, for each $\gl>0$, the function $u:=v^\gl-m^\gl$ is a solution of $\gl u+\gl m^\gl +F[u]=0$ in $\gO$ and $\gamma\cdot Du=g$ on $\bry$. Sending $\gl\to 0$, along an appropriate sequence, yields a solution $(v,c)\in C(\tt)\tim \R$ of \eqref{EN}. The uniqueness of the constant $c$ is a consequence of \eqref{CPN}. \end{proof} \subsection{Representation formulas} Let $(z,\gl)\in\tt\tim[0,\,\infty)$. We define the sets $\cF^{\rN}(\gl)\subset C(\tt\tim\cA)\times C(\tt)$, $\cG^{\rN}(z,\gl)\subset C(\bb)\tim C(\bry)$, respectively, by \begin{align*} &\cF^{\rN}(\gl):=\left\{(\phi,\psi,u)\in \Psi^+\tim C(\tt) \mid u \ \text{is a subsolution of} \ \eqref{N'}\right\}, \\ &\cG^{\rN}(z,\gl):=\left\{(\phi-\gl u(z), \psi)\mid (\phi,\psi,u)\in\cF^{\rN}(\gl)\right\}. \end{align*} \begin{lem} \label{thm2-sec3+0} Assume \eqref{F1}, \eqref{F2}, \eqref{OG} and \eqref{CPN'}. Let $(z,\gl)\in\tt\tim[0,\,\infty)$. The set $\cG^{\rN}(z,\gl)$ is a convex cone in $C(\bb)\tim C(\bry)$ with vertex at the origin. \end{lem} The proof of this is in the same line as that of \cite[Lemma 2.8]{IsMtTr1} with help of the following lemma. Thus, we omit presenting it here. \begin{lem}\label{convexN} Assume \eqref{F1}, \eqref{F2}, \eqref{OG} and \eqref{CPN'}. Let $\gl\in[0,\,\infty)$ and let $(\phi_i,\psi_i, u_i)\in\cF^{\rN}(\gl)$, with $i=1,2$. For $t \in (0,1)$, set $\phi^t=t\phi_1+(1-t)\phi_2$ on $\bb$, $\psi^t=t\psi_1+(1-t)\psi_2$ on $\bry$ and $u^t=t u_1+(1-t)u_2$ on $\tt$. Then, $u^t\in\cF^\rN(\gl)$. \end{lem} \begin{proof} Note first that, given $(\phi,\psi,u)\in\Psi^+\tim C(\tt)$, $(\phi,\psi,u)\in\cF^\rN(\gl)$ if and only if $(\phi-\gl u,\psi,u)\in\cF^\rN(0)$. It is then easily seen that our claim follows from the special case $\gl=0$. Thus we may assume henceforth that $\gl=0$. Let $\eta\in C^2(\tt)$ and $z\in\tt$ be such that $u^t-\eta$ takes a strict maximum at $z$. If $z\in\gO$, then the proof of \cite[Lemma 2.8]{IsMtTr1} ensures that $F_{\phi_t}[\eta](z)\leq 0$. Thus, we only need to show that, if $z\in\pl\gO$, then we have either $F_{\phi^t}[\eta](z)\leq 0\,$ or $\,\gamma(z)\cdot D\eta(z)\leq\psi^t(z)$. To do this, we assume that $z\in\pl\gO$, suppose to the contrary that $F_{\phi^t}[\eta](z)>0$ and $\,\gamma(z)\cdot D\eta(z)>\psi^t(z)$, and obtain a contradiction. We choose $\ep>0$ and an open neighborhood $V$, in $\R^n$, of $z$ so that \begin{equation}\label{super-eta} F_{\phi^t}[\eta](x)>\ep \ \ \text{ in } V\cap\lbar\gO \ \ \text{ and } \ \ \gamma(x)\cdot D\eta(x)\geq \psi^t(x) \ \ \text{ for all }x\in V\cap\pl\gO. \end{equation} Set \[ w(x)=-t^{-1}(1-t)u_2(x)+t^{-1}\eta(x) \ \ \text{ for }x\in\tt, \] and prove that $w$ is a supersolution of \begin{equation}\label{superVN} \begin{cases} F_{\phi_1}[w]=\ep \ \ \ \ \text{ in }V\cap \gO, &\\[3pt] \gamma\cdot Dw=\psi_1 \ \ \text{ on }V\cap\pl\gO. & \end{cases}\end{equation} Once this is done, we apply the comparison principle \eqref{CPN'} to $u_1$ and $w$, to obtain \[ \sup_{V\cap\lbar\gO}(u_1-w)\leq \sup_{\gO\cap\pl V}(u_1-w). \] This gives a contradiction since $t(u_1-w)=u^t-\eta$ attains a strict maximum at $z\in V$ on $V\cap\lbar\gO$. To prove the viscosity property \eqref{superVN} of $w$, we fix $\xi\in C^2(V\cap\lbar\gO)$ and $y\in V\cap\lbar\gO$, and assume that $w-\xi$ takes a minimum at $y$. An immediate consequence of this is that the function $t(1-t)^{-1}(w-\xi)$ has a minimum at $y$ and, thus, the function \[ u_2-(1-t)^{-1}\eta+t(1-t)^{-1}\xi \] has a maximum at $y$. By the viscosity property of $u_2$, we get either \begin{equation}\label{eta-xi1} F_{\phi_2}[(1-t)^{-1}\eta-t(1-t)^{-1}\xi](y)\leq 0, \end{equation} or \begin{equation}\label{eta-xi2} y\in\pl\gO \ \ \text{ and }\ \ \gamma(y)\cdot D((1-t)^{-1}\eta-t(1-t)^{-1}\xi)(y)\leq \psi_2(y). \end{equation} If \eqref{eta-xi1} holds, then, using \eqref{super-eta}, we get \[ \ep\leq F_{\phi^t}[\eta](y) \leq tF_{\phi_1}[\xi](y)+(1-t)F_{\phi_2}[(1-t)^{-1}\eta-t(1-t)^{-1}\xi](y) \leq tF_{\phi_1}[\xi](y). \] On the other hand, if \eqref{eta-xi2} holds, then, using \eqref{super-eta}, we get \[\begin{aligned} \gamma(y)\cdot D\xi(y) &\geq t^{-1}\gamma(y)\cdot D\eta(y)-t^{-1}(1-t)\psi_2(y) \\&\geq t^{-1}\psi_t(y)-t^{-1}(1-t)\psi_2(y)=\psi_1(y). \end{aligned}\] These show, with help of \eqref{OG}, that $w$ is a (viscosity) supersolution of \eqref{superVN}, which completes the proof. \end{proof} \begin{lem}\label{lem-n-3} Assume \eqref{F1}, \eqref{F2}, \eqref{OG} and \eqref{CPN}. Let $(\phi,\psi,u)\in\cF^\rN(\gl)$, with $\phi=tL+\chi$ for some $t>0$ and $\chi\in C(\tt)$. Then, there exists a constant $C>0$, depending only on $\gO$ and $F$, such that \[ \gl u\leq \|\chi\|_{C(\tt)}+(1+\gl)C\|\psi\|_{C(\bry)} \ \ \ \text{ on }\ \tt. \] \end{lem} \begin{proof} Let $\gz\in C^2(\tt)$ be a function that satisfies \[ \gamma\cdot D\gz\geq 1 \ \ \text{ on }\ \bry \ \ \ \text{ and } \ \ \gz\geq 0 \ \ \text{ on } \tt. \] As before, we observe that $v:=t^{-1}u$ is a subsolution of \begin{equation}\label{lem-n-3-1} \gl v+F[v]=t^{-1}\chi \ \ \text{ in }\gO \ \ \ \text{ and } \ \ \ \gamma\cdot Dv=t^{-1}\psi \ \ \text{ on }\ \bry. \end{equation} We set \[ C_1:=\max_{\tt}\gz \ \ \ \text{ and } \ \ \ C_2:=\max_{(x,t)\in\tt\tim[0,\,1]}|F[t\gz](x)|. \] We set $w:=A\gz+B$ for constants $A\geq 0$ and $B\geq 0$, to be fixed in a moment, and note that \[ \gl w+F[w]\geq \gl B+F[A\gz] \ \ \text{ in }\ \gO, \] and \[ \gamma\cdot Dw=A\gamma\cdot D\gz\geq A \ \ \text{ on }\ \bry. \] Now, observe that if $A>1$, then the convexity of $F$ yields \[ -C_2\leq F[\gz]\leq A^{-1}F[A\gz]+(1-A^{-1})F[0] \leq A^{-1}F[A\gz]+C_2, \] and, hence, \[ F[A\gz]\geq -2AC_2 \ \ \text{ on }\ \tt, \] which is obviously true also in the case when $0<A\leq 1$. Thus, putting \[ A:=t^{-1}\|\psi\|_{C(\bry)}\quad \text{ and } \quad B:=\gl^{-1}(t^{-1}\|\chi\|_{C(\tt)}+2AC_2), \] we see that $w$ is a supersolution of \eqref{lem-n-3-1}. Then, \eqref{CPN} implies that $v\leq w$ on $\tt$, which reads \[ \gl u\leq \gl t w\leq \gl t(AC_1+B)=\|\chi\|_{C(\tt)} +(\gl C_1+2 C_2)\|\psi\|_{C(\bry)}, \] which completes the proof. \end{proof} We set \[ \cP_1^\rN:=\cP_{\bb} \ \ \ \text{ and } \ \ \ \cP^\rN:=\cP_1^\rN\tim\cR_{\bry}^+ \] and, for compact subset $K$ of $\cA$, \[ \cP_{1,K}^\rN:=\left\{\mu_1\in\cP_{\bb} \mid\mu_1(\tt\times K)=1\right\}. \] Let $\cG^\rN(z,\gl)'$ denote the dual cone of $\cG^{\rN}(z,\gl)$ in $\cR_L\tim\cR_{\bry}$, that is, \[ \cG^\rN(z,\gl)':=\left\{(\mu_1,\mu_2)\in \cR_{L}\tim\cR_{\bry}\mid \lan\mu_1,\phi\ran +\lan\mu_2,\psi\ran\ge0, \ \ \text{for all } (\phi,\psi)\in\cG^{\rN}(z,\gl)\right\}. \] We use as well the notation: $\cF^\rN(0)=\cF^\rN(z,0)$, $\cG^\rN(0)=\cG^\rN(z,0)$, and $\cG^\rN(0)'=\cG^\rN(z,0)'$. \begin{thm} \label{thm1-n} Assume \eqref{F1}, \eqref{F2}, \eqref{L}, \eqref{OG} and \eqref{CPN'}. \ \emph{(i)}\ Let $(z,\gl)\in\tt\tim(0,\,\infty)$. If $v^\gl\in C(\tt)$ is a solution of \eqref{N}, then \begin{equation} \label{n-min} \gl v^\gl(z)=\min_{(\mu_1,\mu_2)\in \cP^\rN\cap \cG^{\rN}(z,\gl)'}\,\left( \lan \mu_1,\,L\ran+\lan\mu_2,\,g\ran\right). \end{equation} \emph{(ii)} Assume, in addition, \eqref{SLN} and \eqref{EC}. Then \begin{equation}\label{n-min0} -c_\rN=\min_{(\mu_1,\mu_2)\in \cP^\rN\cap \cG^{\rN}(0)'}\,\left( \lan \mu_1,\,L\ran+\lan\mu_2,\,g\ran\right). \end{equation} \end{thm} \begin{definition} We denote the set of minimizers of \eqref{n-min} and that of \eqref{n-min0} by $\cM^{\rN}(z,\gl)$ and $\cM^{\rN}(0)$, respectively. We call any $(\mu_1,\mu_2) \in \cM^{\rN}(0)$ a viscosity Mather measure associated with \eqref{EN}, and any $\gl^{-1}(\mu_1,\mu_2)$, with $(\mu_1,\mu_2) \in \cM^{\rN}(z,\gl)$ and $\gl>0$, a viscosity Green measure associated with \eqref{N}. \end{definition} For $M>0$, $(z,\gl)\in\tt\tim (0,\infty)$, we define $\cF^\rN(M,\gl)$, $\cG^\rN(M,z,\gl)$ and $\cG^\rN(M,z,\gl)'$ by \[ \begin{aligned} \cF^\rN(M,\gl):=&\,\cF^\rN(\gl)\cap(\Psi^+(M)\tim C(\tt)),& \\\cG^\rN(M,z,\gl):=&\,\{(\phi-\gl u(z),\psi)\mid (\phi,\psi,u)\in\cF^\rN(M,\gl)\},& \\\cG^\rN(M,z,\gl)':=&\,\{(\mu_1,\mu_2)\in\cR_L\tim\cR_{\bry} \mid \lan\mu_1,f\ran+\lan\mu_1,\psi\ran\geq 0& \\&& \kern-100pt \text{ for all }\,(f,\psi)\in\cG^\rN(M,z,\gl)\}. \end{aligned} \] It is easily seen by Lemma \ref{thm2-sec3+0} that $\cG^\rN(M,z,\gl)$ is a convex cone in $C(\bb)\tim C(\bry)$ with vertex at the origin. \begin{thm} \label{thm1-n1} Assume \eqref{F1}, \eqref{F2}, \eqref{OG}, \eqref{L}, and \eqref{CPN'}. Let $z\in\tt$ and $\gl,\, M\in(0,\,\infty)$. If $v^\gl\in C(\tt)$ is a solution of \eqref{N}, then \begin{equation} \label{n-inf} \gl v^\gl(z)\geq \inf_{(\mu_1,\mu_2)\in \cP^\rN\cap\cG^{\rN}(M,z,\gl)'} \,\left(\lan \mu_1,\,L\ran+\lan\mu_2,\,g\ran\right). \end{equation} \end{thm} \begin{proof} Let $v^\gl\in C(\tt)$ be a solution of \eqref{N}. We fix any $\ep\in(0,\,1)$, and show that there exist $R>0$ and a compact subset $K$ of $\cA$ such that \begin{equation} \label{n-inf-1} \gl v^\gl(z)+\ep\geq \inf_{(\mu_1,\mu_2)\in \cP^\rN_{K,R}\cap \cG^\rN(M,z,\gl)'}(\lan\mu_1,L\ran+\lan\mu_2,g\ran), \end{equation} where \[ \cP^\rN_{K,R} :=\left\{(\mu_1,\mu_2)\in\cP^\rN \mid \mu_1(\tt\tim K)=1,\, \mu_2(\bry)\leq R\right\}. \] Since $\cP^\rN_{K,R}\subset \cP^\rN$, it follows from \eqref{n-inf-1} that \[ \gl v^\gl(z)+ \ep\geq \inf_{(\mu_1,\mu_2) \in \cP^\rN\cap\cG^\rN(M,z,\gl)'}(\lan\mu_1,L\ran+\lan\mu_2,g\ran), \] which implies that \eqref{n-inf} is valid. We choose a constant $A>0$ so that $g+A\geq 1$ on $\bry$ and, thanks to \eqref{OG} (see also \eqref{gz}), a function $\eta\in C^2(\tt)$ so that \begin{equation}\label{n-inf-2} \gamma\cdot D\eta\leq -A \ \ \text{ on }\ \bry \quad\text{ and } \quad\eta(z)=0. \end{equation} Then, we choose constants $B>0$ and $R>0$ so that \begin{equation}\label{n-inf-3} \gl\eta+F[\eta]\leq B \ \ \ \text{ on }\tt \quad\text{ and }\quad R\geq 1+B+\gl v^\gl(z). \end{equation} Let $K$ be a compact subset of $\cA$ to be specified later, and set \[ I_{K,R}:=\inf_{(\mu_1,\mu_2)\in\cP^\rN_{K,R}\cap\cG^\rN(M,z,\gl)'}(\lan\mu_1,L\ran+\lan\mu_2,g\ran) \] Since $\cG^{\rN}(M,z,\gl)$ is a convex cone with vertex at the origin, we deduce that \[ \inf_{(f,\psi)\in\cG^{\rN}(M,z,\gl)} \left(\lan \mu_1,\,f\ran+\lan\mu_2,\psi\ran\right)= \begin{cases} 0 \ \ &\text{ if }\ (\mu_1,\mu_2)\in\cP^\rN_{K,R} \cap\cG^{\rN}(M,z,\gl)', \\ -\infty &\text{ if }\ (\mu_1,\mu_2)\in \cP^\rN_{K,R}\setminus \cG^{\rN}(M,z,\gl)'. \end{cases} \] and, furthermore, \begin{equation}\begin{aligned} \label{thm1-n-2+} I_{K,R} =\inf_{(\mu_1,\mu_2)\in\cP^{\rN}_{K,R}}\ \sup_{(f,\psi)\in\cG^{\rN}(M,z,\gl)}\,\left(\lan \mu_1, L-f\ran+\lan\mu_2,g-\psi\ran\right). \end{aligned} \end{equation} Note that $\cP^\rN_{K,R}$ is a compact, convex subset of $\cR_{L}\tim\cR_2^{+}$. Sion's minimax theorem implies that \begin{equation}\label{n-inf-4}\begin{aligned} I_{K,R}&\,=\min_{(\mu_1,\mu_2)\in\cP^\rN_{K,R}} \ \sup_{(f,\psi)\in\cG^{\rN}(M,z,\gl)}\,\left(\lan \mu_1, L-f\ran+\lan\mu_2,g-\psi\ran\right) \\&= \sup_{(f,\psi)\in\cG^{\rN}(M,z,\gl)}\ \min_{(\mu_1,\mu_2)\in\cP^\rN_{K,R}} \,\left(\lan \mu_1, L-f\ran+\lan\mu_2,g-\psi\ran\right). \end{aligned} \end{equation} In order to prove \eqref{n-inf-1} for the fixed $R>0$ and a suitably chosen compact set $K\subset\cA$, we argue by contradiction. We suppose that \[ \gl v^\gl(z)+\ep< I_{K,R}, \] which, together with \eqref{n-inf-4}, yields \begin{equation}\label{thm1-n-4} \gl v^\gl(z)+\ep < \ \sup_{(f,\psi)\in\cG^{\rN}(M,z,\gl)}\ \min_{(\mu_1,\mu_2)\in\cP^\rN_{K,R}} \,\left(\lan \mu_1, L-f\ran+\lan\mu_2,g-\psi\ran\right). \end{equation} Hence, we may choose $(\phi,\psi,u)\in\cF^{\rN}(M,\gl)$ and $(t,\chi)\in (0,\,\infty)\tim C(\tt)$ so that $\phi=tL+\chi$, $\|\chi\|_{C(\tt)}< tM$, $\|\psi\|_{C(\bry)}< tM$, and \[ \gl v^\gl(z)+\ep < \inf_{(\mu_1,\mu_2)\in\cP^\rN_{K,R}} \,\left(\lan \mu_1, L-\phi+\gl u(z)\ran+\lan\mu_2,g-\psi\ran\right). \] We get from the above \begin{equation}\label{thm1-n-4++} \gl v^\gl(z)+\ep < \inf_{\mu\in\cP^\rN_{1,K}} \,\lan \mu, L-\phi+\gl u(z)\ran, \end{equation} and, since $(0,R\gd_{x})\in \cP^{\rN}_{K,R}$ for any $x\in\bry$, \[ \gl v^\gl(z)+\ep < \inf_{\mu\in\cP^\rN_{1,K}} \,\lan \mu, L-\phi+\gl u(z)\ran+R\min_{\bry}(g-\psi). \] Thus, setting \[ p:=\inf_{\mu\in\cP^\rN_{1,K}}\lan\mu, L-\phi+\gl (u-v^\gl)(z)\ran \ \ \text{ and } \ \ q:=\min_{\bry}(g-\psi), \] we have \begin{equation} \label{thm1-n-4+} \ep<p \ \ \ \text{ and } \ \ \ \ep<p+Rq. \end{equation} Our choice of $A$, $B$ and $\eta$ ensures that $(L+B,-A,\eta)\in\cF^\rN(\gl)$. Note as well that $(L,g,v^\gl),\ (\phi,\psi,u)\in\cF^{\rN}(\gl)$. Set \[ \nu:=\fr{R\ep}{R\ep+p}\in (0,\,1), \] and observe by the convexity of $\cF^{\rN}(\gl)$ that \[ (L+B\ep,(1-\ep)g-A\ep,(1-\ep)v^\gl+\ep\eta)\in\cF^{\rN}(\gl), \] and \[ (1-\nu)(L+ B\ep,(1-\ep)g-A\ep,(1-\ep)v^\gl+\ep\eta)+\nu(\phi,\psi,u)\in\cF^{\rN}(\gl). \] We set \[\begin{gathered} (\hat\phi,\hat \psi,\hat u):= (1-\nu)(L+ B\ep,(1-\ep)g-A\ep,(1-\ep)v^\gl+\ep\eta) +\nu(\phi,\psi,u), \\ \hat t:=1+\nu(t-1) \ \ \ \text{ and } \ \ \ \hat \chi:=(1-\nu) B\ep+\nu\chi, \end{gathered} \] and note that \[\begin{aligned} \hat\phi&\,=L+\nu(\phi-L)+(1-\nu)B\ep=\hat t L+\hat\chi, \\\hat\psi&\,=(1-\nu)(g-\ep(g+A))+\nu\psi, \\\hat u-v^\gl&\,=\nu(u-v^\gl)+(1-\nu)\ep(\eta-v^\gl), \\\hat t-1&\,=\nu(t-1). \end{aligned} \] Using the facts that $g+A\geq 1$ on $\bry$ and that, by the definition of $q$, $\psi\leq g-q$ on $\bry$, and the second inequality of \eqref{thm1-n-4+}, we compute \[\begin{aligned} \hat\psi&\,\leq (1-\nu)(g-\ep)+\nu(g-q) \\&\,\leq (1-\nu)(g-\ep)+\nu\left(g+\fr{p-\ep}{R}\right) =g-\fr{\nu\ep}{R} \ \ \text{ on }\bry. \end{aligned} \] Also, we compute \[\begin{aligned} L-\hat\phi+&\gl(\hat u-v^\gl)(z) \\ \,=\, &\nu(L-\phi+\gl (u-v^\gl)(z)) -(1-\nu)\ep(B+v^\gl(z)) \ \ \text{ in }\bb, \end{aligned} \] and, by using \eqref{thm1-n-4+} and the second inequality of \eqref{n-inf-3}, that for any $\mu\in\cP^\rN_{1,K}$, \[\begin{aligned} \lan\mu&,L-\hat\phi+\gl(\hat u-v^\gl)(z)\ran \\&\,\geq \nu p-(1-\nu)\ep(B+\gl v^\gl(z)) \\&\,= (1-\nu)R\ep-(1-\nu)\ep(B+\gl v^\gl(z)) \geq (1-\nu)\ep. \end{aligned} \] Thus, setting $\hat\ep:=\ep\min\{R^{-1}\nu,\, 1-\nu\}$, we find that \begin{equation}\label{thm1-n-5} \hat \psi\leq g-\hat\ep \ \ \text{ on }\ \bry, \end{equation} and \begin{equation}\label{thm1-n-5+} \inf_{\mu\in\cP^\rN_{1,K}}\lan\mu, L-\hat\phi +\gl(\hat u-v^\gl)(z)\ran>\hat\ep. \end{equation} We now use these estimates to show that there is positive constant $\gth$ such that $w:=\gth \hat u$ is a subsolution of \begin{equation}\label{thm1-n-6} \begin{cases} \gl w+F[w]=-\gl (v^\gl-w)(z)-\gth\hat\ep\ \ \text{ in }\gO&\\[3pt] \gamma\cdot Dw\leq g \ \ \ \text{ on }\bry. \end{cases} \end{equation} After this is done, we easily get a contradiction as follows: observe that the function $\xi:=w+ (v^\gl-w)(z)+\gl^{-1}\gth\hat\ep$ is a subsolution of $\gl \xi+F[\xi]=0$ in $\gO$ and $\gamma \cdot D\xi=g$ on $\bry$, and, by comparison principle \eqref{CPN}, that $\xi\leq v^\gl$ on $\tt$, which, evaluated at $z$, gives $\,\gl^{-1}\gth \hat \ep\leq 0$. We thus get a contradiction. To show \eqref{thm1-n-6}, we treat first the case when $\cA$ is compact. Select $K=\cA$ and note that $\gd_{(x,\ga)}\in\cP^\rN_{1,K}$ for all $(x,\ga)\in\bb$. Thus, from \eqref{thm1-n-5+}, we get \[ \hat\phi<L-\gl (v^\gl-\hat u)(z)-\hat\ep \ \ \text{ on }\bb, \] and we see that $\hat u$ is a subsolution of \eqref{thm1-n-6}, with $\gth=1$. Consider next the case where $\cA$ is not compact. Choose a point $\ga_0\in\cA$ and a constant $L_0>0$ so that \begin{equation}\label{thm1-n-7} \max_{x\in\tt}L(x,\ga_0)\leq L_0 \ \ \ \text {and } \ \ \ \gl|v^\gl(z)|\leq L_0. \end{equation} Let $L_1>0$ be a constant to be fixed later, and, in view of \eqref{L}, we select a compact set $K_0\subset\cA$ so that \begin{equation}\label{thm1-n-8} L(x,\ga)\geq \max\{L_0,L_1\} \ \ \ \text{ for all }\ (x,\ga)\in\tt\tim(\cA\setminus K_0). \end{equation} Pick a point $\ga_1\in\cA\setminus K_0$ and set $K=K_0\cup\{\ga_1\}$. Since $\gd_{(x,\ga)}\in\tt\tim \cP_{1,K}$ for all $(x,\ga)\in\tt\tim K$, by \eqref{thm1-n-5+}, we get \begin{equation}\label{thm1-n-9} \gl (v^\gl-\hat u)(z)+\hat\ep<(1-\hat t)L(x,\ga)-\hat\chi(x) \ \ \text{ for all }\ (x,\ga)\in\tt\tim K. \end{equation} We divide the argument into two cases. Consider first the case when $\hat t\leq 1$. We repeat the same lines as in the proof of Theorem \ref{thm1-sc}, to deduce that \[ (\hat t-1)L+\hat \chi< L-\gl (v^\gl-\hat u)(z)-\hat \ep \ \text{ on }\bb, \] then that \[ \hat\phi=L+(\hat t-1)L+\hat \chi< -\gl (v^\gl-\hat u)(z)-\hat \ep \ \text{ on }\bb, \] and, hence, that $\hat u$ is a subsolution of \eqref{thm1-n-6}, with $\gth=1$. Secondly, we consider the case when $\hat t\geq 1$. Again, repeating the same lines as in the proof of Theorem \ref{thm1-sc}, we obtain \[ \hat\phi<\hat tL-\hat t\gl v^\gl(z)+\gl \hat u(z)-\hat\ep \ \ \text{ in }\bb, \] which ensures that $w:={\hat t}^{-1}\hat u$ is a subsolution of \begin{equation}\label{thm1-n-10}\left\{\begin{aligned} &\gl w+F[w]=\gl(u-v^\gl)(z)-{\hat t}^{-1}\hat\ep\ \ \text{ in }\ \gO, \\&\gamma\cdot Dw={\hat t}^{-1}(g-\hat \ep) \ \ \text{ on }\ \bry. \end{aligned}\right.\end{equation} Our next step is to show that there exists $L_1$ for which $\hat t^{-1}(g-\hat \ep)\leq g$ on $\bry$. For this, we give an upper bound of $p$, independent of the choice of $L_1$ (and, hence, $K$). Because of \eqref{L}, we may take a positive constant $C_0$ so that $L\geq -C_0$ on $\bb$. Recall that $\nu(t-1)=\hat t-1>0$, and observe that \[ L-\phi =(1-t)L-\chi \leq (t-1)C_0+tM \ \ \text{ on }\ \bb, \] and, moreover, \[ \lan\mu, L-\phi\ran\leq (t-1)C_0+tM \ \ \text{ for all }\ \mu\in\cP^\rN_{1,K}. \] According to Lemma \ref{lem-n-3}, there exists a constant $C_1>0$, depending only on $\gO$, $F$ and $\gl$, such that \[ \gl u\leq C_1Mt \ \ \ \text{on }\ \tt. \] Thus, by the definition of $p$, we get \begin{equation}\label{est-p} p\leq (t-1)C_0+(C_1+1)Mt+\gl |v^\gl(z)|. \end{equation} Next, by \eqref{thm1-n-4++}, we get \[ (\phi-L)(x,\ga_1)\leq\gl(u-v^\gl)(z) \ \ \ \text{ for all }\ x\in\tt, \] and moreover, \[ (t-1)L_1\leq -\chi(x)+\gl(u-v^\gl)(z) \leq (C_1+1)Mt+|\gl v^\gl(z)| \ \ \ \text{ for all }\ x\in\tt. \] Hence, if $L_1>(C_1+1)M$, then we get \begin{equation}\label{est-t} t\leq \fr{L_1+\gl|v^\gl(z)|}{L_1-(C_1+1)M}. \end{equation} Now, set \[ p_0:=C_0+2(C_1+1)M+\gl |v^\gl(z)|,\quad\ep_0:=\min\Big\{\fr{\ep^2}{R\ep+p_0},\,\fr{\ep}{R+1}\Big\}, \] and \[ \tau:=1+\fr{\ep_0}{\|g\|_{C(\bry)}+1}, \] and note that $p_0>0$, $0<\ep_0<1$ and $1<\tau<2$ and that these constants $p_0,\,\ep_0$ and $\tau$ are independent of the choice of $L_1$ and $K$. Then fix $L_1>(C_1+1)M$ large enough so that \[ \fr{L_1+\gl|v^\gl(z)|}{L_1-(C_1+1)M}\leq \tau. \] Observe by \eqref{est-t} that $t\leq \tau$, which implies together with \eqref{est-p} and \eqref{thm1-n-4+} that $\ep<p\leq p_0$. This implies moreover that \[ \fr{R\ep}{R\ep+p_0}\leq \nu<\fr{R}{R+1}, \] and thus, \[ \hat\ep\geq \ep\min\Big\{\fr{\ep}{R\ep+p_0},\fr{1}{R+1}\Big\}=\ep_0. \] Hence, \[\begin{aligned} \hat\ep&\,\geq \ep_0=(\tau-1)\left(\|g\|_{C(\bry)}+1\right) > (t-1)\|g\|_{C(\bry)} \\&\,\geq (\hat t-1)\|g\|_{C(\bry)} \geq -(\hat t-1)g \ \ \text{ on }\ \bry, \end{aligned}\] and therefore, \[ {\hat t}^{-1}(g-\hat\ep)<g \ \ \ \text{ on }\ \bry. \] We now conclude from this and \eqref{thm1-n-10} that $w:=\hat{t}^{-1}\hat u$ is a subsolution of \eqref{thm1-n-6}, with $\gth=\hat{t}^{-1}$, which completes the proof. \end{proof} \begin{lem} \label{est-mu} Assume \eqref{F1}, \eqref{F2}, \eqref{L} and \eqref{OG}. Let $(z,\gl,M)\in\tt\tim(0,\,\infty)\tim(0,\,\infty)$ and $(\mu_1,\mu_2)\in\cP^\rN\cap\cG^\rN(M,z,\gl)'$. For each $\ep>0$ there exists a constant $C>0$, depending only on $\ep$, $\gO$ and $F$, such that if $M>C$, then \[ \mu_2(\bry)\leq \ep \lan\mu_1,L\ran+C(1+\gl). \] \end{lem} \begin{proof} According to \eqref{OG} or \eqref{gz}, there exists a function $\eta\in C^2(\tt)$ such that \[ \gamma\cdot D\eta\leq -1 \ \ \text{ on }\ \bry \quad\text{ and }\quad \eta\leq 0\ \ \text{ on }\ \tt. \] Fix any $\ep>0$, set \[ C_1:=\|F[\ep^{-1}\eta]\|_{C(\tt)}+\ep^{-1}, \] and observe that if $M>C_1$, then \[ (L+C_1,-\ep^{-1},\ep^{-1}\eta)\in\cF^\rN(M,z,\gl). \] Assume that $M>C_1$. Since $(\mu_1,\mu_2)\in\cP^\rN\cap\cG^\rN(M,z,\gl)'$, we get \[ 0\leq \lan\mu_1,L+C_1-\gl \ep^{-1}\eta(z)\ran +\lan\mu_2,-\ep^{-1}\ran, \] which yields \[ \mu_2(\bry)\leq \ep \lan\mu_1,L\ran +\ep C_1+\gl\|\eta\|_{C(\bry)}. \] If we set $C=(1+\ep)C_1+\gl\|\eta\|_{C(\bry)}$ and if $M>C$, then we have \[ \mu_2(\bry)\leq \ep \lan\mu_1,L\ran +C(1+\gl). \] The proof is now complete. \end{proof} \begin{proof}[Proof of Theorem \ref{thm1-n}] We here prove only assertion (ii), since the proof of (i) is similar and slightly simpler. Fix a point $z\in\tt$. For each $\gl>0$, let $v^\gl\in C(\tt)$ be a solution of \eqref{N}. Recall that \begin{equation}\label{thm1-np-1} -c_\rN=\lim_{\gl\to 0+}\gl v^\gl(z). \end{equation} By Lemma \ref{est-mu}, there exists a constant $C_1>0$, depending only on $\|g\|_{C(\bry)}$, $\gO$ and $F$, such that for any $(M,\gl)\in(0,\,\infty)^2$ and $(\mu_1,\mu_2)\in\cP^\rN\cap\cG^\rN(M,z,\gl)'$, if $M\geq C_1$, then \begin{equation}\label{thm1-np-2} \mu_2(\bry)\leq \fr{1}{2\|g\|_{C(\bry)}+1}\lan\mu_1,L\ran+C_1(1+\gl) \end{equation} Fix a sequence $\{\gl_j\}_{j\in\N}\subset(0,\,1)$ converging to zero. Thanks to Theorem \ref{thm1-n1}, for each $j\in\N$, there exists $(\mu_1^j,\mu_2^j)\in\cP^\rN\cap \cG^\rN(j,z,\gl_j)'$ such that \begin{equation}\label{thm1-np-3} \gl_j v^{\gl_j}(z)+\fr{1}{j}\geq \lan\mu_1^j,L\ran+\lan\mu_2^j,g\ran. \end{equation} From this, we get \[ \lan\mu_1^j, L\ran \leq \gl_j v^{\gl_j}(z)+1+\|g\|_{C(\bry)}\mu_2(\bry). \] Combine this with \eqref{thm1-np-2}, to obtain \[ \lan\mu_1^j, L\ran\leq 2\left(\gl_j v^{\gl_j}(z)+1+ \|g\|_{C(\bry)}C_1(1+\gl_j)\right)\ \ \ \text{ if }\ j>C_1, \] and to find that the sequences \[ \{\lan\mu_1^j,L\ran\}_{j\in\N} \quad \text{ and }\quad\{\mu_2^j(\bry)\}_{j\in\N} \] are bounded from above and, hence, bounded. The boundedness of the two sequences above implies, with help of Lemma \ref{basic-cpt}, that the sequence $\{(\mu_1^j,\mu_2^j)\}_{j\in\N}$ of measures has a subsequence, convergent in the topology of weak convergence, which we denote by the same symbol. Let $(\mu_1,\mu_2)\in\cP^\rN$ be the limit of the sequence $\{(\mu_1^j,\mu_2^j)\}_{j\in\N}$. By Lemma \ref{basic-lsc}, we have \[ \lan\mu_1, L\ran\leq \liminf_{j\to\infty}\lan\mu_1^j,L\ran. \] By passing again to a subsequence if necessary, we may assume that the sequence $\{\lan\mu_1^j,L\ran\}_{j\in\N}$ is convergent. We set \[ \rho:=\lim_{j\to\infty}\du{\mu_1^j,L}-\du{\mu,L}\,(\,\geq 0\,). \] Sending $j\to\infty$ in \eqref{thm1-np-3} yields together with \eqref{thm1-np-1} \begin{equation}\label{thm1-np-4} -c_\rN\geq \rho+\du{\mu_1,L}+\du{\mu_2,g}. \end{equation} Note that if $\cA$ is compact, then $\rho=0$. Next, we show that $(\mu_1,\mu_2)\in\cG^\rN(0)'$. Pick $(tL+\chi,\psi,u)\in\cF^\rN(0)$, where $t>0$ and $\chi\in C(\tt)$, and note that $(tL+\chi+\gl_j u,\psi,u)\in\cF^\rN(M,\gl_j)$ for all $j\in\N$ and some constant $M>0$. The dual property of $\cG^\rN(M,\gl_j)'$ yields \[ 0\leq\du{\mu_1^j,tL+\chi+\gl_ju-\gl_ju(z)}+\du{\mu_2^j,\psi} \ \ \ \text{ if }\ j>M. \] Sending $j\to\infty$ gives \begin{equation}\label{thm1-np-5} 0\leq t(\rho+\du{\mu_1,L})+\du{\mu_1,\chi}+\du{\mu_2,\psi}. \end{equation} Owing to Lemma \ref{mod}, we may choose $\tilde \mu_1 \in\cP_1^\rN$ such that \[ \du{\tilde\mu_1,L}=\rho+\du{\mu_1,L} \quad\text{ and }\quad \du{\tilde\mu_1,\gz}=\du{\mu_1,\gz} \ \ \text{ for all }\ \gz\in C(\tt). \] (If $\cA$ is compact, then we choose $\tilde\mu_1=\mu_1$.) This and \eqref{thm1-np-5} give \[ 0\leq \du{\tilde\mu_1,tL+\chi}+\du{\mu_2,\psi} \ \ \ \text{ for any }\ (tL+\chi,\psi,u)\in\cG^\rN(0), \] which ensures that $(\tilde\mu_1,\mu_2)\in \cP^\rN\cap\cG^\rN(0)'$. Also, applying the above to a solution $(u,c_\rN)$ of \eqref{EN}, which implies $(L+c_\rN,g,u)\in\cF^\rN(0)=\cG^\rN(0)$, yields \[ -c_{\rN}\leq\du{\tilde\mu_1,L}+\du{\mu_2,g}. \] Inequality \eqref{thm1-np-4} now reads \[ -c_\rN\geq \du{\tilde \mu_1,L}+\du{\mu_2,g}. \] Thus, we have \[ -c_\rN=\du{\tilde \mu_1,L}+\du{\mu_2,g} =\inf_{(\nu_1,\nu_2)\in\cP^\rN\cap\cG^\rN(0)'} (\du{\nu_1,L}+\du{\nu_2,g}), \] which finishes the proof. \end{proof} \subsection{Convergence with vanishing discount} As an application of Theorem \ref{thm1-n}, we prove here one of main results in this section. \begin{thm}\label{conv-n} Assume \eqref{F1}, \eqref{F2}, \eqref{OG}, \eqref{CPN'}, \eqref{SLN} and \eqref{EC}. For $\gl>0$, let $v^\gl$ be a solution of \eqref{N}. Then \[ \lim_{\gl\to 0+} (v^{\gl}+\gl^{-1}c_\rN)=u \ \ \ \text{ in }\ C(\tt) \] for some $u\in C(\tt)$, and $u$ is a solution of \emph{(EN$_{c_\rN}$)}. \end{thm} \begin{proof} Let $u_0\in C(\tt)$ be a solution of (EN$_{c_\rN}$). Observe that for each $\gl>0$, the functions $u_0+\|u_0\|_{C(\tt)}-\gl^{-1}c_\rN$ and $u_0-\|u_0\|_{C(\tt)}-\gl^{-1}c_\rN$ are a supersolution and a subsolution of \eqref{N}, respectively, and apply \eqref{CPN}, to get $\|v^\gl+\gl^{-1}c_\rN\|_{C(\tt)}\leq 2\|u_0\|_{C(\tt)}$ for all $\gl>0$, which, combined with \eqref{EC}, shows that the family $\{v^\gl+\gl^{-1}c_\rN\}_{\gl>0}$ is relatively compact in $C(\tt)$. Let $\cU$ denote the set of accumulation points of $\{v^\gl+\gl^{-1}c_\rN\}_{\gl>0}$ in $C(\tt)$ as $\gl\to 0$. The relative compactness of the family implies that $\cU\not=\emptyset$. Also, it is a standard observation that any $v\in\cU$ is a solution of (EN$_{c_\rN}$). Now, we prove that $\cU$ has a single element, which ensures that \eqref{conv-n} holds. Let $v,\, w\in\cU$, and select sequences $\{\gl_j\}_{j\in\N}$ and $\{\gd_j\}_{j\in\N}$ of positive numbers so that \[ \begin{cases}\disp \lim_{j\to\infty}\gl_j=\lim_{j\to\infty}\gd_j=0,&\\ \disp \lim_{j\to\infty}\left(v^{\gl_j}+\gl_j^{-1}c_\rN\right)=v \ \ \text{ in }\ C(\tt),&\\ \disp \lim_{j\to\infty}\left(v^{\gd_j}+\gd_j^{-1}c_\rN\right)=w \ \ \text{ in }\ C(\tt). \end{cases} \] Fix any $z\in\tt$, and, owing to Theorem \ref{thm1-n}, choose a sequence $\{(\mu_1^j,\mu_2^j)\}_{j\in\N}$ of measures so that, for every $j\in\N$, $(\mu_1^j,\mu_2^j)\in\cP^\rN\cap \cG^\rN(z,\gl_j)'$ and \begin{equation}\label{conv-np-1} \gl_j v^{\gl_j}(z)=\du{\mu_1^j,L}+\du{\mu_2^j,g}. \end{equation} As in the proof of Theorem \ref{thm1-n}, after passing to a subsequence of $\{\gl_j,\mu_1^j,\mu_2^j\}_{j\in\N}$ if necessary, we may assume that $\{(\mu_1^j,\mu_2^j)\}_{j\in\N}$ converges to some $(\mu_1,\mu_2)\in\cP^\rN$ and that for some $\tilde\mu_1\in\cP_1^\rN$, \begin{align} &\lim_{j\to\infty}\du{\mu_1^j,L}=\du{\tilde\mu_1,L},\notag \\&(\tilde\mu_1,\mu_2)\in\cG^\rN(0)',\notag \\&\du{\tilde \mu_1,\psi}=\du{\mu_1,\psi} \ \ \ \text{ for all }\ \psi\in C(\tt),\notag \\&-c_\rN=\du{\tilde\mu_1,L}+\du{\mu_2,g}.\label{conv-np-3} \end{align} Since $w$ is a solution of (EN$_{c_\rN}$), we have $(L+\gl_j w,g,w-\gl_j^{-1}c_\rN)\in\cF^\rN(\gl_j)$ for all $j\in\N$. Hence, by the dual cone property of $\cG^\rN(z,\gl_j)'$, we get \[ 0\leq \du{\mu_1^j,L+\gl_j w-\gl_j(w(z)-\gl_j^{-1}c_\rN)} +\du{\mu_2^j,g}. \] Combining this with \eqref{conv-np-1} yields \[ 0\leq \gl_j \left(v^{\gl_j}(z)+\gl_j^{-1}c_\rN -w(z)+\du{\mu_1^j,w}\right) \ \ \ \text{ for all }\ j\in\N, \] which implies in the limit $j\to\infty$ that \begin{equation}\label{conv-np-2} w(z)\leq v(z)+\du{\mu_1,w}=v(z)+\du{\tilde \mu_1,w}. \end{equation} Next, we note that $(L-\gd_j v^{\gd_j},g,v^{\gd_j}) \in\cF^\rN(0)$ and use the fact that $(\tilde\mu_1,\mu_2)\in\cG^\rN(0)'$ and then \eqref{conv-np-3}, to get \[ 0\leq \du{\tilde\mu_1,L-\gd_j v^{\gd_j}} +\du{\mu_2,g}=-\gd_j\du{\tilde\mu_1,v^{\gd_j}+\gd_j^{-1}c_\rN}, \] which yields, after division by $\gd_j$ and taking the limit $j\to\infty$, \[ \du{\tilde\mu_1,w}\leq 0. \] From this and \eqref{conv-np-2}, we get $\ w(z)\leq v(z)$, and, since $z\in\tt$ and $v,w\in\cU$ are arbitrary, we conclude that $\cU$ has only one element. \end{proof} \section{Examples} \label{sec-ex} In this section, we give some examples to which the main theorems in this paper, Theorems \ref{thm2-sc}, \ref{thm2-d}, \ref{conv-n}, can be applied. For the case of Theorem \ref{conv-n}, we always assume that \eqref{OG} holds. \subsection{First-order Hamilton-Jacobi equations} We consider the first-order Hamilton-Jacobi equations \eqref{DP} and \eqref{E}, where the function $F$ is replaced by $H=H(x,p)$ on $\tt\tim\R^n$. The Hamiltonian $H$ is assumed to be continuous and, moreover, satisfies two conditions: $H$ is coercive and convex, that is, \[ \lim_{|p|\to\infty}H(x,p)=+\infty \ \ \ \text{ uniformly for }\ x\in \tt, \] and \[ \text{ for every $x\in\tt$, the function $p\mapsto H(x,p)$ is convex in $\R^n$.} \] We assume as well that $\gO$ has a $C^1$-boundary. Thanks to \cite{Ishi, BaMi}, for every $\gl>0$, the solution $v^\gl$ of each of \eqref{S}, \eqref{D} and \eqref{N} exists and the family $\{v^\gl\mid \gl>0\}$ is equi-Lipschitz on $\tt$, which, in particular, implies that \eqref{EC} holds. Also, the comparison principles \eqref{CP}, \eqref{CPS}, \eqref{CPD} and \eqref{CPN'} hold. We take the advantage of the equi-Lipschitz property of $\{v^\gl\mid\gl>0\}$ and replace $H$ by another convex and coercive Hamiltonian $\widetilde H\in C(\tt\tim\R^n)$ that satisfies: (i) $\widetilde H(x,p)=H(x,p)$ for all $(x,p)\in \tt\tim B_{M_0}$, where $M_0>0$ is a Lipschitz bound for $\{v^\gl\mid\gl>0\}$ and $B_{M_0}$ denotes the ball of radius $M_0$ with center at the origin, and (ii) $\lim_{|p|\to\infty}\widetilde H(x,p)/|p|=+\infty$ uniformly on $\tt$. Let $L$ be the Lagrangian, corresponding to $\widetilde H$, given by \[ L(x,\ga)=\max_{p\in\R^n}(\ga\cdot p-\widetilde H(x,p)), \] which is a continuous function on $\tt\tim\R^n$. Moreover, we have \[ \widetilde H(x,p)=\max_{\ga\in\R^n}(p\cdot \ga-L(x,\ga)), \] and \[ \lim_{|\ga|\to\infty}L(x,\ga)=+\infty \ \ \ \text{ uniformly for }x\in\tt. \] Thus, in the present case, our choice of $\cA$, $a$ and $b$ are $\R^n$, $a(x,\ga)=0$ and $b(x,\ga)=-\ga$, respectively, and $L$ satisfies condition \eqref{L}. Furthermore, the solution $v^\gl$ of any of \eqref{S}, \eqref{D} or \eqref{N}, with $F=H$, is again a solution of the respective problem \eqref{S}, \eqref{D} or \eqref{N}, with $F$ replaced by $\widetilde H$. All the conditions and, hence, the convergence results of Theorems \ref{thm2-sc}, \ref{thm2-d}, and \ref{conv-n}, with $F=\widetilde H$, hold. This implies that the convergence assertions of Theorems \ref{thm2-sc}, \ref{thm2-d}, and \ref{conv-n}, with $F=H$, hold. \subsection{Fully nonlinear, possibly degenerate elliptic equations with superquadratic growth in the gradient variable} \label{ss-superquad} Firstly, set \[ F_0(x,p,X):=\max_{\ga \in \cA_0} \left( - \tr\gs^t(x,\ga)\gs(x,\ga) X - b(x,\ga)\cdot p - L_0(x,\ga) \right), \] where $\cA_0$ is a compact subset of $\R^l$ for some $l \in \N$, $\sigma \in \Lip(\tt \times \cA_0, \R^{k \times n})$ for some $k \in \N$, $b \in \Lip(\tt\times \cA_0,\R^n)$, and $L_0\in C(\tt \times \cA_0,\R)$. Consider the operator $F$ of the form \begin{align*} &F(x,p,X):= \frac{c(x)}{m}|p|^m + F_0(x,p,X)\\ =&\, \sup_{(q,\ga)\in\R^n\times \cA_0}\left(- \tr\gs^t(x,\ga)\gs(x,\ga) X+q\cdot p - b(x,\ga)\cdot p -\frac{c(x)^{\frac{-1}{m-1}}}{m'}|q|^{m'}- L_0(x,\ga)\right), \end{align*} where $m>2$ and $c \in \Lip(\tt,\R)$ with $c>0$. We set here $m':=m/(m-1)$. Assume further that $\partial \gO \in C^3$. It is worth pointing out that in this superquadratic case ($m>2$), the scale of the gradient term dominates the scale of the diffusion term. Because of this feature, for any subsolution $u$ of $\gl u +F[u] =0$ in $\gO$ with $\gl \geq 0$ such that $u$ is bounded in $\tt$, we have $u \in C^{0,\frac{m-2}{m-1}}(\tt)$ and the H\"older constant is dependent on $m$, $\|\sigma\|_{C(\tt\times \cA_0)}$, $\|b\|_{C(\tt\times \cA_0)}$, $\|L_0\|_{C(\tt\times \cA_0)}$, $\min_{\tt} c$, $\gl\|u^-\|_{L^\infty(\tt)}$ and $\partial \gO$ (see \cite{CaLePo}, \cite{Barl}, \cite{ArTr}). This important point helps verifying the equi-continuity assumption \eqref{EC}. It is clear that \eqref{F1}, \eqref{F2} and \eqref{L} hold. Thanks to Theorems 3.1, 3.2 and 4.1 in \cite{Barl}, \eqref{CP}, \eqref{CPS}, \eqref{CPD}, \eqref{SLS}, \eqref{SLD} and \eqref{EC} are valid. Besides, if $u$ is a subsolution of \eqref{D'} with $\gl \geq 0$, then $u \leq \psi$ pointwise on $\partial \gO$ (see \cite[Proposition 3.1]{DaL} or \cite[Proposition 3.1]{Barl} for the proof). For the Neumann problem, we assume further that $\gamma \in C^2(\partial \gO)$. By using the ideas in \cite{Barl} with careful modifications, we verify that \eqref{CPN'} and \eqref{SLN} are valid. Therefore, Theorems \ref{thm2-sc}, \ref{thm2-d}, \ref{conv-n} hold true. \subsection{Fully nonlinear, possibly degenerate elliptic equations with superlinear growth in the gradient variable} This example is only applicable to Neumann problem. Let $F_0$ be as in Subsection \ref{ss-superquad}. Suppose that \begin{align*} &F(x,p,X):= \frac{|p|^m }{m}+ F_0(x,p,X)\\ =&\, \sup_{(q,\ga)\in\R^n\times \cA_0}\left(- \tr\gs^t(x,\ga)\gs(x,\ga) X+q\cdot p - b(x,\ga)\cdot p -\frac{|q|^{m'}}{m'}- L_0(x,\ga)\right), \end{align*} where $m>1$ and $m':=m/(m-1)$. Assume further that $\partial \gO \in C^{1,1}$ and $\gamma \in C^{1,1}(\partial \gO)$. This will be studied in details in the forthcoming paper \cite{IsMtTr3}. It is verified there that all conditions in Theorem \ref{conv-n} are satisfied and thus Theorem \ref{conv-n} holds true. As far as the authors know, this example has not been studied in the literature. \subsection{Fully nonlinear, uniformly elliptic equations with linear growth in the gradient variable} \label{ss-fully} We note that this example is only applicable to Dirichlet problem and Neumann problem. We consider the operator \[ F(x,p,X):= \max_{\ga \in \cA} \left(- \tr a(x,\ga) X -b(x,\ga)\cdot p-L(x,\ga)\right), \] where $\cA$ is a compact metric space, $a\in \Lip(\tt \times \cA, \bS^n)$, $b\in \Lip(\tt \times \cA,\R^n)$, $L\in \Lip(\tt \times \cA,\R)$. Assume that $a$ is uniformly elliptic, that is, there exists $\theta>0$ such that \[ \frac{1}{\theta} I \le a(x,\ga)\le \theta I \quad\text{for all} \ (x,\ga) \in\tt \times \cA, \] where $I$ denotes the identity matrix of size $n$. It is obvious to see that \eqref{F1} and \eqref{F2} hold. For the Dirichlet problem, we assume that $\partial \gO \in C^{1,\beta}$ and $g \in C^{1,\beta}(\tt)$ for some $\beta \in (0,1)$. Comparison principles \eqref{CP}, \eqref{CPD} are consequences of \cite[Theorem III.1 (1)]{IsLi}, \cite[Theorem 7.9]{CrIsLi}, respectively, under the Lipschitz continuity assumption on $a$, $b$ and $L$. By \cite[Theorem 3.1]{Tr}, \eqref{SLD} holds true. In view of the Krylov-Safonov estimate (see \cite{KrSa} and also \cite{Ca,Tr}), we have that $\{v^\lambda\}_{\lambda>0}$ is equi-H\"older continuous on $\tt$, which implies (EC). Therefore, Theorem \ref{thm2-d} is valid. For the Neumann problem, we assume that $\partial \gO \in C^2$, $\gamma \in C^{0,1}(\partial \gO)$, and $g\in C^{0,\beta}(\partial \gO)$ for some $\beta \in (0,1)$. Thanks to \cite[Theorem 7.12]{CrIsLi}, \eqref{CPN'} and \eqref{SLN} are valid. Furthermore, in view of \cite[Theorem A.1]{BaDa}, $\{v^\lambda\}_{\lambda>0}$ is equi-H\"older continuous on $\tt$, which implies (EC). These yield the validity of Theorem \ref{conv-n}. \begin{bibdiv} \begin{biblist} \bib{AlAlIsYo}{article}{ author={Al-Aidarous, E. S.}, author={Alzahrani, E. O.}, author={Ishii, H.}, author={Younas, A. M. M.}, title={A convergence result for the ergodic problem for Hamilton--Jacobi equations with Neumann-type boundary conditions}, journal={Proc. Roy. Soc. Edinburgh Sect. A}, volume={146}, date={2016}, number={2}, pages={225--242}, } \bib{ArTr}{article}{ author={Armstrong, S. N.}, author={Tran, H. V.}, title={Viscosity solutions of general viscous Hamilton-Jacobi equations}, journal={Math. Ann.}, volume={361}, date={2015}, number={3-4}, pages={647--687}, issn={0025-5831}, } \bib{Barl}{article}{ AUTHOR = {Barles, G.}, TITLE = {A short proof of the {$C^{0,\alpha}$}-regularity of viscosity subsolutions for superquadratic viscous {H}amilton-{J}acobi equations and applications}, JOURNAL = {Nonlinear Anal.}, VOLUME = {73}, YEAR = {2010}, NUMBER = {1}, PAGES = {31--47}, ISSN = {0362-546X}, } \bib{BaDa}{article}{ AUTHOR = {Barles, G.}, AUTHOR={Da Lio, F.} TITLE = {On the boundary ergodic problem for fully nonlinear equations in bounded domains with general nonlinear {N}eumann boundary conditions}, JOURNAL = {Ann. Inst. H. Poincar\'e Anal. Non Lin\'eaire}, FJOURNAL = {Annales de l'Institut Henri Poincar\'e. Analyse Non Lin\'eaire}, VOLUME = {22}, YEAR = {2005}, NUMBER = {5}, PAGES = {521--541}, ISSN = {0294-1449}, } \bib{BaMi}{article}{ AUTHOR = {Barles, G.}, AUTHOR={Mitake, H.}, TITLE = {A {PDE} approach to large-time asymptotics for boundary-value problems for nonconvex {H}amilton-{J}acobi equations}, JOURNAL = {Comm. Partial Differential Equations}, FJOURNAL = {Communications in Partial Differential Equations}, VOLUME = {37}, YEAR = {2012}, NUMBER = {1}, PAGES = {136--168}, } \bib{Ca}{article}{ author={Caffarelli, L. A.}, title={Interior a priori estimates for solutions of fully nonlinear equations}, journal={Ann. of Math. (2)}, volume={130}, date={1989}, number={1}, pages={189--213}, issn={0003-486X}, } \bib{CaLePo}{article}{ author={Capuzzo-Dolcetta, I.}, author={Leoni, F.}, author={Porretta, A.}, title={H\"older estimates for degenerate elliptic equations with coercive Hamiltonians}, journal={Trans. Amer. Math. Soc.}, volume={362}, date={2010}, number={9}, pages={4511--4536}, } \bib{CrIsLi}{article}{ author={Crandall, M. G.}, author={Ishii, H.}, author={Lions, P.-L.}, title={User's guide to viscosity solutions of second order partial differential equations}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={27}, date={1992}, number={1}, pages={1--67}, } \bib{DaL}{article}{ AUTHOR = {Da Lio, F.}, TITLE = {Comparison results for quasilinear equations in annular domains and applications}, JOURNAL = {Comm. Partial Differential Equations}, VOLUME = {27}, YEAR = {2002}, NUMBER = {1-2}, PAGES = {283--323}, ISSN = {0360-5302}, } \bib{DFIZ}{article}{ author={Davini, A.}, author={Fathi, A.}, author={Iturriaga, R.}, author={Zavidovique, M.}, title={Convergence of the solutions of the discounted equation}, journal={Invent. Math. First online January, 2016 (Preprint is also available in arXiv:1408.6712)}, } \bib{Ev}{article}{ author={Evans, L. C.}, title={Classical solutions of fully nonlinear, convex, second-order elliptic equations}, journal={Comm. Pure Appl. Math.}, volume={35}, date={1982}, number={3}, pages={333--363}, issn={0010-3640}, } \bib{Go}{article}{ author={Gomes, D. A.}, title={Duality principles for fully nonlinear elliptic equations}, conference={ title={Trends in partial differential equations of mathematical physics}, }, book={ series={Progr. Nonlinear Differential Equations Appl.}, volume={61}, publisher={Birkh\"auser, Basel}, }, date={2005}, pages={125--136}, } \bib{GMT}{article}{ author={Gomes, D. A.}, author={Mitake, H.}, author={Tran, H. V.}, title={The selection problem for discounted Hamilton-Jacobi equations: some non-convex cases}, journal={preprint}, } \bib{Ishi}{article}{ author={Ishii, H.}, title={Weak {KAM} aspects of convex Hamilton-Jacobi equations with Neumann type boundary conditions}, journal={J. Math. Pures Appl. (9)}, volume={95}, date={2011}, number={1}, pages={99--135}, } \bib{IsLi}{article}{ author={Ishii, H.}, author={Lions, P.-L.}, title={Viscosity solutions of fully nonlinear second-order elliptic partial differential equations}, journal={J. Differential Equations}, volume={83}, date={1990}, number={1}, pages={26--78}, } \bib{IsMtTr1}{article}{ author={Ishii, H.}, author={Mitake, H.}, author={Tran, H. V.}, title={The vanishing discount problem and viscosity Mather measures. Part 1: the problem on a torus}, journal={submitted, (Preprint is available in arXiv:1603.01051)}, } \bib{IsMtTr3}{article}{ author={Ishii, H.}, author={Mitake, H.}, author={Tran, H. V.}, title={work under preparation}, } \bib{Kr}{article}{ author={Krylov, N. V.}, title={Boundedly inhomogeneous elliptic and parabolic equations}, language={Russian}, journal={Izv. Akad. Nauk SSSR Ser. Mat.}, volume={46}, date={1982}, number={3}, pages={487--523, 670}, } \bib{KrSa}{article}{ author={Krylov, N. V.}, author={Safonov, M. V.}, title={An estimate for the probability of a diffusion process hitting a set of positive measure}, language={Russian}, journal={Dokl. Akad. Nauk SSSR}, volume={245}, date={1979}, number={1}, pages={18--20}, } \bib{LPV}{article}{ author={P.-L. Lions}, author={G. Papanicolaou}, author={S. R. S. Varadhan}, title={Homogenization of Hamilton--Jacobi equations}, journal={unpublished work (1987)}, } \bib{Man}{article}{ author={R. Ma\~n\'e}, title={Generic properties and problems of minimizing measures of Lagrangian systems}, journal={Nonlinearity}, volume={9}, date={1996}, NUMBER = {2}, pages={273--310}, } \bib{Mat}{article}{ author={J. N. Mather}, title={Action minimizing invariant measures for positive definite Lagrangian systems}, journal={Math. Z.}, volume={207}, date={1991}, NUMBER = {2}, pages={169--207}, } \bib{MiTr}{article}{ author={Mitake, H.}, author={Tran, H. V.}, title={Selection problems for a discounted degenerate viscous Hamilton--Jacobi equation}, journal={submitted, (Preprint is available in arXiv:1408.2909)}, } \bib{Si}{article}{ author={Sion, M.}, title={On general minimax theorems}, journal={Pacific J. Math.}, volume={8}, date={1958}, pages={171--176}, } \bib{Te}{article}{ author={Terkelsen, F.}, title={Some minimax theorems}, journal={Math. Scand.}, volume={31}, date={1972}, pages={405--413 (1973)}, } \bib{Tr}{article}{ author={Trudinger, N. S.}, title={On regularity and existence of viscosity solutions of nonlinear second order, elliptic equations}, conference={ title={Partial differential equations and the calculus of variations, Vol.\ II}, }, book={ series={Progr. Nonlinear Differential Equations Appl.}, volume={2}, publisher={Birkh\"auser Boston, Boston, MA}, }, date={1989}, pages={939--957}, } \end{biblist} \end{bibdiv} \bye
455ebde23200f161d094cc2901e235006d738cdd
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Pre-class preparation positions students to get the most out of face-to-face class time \cite{Schwartz1998}. This is particularly important for active learning classrooms, where students rely on their prior learning to participate in peer discussion and construct their own further understanding of key concepts. The standard approach to foster student preparedness is the assignment of pre-class reading. This is an evidence-based practice: in courses using pre-assignments coupled with quizzes, students complete the reading more consistently \cite{HeinerAJP2014,MarcellIJSoTL2008}, ask more---and more challenging---questions during class \cite{MarcellIJSoTL2008,Narloch2006}, and perform better on course assessments and exams \cite{HeinerAJP2014,Narloch2006,Dobson2008,Johnson2009}. Notably, a common thread among most studies is in the use of a relatively passive modality in pre-class assignments---textbook reading or watching videos. While it is possible that students may actively engage---by taking notes, developing questions, and making connections with the material---there is no guarantee that they are using these types of strategies. Additionally, they may have low intrinsic motivation to deeply engage, given the minimal opportunity for inquiry and exploration. As a possible solution, multimedia learning modules used for pre-lecture assignments were shown to promote learning and to improve student attitudes \cite{StelzerAJP2009,StelzerAJP2010}. Thus, there is a clear opportunity to deliberately use non-passive tools in the context of pre-class preparation assignments to support active engagement, inquiry, and other positive student outcomes. Simulations offer one such tool. The PhET Interactive Simulations have been carefully designed \cite{PerkinsPhysTeach2006} with dual goals of student engagement and learning. The results have been positive: PhETs promote learning, especially when students are exploring in a manner driven by their own questioning \cite{WiemanScience,AdamsJILC2008}. Additionally, the nature of PhET-based assignments that promote productive engagement has been investigated. In interviews, maximal engagement was seen in the presence of minimal---but non-zero---guidance \cite{AdamsAIP2008}. A separate in-class study provided students with either light-, moderate-, or heavy-guidance PhET-based assignments; students in the light guidance condition explored the PhET more and paid more attention to their interactions with the simulation \cite{ChamberlainChem2014}. Adams {\it et al.}~found that students can learn from PhETs at home, in an unstructured environment, and that moderate, question-driven guidance may be best in this context \cite{AdamsPERC2015}. \begin{table*}[htbp] \caption{Pre- and post-test scores by topic. The pre-test was administered at the beginning of the semester, and the identical post-test occurred immediately after the pre-class assignment. $N$ is the number of students who completed all of the pre-test, post-test, and in-class test for the topic. A Wilcoxon signed-rank test shows that, for all topics, students learned during the pre-class assignments.}\label{tab1} \begin{ruledtabular} \begin{tabular}{lccccccccc} && \multicolumn{3}{c}{{\bf Pre-test}} & \multicolumn{3}{c}{{\bf Post-test}} \\ {\bf Topic} & $N$ & {\bf Mean} & {\bf Median} & {\bf SD} & {\bf Mean} & {\bf Median} & {\bf SD} & $p$ & {\bf Effect size}\\ \hline Blackbody Radiation & 395 & 42\% & 33\% & 29\% & 53\% & 66\% & 29\% & $<.001$ & -.22 \\ Masses and Springs & 324 & 21\% & 0\% & 32\% & 43\% & 50\% & 40\% & $<.001$ & -.33 \\ Resonance & 342 & 31\% & 33\% & 28\% & 44\% & 33\% & 32\% & $<.001$ & -.22 \end{tabular} \end{ruledtabular} \end{table*} Our goal in this work is to leverage the capacity of PhET simulations to create productive engagement and learning to help better prepare students for class. We add active learning, rather than just passive reading, to the pre-class assignments by including inquiry-oriented PhET-based activities, giving students an opportunity to explore within the constrained environment of the PhET. It is plausible that this type of hands-on exploration can promote a conceptual understanding that helps students piece together the knowledge they encounter in the textbook reading and during in-class instruction. Specifically, we investigate if the use of PhET-based activities in conjunction with pre-class reading assignments can improve both the preparation of students for in-class learning and student attitudes towards and engagement with pre-class assignments. \section{Method} This work took place at a large research-intensive university in the Pacific Northwest. The study was conducted in a one-semester, multi-section first-year physics course for engineering students on the subject of thermodynamics and periodic motion. Each of three lecture sections for the course contained about 260 students, with a total of 779 students in the course. Within this course, for each of three different modules (Blackbody Radiation, Masses and Springs, and Resonance), we created assignments with two pieces, based on either the textbook or a PhET simulation on the same topic. The text pre-reading was a standard pre-class reading assignment, which included targeted questions that directed students to engage with specific passages, vocabulary, concepts, and figures in the book \cite{HeinerAJP2014}. The PhET assignment directed students to interact with the relevant PhET simulation by providing a hyperlink to the PhET and prompting them to explore particular relationships within the simulation and record their observations in an open-response text box. For example, for the topic Masses and Springs, the prompt was: ``Explore the relationships between the spring softness, mass, amplitude, and period of oscillation. In the space below, describe 2-3 interesting things you noticed." In comparison to the different levels of guidance described in \cite{ChamberlainChem2014}, our PhET assignments were closest to their ``light guidance'' condition. While textbook-based pre-reading assignments are standard in this course, and PhETs have been recommended to students in previous iterations of the course, the explicit use of the PhET assignment was novel in this course. These two styles of activities---Text and PhET---engage the students differently with a topic: reading a textbook transmits knowledge whereas PhET simulations reveal knowledge through exploration of the concepts. We sought to test the hypotheses that 1) inclusion of the PhET assignment would improve student performance and engagement over a text-only pre-reading, and 2) when both PhET and text are assigned, the relative order of the two would impact student performance. Students were randomly assigned to three treatment groups: text pre-reading assignment alone; PhET activity before text pre-reading; and text pre-reading before PhET activity. The experiment was repeated over three course modules and each student group experienced each treatment once over the course of the study. To ensure the correct order of completion, students were not given access to their second part of the assignment (text or PhET) until they had completed the first. Student performance and perspectives data were collected from these students throughout the semester. To assess baseline student understanding ("pre-test"), students completed an online 8-question multiple choice pre-test early in the semester, prior to any exposure to the study topics. The test, graded for participation, included three questions each related to Blackbody Radiation and Resonance and two questions related to Masses and Springs. Subsequently, within each module, student learning was assessed immediately following the experimental treatment, but before in-class instruction ("post-test"), and again following in-class instruction ("in-class test"). The post-test was completed as part of a normal online pre-class quiz, and consisted of the identical questions from the cognate topics in the pre-test. The in-class test was a series of 3-4 clicker questions, given in class 2-3 days later. Students were directed to work independently for these questions. Student perspectives data were collected using an online survey (for participation grades) administered with the post-test. Students were asked 1) to self-report the amount of time spent on both parts of the assignment, and 2) their perceived enjoyment and learning from the entire pre-class assignment. The perspectives data were collected only for the latter two modules (Masses and Springs; Resonance). For the analysis of learning outcomes, only students who completed all performance assessments for a topic (pre-test, post-test, and in-class test) were included in the study cohort. For each student, a raw score (percent correct) was computed for each assessment, for each topic. Only students who completed all survey questions for a topic were included in the analysis of perceptions ($N=408$ for Masses and Springs and $N=407$ for Resonance). \section{Results} \subsection{Learning} Over all three topics, students learned during the pre-class assignment, as measured by an increase in score from pre-test to post-test. These results are shown in Table~\ref{tab1}. To evaluate whether or not student learning during the pre-class assignment depended on the treatment (text reading alone, PhET activity then text reading, or text reading then PhET activity), the following linear model was used \begin{align}\label{model} &\textrm{Post-test}_{ijk}=\beta_0+\beta_{1}\times\textrm{Pre-test}_{ij}+\beta_{2,j}\times\textrm{Topic}_j\\ &\quad\;\;+\beta_{3,k}\times\textrm{Treatment}_k+\beta_{4,jk}\times\textrm{Topic}_j\textrm{Treatment}_k+\varepsilon_i,\nonumber \end{align} where $\textrm{Post-test}_{ijk}$ is the post-test score of student $i$ on $\textrm{Topic}_j$, $\textrm{Pre-test}_{ij}$ is the pre-test score for student $i$ on $\textrm{Topic}_j$, $\textrm{Topic}_j$ is a categorical variable representing the topic (Blackbody Radiation, Masses and Springs, or Resonance), $\textrm{Treatment}_k$ is a categorical variable representing the treatment condition, and $\varepsilon_i$ is a random intercept for student $i$ which accounts for differences in students. We include a $\textrm{Topic}_j\textrm{Treatment}_k$ interaction term to take into account that the different PhETs may promote learning differently. There was no significant effect for treatment, when controlling for pre-test and topic, $F(2, 695) = 1.1$, $p=.33$, nor for the interaction between treatment and topic, $F(2, 695) = 0.041$, $p=.96$. As expected, pre-test was a highly significant predictor of learning, $F(1, 695) = 50$, $p<.001$, $\eta^2 = .067$, which corresponds to a medium effect size. To evaluate whether or not the type of pre-class assignment influenced student learning in-class, we used a similar model as (\ref{model}), replacing post-test with the in-class clicker test as the dependent variable. As for the post-test, no evidence for an effect of the treatment on the in-class test results was found. \subsection{Engagement and perceptions} Over all treatment conditions, the students reported spending a mean time of 34.4 min (median = 30 min, SD = 29.0 min) on the textbook reading. Over the two conditions involving PhET assignments, students reported spending a mean time of 10.3 min (median = 10 min, SD = 10.7 min) on the PhET activity. For the topic Resonance, when their pre-class assignment included the PhET activity, students reported spending more time overall on the pre-class assignment (Table~\ref{tab2}). \begin{table}[htbp] \caption{Time on task by topic and condition. The overall time for each PhET treatment is compared to the text-only treatment with a Mann-Whitney test, with effect size (rank-biserial correlation) $r$. For the Resonance topic, when the PhET activity was included, students reported longer time-on-task with the pre-class activity.\label{tab2}} \begin{ruledtabular} \begin{tabular}{llcccccc} & \multicolumn{5}{c}{{\bf Overall Reported Time on task (min)}} \\ {\bf Treatment} & $N$ & {\bf Mean} & {\bf SD} & $p$ & $r$\\ \hline \textbf{Masses \& Springs} \\ Text only & 144 & 45.9 & 31.7 \\ PhET then text & 137 & 51.4 & 42.6 & .47 & -.050 \\ Text then PhET & 127 & 44.0 & 29.6 & .65 & .032\\ \hline \textbf{Resonance} \\ Text only & 128 & 29.5 & 19.8 \\ PhET then text & 143 & 35.5 & 24.3 & .009 & -.18\\ Text then PhET & 136 & 40.8 & 33.8 & <.001 & -.29 \end{tabular} \end{ruledtabular} \end{table} The time students spent engaged with the PhET activity did not seem to depend on the order of the assignments: Students did spend slightly more time on PhET if it was before the Text assignment, though not significantly, Mann-Whitney $p=.087$, $r=.08$. To investigate if engagement with the PhET assignment promoted learning during the pre-class assignment, the model in (\ref{model}) was run with PhET engagement as a predictor. No evidence for an effect of PhET engagement on learning from the pre-class assignment was found (as measured on either the post-test or in-class test). Students reported that they enjoyed the pre-class assignments and found them beneficial to their learning. Splitting by treatment, 71\% of students whose assignment included a PhET activity rated the pre-class assignment as enjoyable compared to 62\% of students in the textbook-only group. Overall, 74\% of students found the pre-class assignments beneficial to their learning, with no difference in perceived learning benefit across treatment groups. Within the two PhET treatments, there is a correlation between engagement with the PhET activity and how beneficial to their learning and enjoyable students found the pre-class assignment (Table~\ref{tab3}): Students who reported the assignment as beneficial to their learning (enjoyable) spent more time with the PhET activity than students who reported the assignment as not beneficial to their learning (not enjoyable). \begin{table}[htbp] \caption{Time on task with the PhET activity split on perceptions of the pre-class assignment. By a Mann-Whitney test, students who reported the assignment as beneficial to their learning (enjoyable) spent more time with the PhET activity than students who reported the assignment as not beneficial to their learning (not enjoyable).\label{tab3}} \begin{ruledtabular} \begin{tabular}{lccccc} & \multicolumn{5}{c}{{\bf PhET Reported Time on task (min)}} \\ {\bf Perception} & $N$ & {\bf Mean} & {\bf SD} & $p$ & $r$\\ \hline Beneficial to learning & 404 & 11.1 & 11.8\\ Not beneficial to learning & 139 & 7.96 & 6.25 & <.001 & -.24\\\hline Enjoyable & 384 & 10.7 & 8.35\\ Not enjoyable & 159 & 9.20 & 14.9 & <.001 & -.22 \end{tabular} \end{ruledtabular} \end{table} \section{Discussion} From pre- to post-test, students learned during the pre-class assignments, with a pedagogically significant effect size. This indicates that the assignments were indeed successful in helping prepare students for class. Although we hypothesized that PhETs could contribute to learning during the pre-class assignments, our analysis revealed no effect on our measures of learning relative to the textbook-only assignments. It could be that, for the at-home pre-class assignment, our light guidance did not provide the scaffolding students would need to productively explore within the PhET activity. If this were the case, the Text then PhET treatment, for which reading the textbook before the PhET activity may provide extra scaffolding, may have had a better chance of providing the structure students would need for productive engagement with the PhET. However, no learning effect was seen whether students were constrained to do the PhET activity before or after the textbook reading, so it may be that, no matter the order, the level of guidance was too light for this assignment. In their study of PhET activities in unstructured environments, Adams {\it et al.}~were unable to create effective activities with light guidance \cite{AdamsPERC2015}; it may be that more direct instruction is necessary for pre-class activities as well. It is possible that, although the assignment was constrained, students referred to the textbook through all parts of the assignment. Then, by also having a textbook assignment, productive exploration may have been undermined. Because the textbook part of the assignment is important to students' preparation for the class---so that they have encountered basic definitions and concepts---it may be difficult to disentangle the learning specific to the PhET activity. It may also be that our assessment questions, while appropriate for the topic in general, did not target the specific concepts addressed in the PhETs, or that these particular PhETs did not provide exploration opportunities at a level appropriate for this audience. Including a PhET activity in the pre-class assignment influenced both student engagement with and perceptions of the learning benefits of the assignments. When PhET was included, more students reported the pre-class assignments as enjoyable and as many students as the text-only condition rated them as beneficial to their learning. Additionally, when PhET was included in the Resonance topic, students reported spending 5--10 minutes longer with the assignments. This is notable, as Resonance was the most difficult of the topics and was encountered late in the semester, when competing priorities and general fatigue may have prevented students from engaging. The affective benefits of including PhET may have helped to sustain the motivation of these students. Interestingly, the reported time spent on the PhET activity did not seem to depend on the ordering of activities. If completing the textbook reading before the PhET activity affected how students interacted with the PhET, it did not show up here. Overall, students in both PhET conditions reported spending marginally less time on the textbook part of the assignment. This is odd, because students in both the Text or Text then PhET treatment did not know if they were going to have a PhET activity, so we would expect the textbook engagement for these treatments to match. Since their perceived time spent on the assignments was recorded after finishing all activities, it appears that having PhET as part of the assignment altered how students thought about the assignment as a whole. There was an interesting correlation between time spent on the PhET and perceptions of the pre-class assignments. It could be that students who enjoyed the assignment, topic, or PhET more were pre-disposed to spend more time with the simulation, or it could be that students who spent more time with the PhET ended up enjoying the assignment more. More study is needed to understand how student engagement and attitudes interact. No effect of PhET engagement on learning was observed, indicating that the extra time spent did not translate into learning gains on our measures. \section{Conclusion} PhETs are not a magic bullet: They have been shown to promote learning with carefully designed activities in particular contexts. This report shows that even purposeful inclusion of PhETs in pre-class assignments does not automatically translate into learning. In the case of pre-class assignments, more study is required to see if it is possible to design supplemental PhET activities that improve student preparation beyond that from a text-only assignment. The inclusion of PhET did have an effect on both student engagement with the pre-class assignment and student perceptions of the assignments. These affective results suggest that PhETs did contribute meaningfully to these assignments. Sustaining student motivation can be a challege, so the potential for PhETs to contribute in this domain is an important consideration. \acknowledgments{We thank Don Witt and Andrzej Kotlicki for early conversations and support in this project.}
ec5d1320adc0905ea18c0f8c4ffbb5b753e95f23
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} We investigate the problem of simulation from a multivariate normal (MVN) distribution whose samples are restricted to the intersection of a set of hyperplanes, which is shown to be inherently related to the simulation of a conditional distribution of a MVN distribution. A naive approach, which linearly transforms a random variable drawn from the conditional distribution of a related MVN distribution, requires a large number of intermediate variables that are often computationally expensive to instantiate. To address this issue, we propose a fast and exact simulation algorithm that directly projects a MVN random variable onto the intersection of a set of hyperplanes. We further show that sampling from a MVN distribution, whose covariance (precision) matrix can be decomposed as the sum (difference) of a positive-definite matrix, whose inversion is known or easy to compute, and a low-rank symmetric matrix, may also be made significantly fast by exploiting this newly proposed stimulation algorithm for hyperplane-truncated MVN distributions, avoiding the need of Cholesky decomposition that has a computational complexity of $O(k^3)$ \citep{golub2012matrix}, where $k$ is the dimension of the MVN random variable. Related to the problems under study, the simulation of MVN random variables subject to certain constraints \citep{gelfand1992bayesian} has been investigated in many other different settings, such as multinomial probit and logit models \citep{albert1993bayesian,mcculloch2000bayesian,imai2005bayesian,train2009discrete,holmes2006bayesian,johndrow2013diagonal}, Bayesian isotonic regression \citep{neelon2004bayesian}, Bayesian bridge \citep{polson2014bayesian}, blind source separation \citep{schmidt2009linearly}, and unmixing of hyperspectral data \citep{altmann2014sampling, dobigeon2009joint}. A typical example arising in these different settings is to sample a random vector $\xv\in\mathbb{R}^k$ from a MVN distribution subject to $k$ inequality constraints as \begin{equation} \xv\sim\Nor_{\mathcal{S}}(\muv,\Sigmamat),~~\mathcal{S}=\{\xv:\lv\le \Gmat \xv \le \uv\}, \end{equation} where $\Nor_{\mathcal{S}}(\muv,\Sigmamat)$ represents a MVN distribution truncated on the sample space $\mathcal{S}$, $\muv\in\mathbb{R}^k$ is the mean, $\Sigmamat\in\mathbb{R}^{k\times k}$ is the covariance matrix, $\Gmat\in\mathbb{R}^{k\times k}$ is a full-rank matrix, $\lv\in \mathbb{R}^k$, $\uv\in \mathbb{R}^k$, and $\lv<\uv$. If the elements of $\lv$ and $\uv$ are permitted to be $-\infty$ and $+\infty$, respectively, then both single sided and fewer than $k$ inequality constraints are allowed. Equivalently, as in \citet{geweke1991efficient,geweke1996bayesian}, one may let $\xv = \muv + \Gmat^{-1} \zv$ and use Gibbs sampling \citep{Geman,gelfand1990sampling} to sample the $k$ elements of $\zv$ one at a time conditioning on all the others from a univariate truncated normal distribution, for which efficient algorithms exist \citep{robert1995simulation,damien2001sampling,chopin2011fast}. To deal with the case that the number of linear constraints imposed on $\xv$ exceed its dimension $k$ and to obtain better mixing, one may consider the Gibbs sampling algorithm for truncated MVN distributions proposed in \citet{rodriguez2004efficient}. In addition to Gibbs sampling, to sample truncated MVN random variables, one may also consider Hamiltonian Monte Carlo \citep{pakman2014exact,lan2014spherical} and a minimax tilting method proposed in \citet{botev2016normal}. \section{Hyperplane-truncated and conditional MVNs}\label{sec:2} For the problem under study, we express a $k$-dimensional MVN distribution truncated on the intersection of $k_2<k$ hyperplanes as \begin{equation}\label{eq:x_S} \xv\sim\Nor_{\mathcal{S}}(\muv,\Sigmamat),~~\mathcal{S}=\{\xv: \Gmat \xv = \rv\}, \end{equation} where $$ \Gmat\in\mathbb{R}^{k_2\times k}, ~~\rv\in\mathbb{R}^{k_2}, $$ and $\mbox{Rank}(\Gmat)=k_2$. The probability density function can be expressed as \begin{align}\label{eq:px} p( \xv \,|\, \muv,\Sigmamat, \Gmat,\rv) =\frac{1}{Z} \exp\left[-\frac{1}{2}(\xv- \muv)^T \Sigmamat^{-1} (\xv- \muv) \right] \delta(\Gmat\xv = \rv), \end{align} where $Z$ is a constant ensuring $\int p( \xv \,|\, \muv,\Sigmamat, \Gmat,\rv)d\xv = 1$, and $\delta(x)=1$ if the condition $x$ is satisfied and $\delta(x)=0$ otherwise. Let us partition $\Gmat$, $\xv$, $\muv$, $\Sigmamat$, and $\Lambdamat = \Sigmamat^{-1}$ as $$ \Gmat = (\Gmat_1,\Gmat_2),~ \xv=\begin{bmatrix} \xv_1 \\ \xv_2 \end{bmatrix},~ \muv = \begin{bmatrix} \muv_1 \\ \muv_2 \end{bmatrix},~ \Sigmamat = \begin{bmatrix} \Sigmamat_{11} &\Sigmamat_{12} \\ \Sigmamat_{21} & \Sigmamat_{22} \end{bmatrix},~ \mbox{and}~ \Lambdamat = \begin{bmatrix} \Lambdamat_{11} & \Lambdamat_{12} \\ \Lambdamat_{21} & \Lambdamat_{22} \end{bmatrix}, $$ whose sizes are $$ (k_2\times k_1,k_2\times k_2),~ \begin{bmatrix} k_1 \times 1 \\ k_2 \times 1 \end{bmatrix},~ \begin{bmatrix} k_1 \times 1\\ k_2\times 1 \end{bmatrix},~ \begin{bmatrix} k_1\times k_1&k_1\times k_2 \\ k_2\times k_1& k_2\times k_2 \end{bmatrix},~\mbox{and}~ \begin{bmatrix} k_1\times k_1&k_1\times k_2 \\ k_2\times k_1& k_2\times k_2 \end{bmatrix}, $$ respectively, where $k=k_1+k_2$, $\Sigmamat_{21} = \Sigmamat_{12}^T$, and $\Lambdamat_{21} = \Lambdamat_{12}^T$. A special case that frequently arises in real applications is when $\Gmat_1=\mathbf{0}_{k_2\times k_1}$ and $\Gmat_2= \Imat_{k_2}$, which means $(\mathbf{0}_{k_2\times k_1},\Imat_{k_2})\xv=\xv_2=\rv$ and the need is to simulate $\xv_1$ given $\xv_2=\rv$. For a MVN random variable $ \xv\sim\mathcal{N}(\muv,\Sigmamat) $, it is well known, $e.g.$, in \citet{tong2012multivariate}, that the conditional distribution of $\xv_1$ given $\xv_2=\rv$, $i.e.$, the distribution of $\xv$ restricted to $\mathcal{S}=\{\xv: (\mathbf{0}_{k_2\times k_1},\Imat_{k_2})\xv = \rv\}$, can be expressed as \begin{equation}\label{eq:conditional} \xv_1 \,|\, \xv_2 = \rv \,\sim\,\Nor \left[\muv_1 + \Sigmamat_{12}\Sigmamat_{22}^{-1}(\rv-\muv_2), ~~\Sigmamat_{11} - \Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}\right]. \end{equation} Alternatively, applying the Woodbury matrix identity to relate the entries of the covariance matrix $\Sigmamat$ to those of the precision matrix $\Lambdamat$, one may obtain the following equivalent expression as \begin{equation}\label{eq:conditionalPrecision} \xv_1 \,|\, \xv_2 = \rv \,\sim\,\Nor \left[\muv_1 - \Lambdamat_{11}^{-1} \Lambdamat_{12} (\rv-\muv_2), ~\Lambdamat_{11}^{-1}\right]. \end{equation} In a general setting where $\Gmat\neq ( \mathbf{0}_{k_2\times k_1}, \Imat_{k_2})$, let us define a full rank linear transformation matrix $\Hmat\in\mathbb{R}^{k\times k}$, with $(\Hmat_1,\Hmat_2)$ as the $(k\times k_1,k\times k_2)$ partition of $\Hmat$, where the columns of $\Hmat_1\in\mathbb{R}^{k\times k_1}$ span the null space of the $k_2$ rows of $\Gmat$, making $\Gmat\Hmat=(\Gmat\Hmat_1,\Gmat\Hmat_2)=(\mathbf{0}_{k_2\times k_1},\Gmat\Hmat_2)$, where $\Gmat\Hmat_2$ is a $k_2\times k_2$ full rank matrix. \linebreak[4] For example, a linear transformation matrix $\Hmat$ that makes $\Gmat\Hmat=(\mathbf{0}_{k_2\times k_1},\Imat_{k_2})$ can be constructed using the command $\mathrm{\Hmat= inv([null(\Gmat)';\Gmat]}$) in Matlab and \linebreak[4] $\mathrm{\Hmat<\!\!-solve(rbind(t(Null(t(\Gmat))),\Gmat))}$ in R. With $\Hmat $ and $\Hmat^{-1}$, one may re-express the constraints as $\mathcal{S}=\{\xv: (\mathbf{0}_{k_2\times k_1},\Gmat\Hmat_2) (\Hmat^{-1}\xv) = \rv\}$. Denote $\zv=\Hmat^{-1}\xv$, then we can generate $\xv$ by letting $\xv=\Hmat \zv$, where \begin{equation} \label{eq:pztruccond} \zv\sim\Nor_{\mathcal{D}}[\Hmat^{-1}\muv, \Hmat^{-1}\Sigmamat (\Hmat^{-1})^T],~~ \mathcal{D}=\{\zv: \Gmat\Hmat_2\zv_2=\rv\}=\{\zv: \zv_2=(\Gmat\Hmat_2)^{-1}\rv\}. \end{equation} More specifically, denoting $\Lambdamat = [\Hmat^{-1}\Sigmamat(\Hmat^{-1})^T]^{-1} = \Hmat^T \Sigmamat^{-1}\Hmat$ as the precision matrix for $\zv$, we have \begin{align}\label{eq:precision} \begin{bmatrix} \Lambdamat _{11} &\Lambdamat _{12} \\ \Lambdamat _{21} & \Lambdamat _{22} \end{bmatrix} &= \Hmat^T\Sigmamat^{-1} \Hmat = \begin{bmatrix} \Hmat_1^T\Sigmamat^{-1}\Hmat_1 &\Hmat^T_1\Sigmamat^{-1} \Hmat_2\\ \Hmat_2^T\Sigmamat^{-1}\Hmat_1 &\Hmat_2^T\Sigmamat^{-1} \Hmat_2 \end{bmatrix}, \end{align} and hence $\xv$ truncated on $\mathcal{S}$ can be naively generated using the following algorithm, whose computational complexity is described in Table \ref{tab:CompAlg1} of the Appendix. \begin{algorithm}[H] \begin{itemize} \item Find $\Hmat=(\Hmat_1,\Hmat_2)$ that satisfies $\Gmat\Hmat=(\Gmat\Hmat_1,\Gmat\Hmat_2)=(\mathbf{0}_{k_2\times k_1},\Gmat\Hmat_2)$, where $\Gmat\Hmat_2$ is a full rank matrix; \item Let $\zv_2 =(\Gmat \Hmat_2)^{-1}\rv$, $\Lambdamat _{11} = \Hmat_1^T\Sigmamat^{-1}\Hmat_1$, and $\Lambdamat _{12}=\Hmat_1^T\Sigmamat^{-1}\Hmat_2$; \item Sample $\zv_1\,|\, \zv_2=(\Gmat \Hmat_2)^{-1}\rv \sim\Nor( \muv_{z_1}, \Lambdamat^{-1}_{11})$, where $$\muv_{z_1}=(\Imat_{k_1},\mathbf{0}_{k_1\times k_2})\Hmat^{-1}\muv - \Lambdamat^{-1}_{11}\Lambdamat_{12}\left[(\Gmat\Hmat_2)^{-1}\rv - (\mathbf{0}_{k_2\times k_1},\Imat_{k_2})\Hmat^{-1}\muv\right];$$ \item Return $\xv = \Hmat\zv=\Hmat_1 \zv_1 + \Hmat_2 (\Gmat \Hmat_2)^{-1}\rv $. \end{itemize} \caption{\label{alg:1} Simulation of the hyperplane truncated MVN distribution $\xv\sim\Nor_{\mathcal{S}}(\muv,\Sigmamat)$, where $\mathcal{S}=\{\xv: \Gmat \xv = \rv\}$, by transforming a random variable drawn from the conditional distribution of another MVN distribution.} \end{algorithm}\vspace{2mm} For illustration, we consider a simple 2-dimensional example with $\muv = (1,1.2)^T$, $\Sigmamat = [(1,0.3)^T,(0.3,1)^T] $, $\Gmat = (1,1)$, and $\rv = 1$. If we choose $\Hmat_1 = (-0.7071,0.7071)^T$ and $\Hmat_2=(1.3,1.3)^T$, then we have $z_2 = (\Gmat \Hmat_2)^{-1}\rv = (2.6)^{-1} = 0.3846$, $\Lambdamat_{11} = 1.4285$, and $\Lambdamat_{12}=0$; as shown in Figure~\ref{fig:PxPz}, we may generate $\xv$ using $$\xv = (-0.7071,0.7071)^T z_1 + (1.3,1.3)^Tz_2$$ where $z_1\sim\Nor(0.1414,0.7)$ and $z_2=0.3846$. \begin{figure}[!t] \centering \subfigure[]{ \includegraphics[height = 4.5 cm]{pxOri.pdf} } \subfigure[]{ \includegraphics[height = 4.5 cm]{pzIndep.pdf} } \caption{\small Illustration of (a) $p(\xv)$ in \eqref{eq:px}, where $\muv = (1,1.2)^T$, $\Sigmamat = [(1,0.3)^T,(0.3,1)^T] $, $\Gmat = (1,1)$, and $\rv = 1$, and (b) $p(\zv)$ in \eqref{eq:pztruccond}, where $\Hmat_1 = (-0.7071,0.7071)^T$, $\Hmat_2=(1.3,1.3)^T$, and $\Hmat^{-1}=[(-0.7071,0.3846)^T, (0.7071,0.3846)^T]$. The coordinate systems of $\xv$ and $\zv$ are shown in black and red, respectively, and the first and second axes of a coordinate system are shown as dotted and dashed lines, respectively.} \label{fig:PxPz} \end{figure} For high dimensional problems, however, Algorithm \ref{alg:1} in general requires a large number of intermediate variables that could be computationally expensive to compute. In the following discussion, we will show how to completely avoid instantiating these intermediate variables. \section{Fast and exact simulation of MVN distributions } Instead of using Algorithm \ref{alg:1}, we first provide a theorem to show how to efficiently and exactly simulate from a hyperplane-truncated MVN distribution. In the Appendix, we provide two different proofs. The first proof facilitates the derivations by employing an existing algorithm of \citet{hoffman1991constrained} and \citet{doucet2010note}, which describes how to simulate from the conditional distribution of a MVN distribution shown in \eqref{eq:conditional} without computing $\Sigmamat_{11} - \Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}$ and its Cholesky decomposition. Note it is straightforward to verify that the algorithm in \citet{hoffman1991constrained} and \citet{doucet2010note}, as shown in the Appendix, can be considered as a special case of the proposed algorithm with $\Gmat = [\bds 0, \Imat]$. \vspace{2mm} \begin{algorithm}[H] \begin{itemize} \item Sample $\yv\sim\Nor (\muv,\Sigmamat)$; \item Return $\xv = \yv + \Sigmamat \Gmat^T ( \Gmat \Sigmamat \Gmat^T )^{-1} (\rv - \Gmat \yv)$, which can be realized using \begin{itemize} \item Solve $\alphav$ such that $( \Gmat \Sigmamat \Gmat^T ) \alphav = \rv - \Gmat \yv$; \item Return $\xv = \yv + \Sigmamat \Gmat^T \alphav$. \end{itemize} \end{itemize} \caption{\label{alg:2} Simulation of the hyperplane truncated MVN distribution $\xv\sim\Nor_{\mathcal{S}}(\muv,\Sigmamat)$, where $\mathcal{S}=\{\xv: \Gmat \xv = \rv\}$, by transforming a random variable drawn from $\yv\sim\Nor(\muv,\Sigmamat)$. } \end{algorithm}\vspace{2mm} \begin{thm} \label{maintheorem1} Suppose $\xv$ is simulated with Algorithm \ref{alg:2}, then it is distributed as $ \xv\sim\Nor_{\mathcal{S}}(\muv,\Sigmamat),~~\mathcal{S}=\{\xv: \Gmat \xv = \rv\}, $ where $ \Gmat\in\mathbb{R}^{k_2\times k}, ~~\rv\in\mathbb{R}^{k_2}, $ and $\mbox{Rank}(\Gmat)=k_2<k$. \end{thm} The above algorithm and theorem, whose computational complexity is described in Table \ref{tab:CompAlg2} of the Appendix, show that one may draw $\yv$ from the unconstrained MVN as $\yv\sim\Nor ( \muv , \Sigmamat )$ and directly map it to a vector $\xv$ on the intersection of hyperplanes using $ \xv =\Sigmamat \Gmat^T (\Gmat \Sigmamat \Gmat^T )^{-1} \rv +\left[ \Imat-\Sigmamat \Gmat^T ( \Gmat \Sigmamat \Gmat^T )^{-1}\Gmat\right]\yv.\notag $ For illustration, with the same $\muv$, $\Sigmamat$, $\Gmat$, and $\rv$ as those in Figure \ref{fig:PxPz}, we show in Figure \ref{fig:DemoAlg2} a simple two dimensional example, where the unrestricted Gaussian distribution $ \Nor ( \muv , \Sigmamat )$ is represented with a set of ellipses, and the constrained sample space $\mathcal{S}$ is represented as a straight line in the two-dimensional setting. With $\Sigmamat \Gmat^T (\Gmat \Sigmamat \Gmat^T )^{-1} \rv = (0.5,0.5)^T$, $\left[ \Imat-\Sigmamat \Gmat^T ( \Gmat \Sigmamat \Gmat^T )^{-1}\Gmat\right] = [(0.5,-0.5)^T,(-0.5,0.5)^T]$, one may directly maps a sample $\yv \sim \Nor ( \muv , \Sigmamat )$ to a vector on the constrained space. For example, if $\yv = (1,2)^T$, then it would be mapped to $\xv = (0,1)^T$ on the straight line. \begin{figure}[h] \centering \includegraphics[height = 5 cm]{pxOriProjYtoX.pdf}\\ \caption{A two dimensional demonstration of Algorithm \ref{alg:2} that maps a random sample from $\yv\sim\Nor ( \muv , \Sigmamat )$ to a sample in the constrained space using $ \xv =\Sigmamat \Gmat^T (\Gmat \Sigmamat \Gmat^T )^{-1} \rv +\left[ \Imat-\Sigmamat \Gmat^T ( \Gmat \Sigmamat \Gmat^T )^{-1}\Gmat\right]\yv $. For example, if $\muv = (1,1.2)^T$, $\Sigmamat = [(1,0.3)^T,(0.3,1)^T]$, $\Gmat = (1,1)$, and $\rv = 1$, then $\yv = (1,2)^T$ would be mapped to $\xv=(0,1)^T$ on a straight line using Algorithm \ref{alg:2}. } \label{fig:DemoAlg2} \end{figure} \subsection{Fast simulation of MVN distributions with structured covariance or precision matrices} For fast simulation of MVN distributions with structured covariance or precision matrices, our idea is to relate them to higher-dimensional hyperplane-truncated MVN distributions, with block-diagonal covariance matrices, that can be efficiently simulated with Algorithm \ref{alg:2}. We first introduce an efficient algorithm for the simulation of a MVN distribution, whose covariance matrix is a positive-definite matrix subtracted by a low-rank symmetric matrix. Such kind of covariance matrices commonly arise in the conditional distributions of MVN distributions, as shown in~\eqref{eq:conditional}. We then further extend this algorithm to the simulation of a MVN distribution whose precision (inverse covariance) matrix is the sum of a positive-definite matrix and a low-rank symmetric matrix. Such kind of precision matrices commonly arise in the conditional posterior distributions of the regression coefficients in both linear regression and generalized linear models. \begin{thm}\label{thm_condition} The probability density function (PDF) of the MVN distribution \begin{equation}\label{eq:px1} \xv_1\sim \Nor ( \muv_1,~ \Sigmamat_{11} - \Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}), \end{equation} is the same as the PDF of the marginal distribution of $\xv_1=(x_1,\ldots,x_{k_1})^T$ in $\xv=(\xv_1^T,x_{k_1+1},\ldots,x_{k})^T$, whose PDF is expressed as \begin{align} p( \xv\,|\, \muv,\tilde \Sigmamat,\Gmat,\rv ) &= \Nor_{\{ \xv: \Gmat \xv = \rv\}}\left( \muv, \tilde \Sigmamat \right) \notag\\ &=\frac{1}{Z} \exp\left[-\frac{1}{2}(\xv- \muv)^T\tilde \Sigmamat^{-1} (\xv- \muv) \right] \delta(\Gmat\xv = \rv), \end{align} where $Z$ is a normalization constant, $\Gmat_1 = \Sigmamat_{21}\Sigmamat_{11}^{-1} $ is a matrix of size $k_2\times k_1$, $\Gmat_2 $ is a user-specified full rank invertible matrix of size $k_2\times k_2$, $\rv\in\mathbb{R}^{k_2}$ is a user-specified vector, and \begin{equation}\label{eq:14} \Gmat=(\Gmat_1,\Gmat_2) \in\mathbb{R}^{k_2\times k},~~~~ \muv = \begin{bmatrix} \muv_1\\ \muv_2 \end{bmatrix}\in\mathbb{R}^{k}, ~~~~\tilde \Sigmamat = \begin{bmatrix} \Sigmamat_{11} & \mathbf{0} \\ \mathbf{0} & \tilde \Sigmamat_{22} \end{bmatrix}\in\mathbb{R}^{k\times k}, \end{equation} where \begin{align}\label{eq:tilde_mu2} & \muv_2= \Gmat_{2}^{-1}(\rv - \Sigmamat_{21}\Sigmamat_{11}^{-1} \muv_1),\\%\in \mathbb{R}^{ k_2},\\ \label{eq:tilde_sig22} &\tilde \Sigmamat_{22} = \Gmat_2^{-1}\left(\Sigmamat_{22} - \Sigmamat_{21} \Sigmamat_{11}^{-1} \Sigmamat_{12}\right)(\Gmat_{2}^{-1})^T. \end{align} \end{thm} The above theorem shows how the simulation of a MVN distribution, whose covariance matrix is a positive-definite matrix minus a symmetric matrix, can be realized by the simulation of a higher-dimensional hyperplane-truncated MVN distribution. By construction, it makes the covariance matrix $\tilde \Sigmamat$ of the truncated-MVN be block diagonal, but still preserves the flexibility to customize the full-rank matrix $\Gmat_2$ and the vector~$\rv$. While there are infinitely many choices for both $\Gmat_2$ and $\rv$, in the following discussion, we remove that flexibility by specifying $\Gmat_2 = \Imat_{k_2} $, leading to $\Gmat=(\Gmat_1,\Gmat_2)=(\Sigmamat_{21}\Sigmamat_{11}^{-1} ,\Imat_{k_2})$, and $\rv = \Sigmamat_{21}\Sigmamat_{11}^{-1} \muv_1$. This specific setting of $\Gmat_2$ and $\rv$ leads to the following Corollary that is a special case of Theorem~\ref{thm_condition}. Note that while we choose this specific setting in the paper, depending on the problems under study, other settings may lead to even more efficient simulation algorithms. \begin{cor}\label{cor1} The PDF of the MVN distribution \begin{equation} \xv_1 \sim \Nor ( \muv_1, \Sigmamat_{11} - \Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}) \end{equation} is the same as the PDF of the marginal distribution of $\xv_1$ in $\xv=(\xv_1^T,x_{k_1+1},\ldots,x_{k})^T$, whose PDF is expressed as \begin{align} p( \xv) &= \Nor_{\{ \xv: \, \Sigmamat_{21}\Sigmamat_{11}^{-1}\xv_1 + \xv_2 = \Sigmamat_{21}\Sigmamat_{11}^{-1} \muv_1\}}\left( \muv, \tilde \Sigmamat \right) \notag\\ &=\frac{1}{Z} \exp\left[-\frac{1}{2}(\xv- \muv)^T\tilde \Sigmamat^{-1} (\xv- \muv) \right] \delta( \Sigmamat_{21}\Sigmamat_{11}^{-1}\xv_1 + \xv_2 = \Sigmamat_{21}\Sigmamat_{11}^{-1} \muv_1), \end{align} where $\xv_2 = (x_{k_1+1},\ldots,x_{k})^T$, $Z$ is a normalization constant, and \begin{equation} \muv = \begin{bmatrix} \muv_1\\ \mathbf{0} \end{bmatrix}\in\mathbb{R}^{k}, ~~~~\tilde \Sigmamat = \begin{bmatrix} \Sigmamat_{11} & \mathbf{0} \\ \mathbf{0} & \Sigmamat_{22} - \Sigmamat_{21} \Sigmamat_{11}^{-1} \Sigmamat_{12} \end{bmatrix}\in\mathbb{R}^{k\times k}. \end{equation} \end{cor} Further applying Theorem \ref{maintheorem1} to Corollary \ref{cor1}, as described in detail in the Appendix, a MVN random variable $\xv$ with a structured covariance matrix can be generated as in Algorithm \ref{alg:3}, where there is no need to compute $\Sigmamat_{11}-\Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}$ and its Cholesky decomposition. Suppose the covariance matrix $\Sigmamat_{11}$ admits some special structure that makes it easy to invert and computationally efficient to simulate from $\Nor (\mathbf{0},\Sigmamat_{11})$, then Algorithm \ref{alg:3} could lead to a significant saving in computation if $k_2\ll k_1$. On the other hand, when $k_2\gg k_1$ and $\Sigmamat_{22}- \Sigmamat_{21} \Sigmamat_{11}^{-1} \Sigmamat_{12}$ admits no special structures, Algorithm \ref{alg:3} may not bring any computational advantage and hence one may resort to the naive Cholesky decomposition based procedure. Detailed computational complexity analyses for both methods are provided in Tables \ref{tab:CompNaive3} and \ref{tab:CompAlg3} of the Appendix, respectively. \vspace{2mm} \begin{algorithm}[H] \begin{itemize} \item Sample $\yv_1\sim\Nor (\mathbf{0},\Sigmamat_{11})$ and $\yv_2\sim\Nor (\mathbf{0},\Sigmamat_{22}- \Sigmamat_{21} \Sigmamat_{11}^{-1} \Sigmamat_{12})$ ; \item Return $\xv_1 = \muv_1+\yv_1 - \Sigmamat_{12} \Sigmamat_{22}^{-1} (\Sigmamat_{21} \Sigmamat_{11}^{-1}\yv_1 + \yv_2)$, which can be realized using \begin{itemize} \item Solve $\alphav$ such that $ \Sigmamat_{22}\alphav = \Sigmamat_{21} \Sigmamat_{11}^{-1}\yv_1 + \yv_2$; \item Return $\xv_1 = \muv_1+\yv_1 - \Sigmamat_{12} \alphav$. \end{itemize} \end{itemize} \caption{\label{alg:3} Simulation of the MVN distribution $$\xv_1\sim\Nor(\muv_1,\Sigmamat_{11}-\Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}).$$ } \end{algorithm} \vspace{2mm} \begin{cor} \label{cor2} A random variable simulated with Algorithm \ref{alg:3} is distributed as $ \xv_1\sim\Nor (\muv_1, \Sigmamat_{11} - \Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21}). $ \end{cor} The efficient simulation algorithm for a MVN distribution with a structured covariance matrix can also be further extended to a MVN distribution with a structured precision matrix, as described below, where $\betav \in \Rbb^p$, $\muv_{\beta} \in \Rbb^{p}$, $\Phimat \in \Rbb^{n \times p}$, and both $\Amat \in \Rbb^{p \times p}$ and $\Omegamat \in \Rbb^{n \times n}$ are positive-definite matrices. Computational complexity analyses for both the naive Cholesky decomposition based implementation and Algorithm \ref{alg:4} are provided in Table \ref{tab:CompNaive4} and \ref{tab:CompAlg4} of the Appendix, respectively. Similar to Algorithm~\ref{alg:3}, Algorithm \ref{alg:4} may bring a significant saving in computation when $p \gg n$ and $\Amat$ admits some special structure that makes it easy to invert and computationally efficient to simulate $\yv_1$. \vspace{2mm} \begin{algorithm}[H] \begin{itemize} \item Sample $\yv_1\sim\Nor (\mathbf{0},\Amat^{-1})$ and $\yv_2\sim\Nor (\mathbf{0},\Omegamat^{-1})$ ; \item Return $ \betav =\muv_{\beta}+\yv_1 - \Amat^{-1} \Phimat^T (\Omegamat^{-1}+\Phimat\Amat^{-1}\Phimat^T)^{-1} \left( \Phimat \yv_1+\yv_2\right)$, which can be realized using \begin{itemize} \item Solve $\alphav$ such that $ (\Omegamat^{-1}+\Phimat\Amat^{-1}\Phimat^T)\alphav = \Phimat \yv_1+\yv_2$. \item Return $\betav =\muv_{\beta}+\yv_1 - \Amat^{-1} \Phimat^T \alphav$. \end{itemize} \end{itemize} \caption{\label{alg:4} Simulation of the MVN distribution $$\betav\sim\Nor\left[\muv_{\beta}, (\Amat + \Phimat^T \Omegamat \Phimat)^{-1}\right].$$ } \end{algorithm} \vspace{2mm} \begin{cor}\label{cor3} The random variable obtained with Algorithm \ref{alg:4} is distributed as $\betav\sim\Nor(\muv_{\beta},\Sigmamat_{\beta})$, where $ \Sigmamat_{\beta} = (\Amat + \Phimat^T \Omegamat \Phimat)^{-1}$. \end{cor} \section{Illustrations} Below we provide several examples to illustrate Theorem \ref{maintheorem1}, which shows how to efficiently simulate from a hyperplane-truncated MVN distribution, and Corollary \ref{cor2} (Corollary \ref{cor3}), which shows how to efficiently simulate from a MVN distribution with a structured covariance (precision) matrix. We run all our experiments on a 2.9 GHz computer. \subsection{Simulation of hyperplane-truncated MVNs} We first compare Algorithms \ref{alg:1} and \ref{alg:2}, whose generated random samples follow the same distribution, as suggested by Theorem \ref{maintheorem1}, to highlight the advantages of Algorithm~\ref{alg:2} over Algorithm \ref{alg:1}. We then employ Algorithm \ref{alg:2} for a real application whose data dimension is high and sample size is large. \subsubsection{Comparison of Algorithms \ref{alg:1} and \ref{alg:2}} \label{sec:A1vsA2} We compare Algorithms \ref{alg:1} and \ref{alg:2} in a wide variety of settings by varying the data dimension $k$, varying the number of hyperplane constraints $k_2$, and choosing either a diagonal covariance matrix $\Sigmamat$ or a non-diagonal one. We generate random diagonal covariance matrices using the MATLAB command $\diag(0.05+\text{rand}(k,1))$ and random non-diagonal ones using $U.'*\diag(0.05+\text{rand}(k,1))*U$, where $\text{rand}(k,1)$ is a vector of $k$ uniform random numbers and $U$ consists of a set of $k$ orthogonal basis vectors. The elements of $\muv$, $\rv$, and $\Gmat$ are all sampled from $\Nor(0,1)$, with the singular value decomposition applied to $\Gmat$ to check whether $\mbox{Rank}(\Gmat)=k_2$. First, to verify Theorem \ref{maintheorem1}, we conduct an experiment with $k=5000$ data dimension, $k_2=20$ hyperplanes, and a diagonal $\Sigmamat$. Contour plots of two randomly selected dimensions of the 10,000 random samples simulated with Algorithms \ref{alg:1} and \ref{alg:2} are shown in the top and bottom rows of Figure \ref{fig:A1vsA2CutPatch}, respectively. The clear matches between the contour plots of these two different algorithms suggest the correctness of Theorem \ref{maintheorem1}. \begin{figure}[t] \centering \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D515D4164A1.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D2249D4294A1.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D2253D834A1.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D3138D3871A1.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial2D550D1011A1.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D515D4164A2.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D2249D4294A2.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D2253D834A2.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial1D3138D3871A2.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{ZA1A2Trial2D550D1011A2.pdf} } \caption{\small Comparison of the contour plots of two randomly selected dimensions of the 10,000 $k=5000$ dimensional random samples simulated with Algorithm \ref{alg:1} (top row) and Algorithm \ref{alg:2} (bottom row). Each of the five columns corresponds to a random trial. } \label{fig:A1vsA2CutPatch} \end{figure} To demonstrate the efficiency of Algorithm \ref{alg:2}, we first carry out a series of experiments with the number of hyperplane constraints fixed at $k_2=20$ and the data dimension increased from $k=50$ to $k=5000$. The computation time of simulating 10,000 samples averaged over five random trials is shown in Figure \ref{fig:A1A2K220GenCov} for non-diagonal $\Sigmamat$'s and in Figure \ref{fig:A1A2K220DiagCov} for diagonal ones. It is clear that, when the data dimension $k$ is high, Algorithm \ref{alg:2} has a clear advantage over Algorithm \ref{alg:1} by avoiding computing unnecessary intermediate variables, which is especially evident when $\Sigmamat$ is diagonal. We then carry out a series of experiments where we vary not only $k$, but also $k_2$ from $0.1k$ to $0.9k$ for each~$k$. As shown in Figure \ref{fig:A1vsA2Time}, it is evident that Algorithm \ref{alg:2} dominates Algorithm \ref{alg:1} in all scenarios, which can be explained by the fact that Algorithm \ref{alg:2} needs to compute much fewer intermediate variables. Also observed is that a larger $k_2$ leads to slower simulation for both algorithms, but to a much lesser extent for Algorithm \ref{alg:2}. Moreover, the curvatures of those curves indicate that Algorithm \ref{alg:2} is more practical in a high dimensional setting. Note that since Algorithm \ref{alg:2} can naturally exploit the structure of the covariance matrix $\Sigmamat$ for fast simulation, it is clearly more capable of benefiting from having a diagonal or block-diagonal $\Sigmamat$, demonstrated by comparing Figures \ref{fig:Alg1GenCovVaryK2} and \ref{fig:Alg2GenCovVaryK2} with Figures \ref{fig:Alg1DiagCovVaryK2} and \ref{fig:Alg2DiagCovVaryK2}. All these observations agree with our computational complexity analyses for Algorithms \ref{alg:1} and \ref{alg:2}, as shown in Table \ref{tab:CompAlg1} and \ref{tab:CompAlg2} of the Appendix, respectively. \begin{figure}[t] \centering \subfigure[]{\label{fig:A1A2K220GenCov} \includegraphics[width = 4.1 cm]{A1A2K220GenCov.pdf} } \subfigure[]{\label{fig:Alg1GenCovVaryK2} \includegraphics[width = 4.1 cm]{Alg1GenCovVaryK2.pdf} } \subfigure[]{\label{fig:Alg2GenCovVaryK2} \includegraphics[width = 4.1 cm]{Alg2GenCovVaryK2.pdf} } \subfigure[]{\label{fig:A1A2K220DiagCov} \includegraphics[width = 4.1 cm]{A1A2K220DiagCov.pdf} } \subfigure[]{\label{fig:Alg1DiagCovVaryK2} \includegraphics[width = 4.1 cm]{Alg1DiagCovVaryK2.pdf} } \subfigure[]{\label{fig:Alg2DiagCovVaryK2} \includegraphics[width = 4.1 cm]{Alg2DiagCovVaryK2.pdf} } \caption{\small Average time of simulating 10,000 hyperplane-truncated MVN samples over five random trials in different dimensions with non-diagonal covariance matrixes (top row) and diagonal ones (bottom row). (a)(d) Comparison with fixed $k_2 = 20$. (b)(e) Algorithm \ref{alg:1} with varying $k_2$. (c)(f) Algorithm~\ref{alg:2} with varying $k_2$. } \label{fig:A1vsA2Time} \end{figure} \subsubsection{A practical application of Algorithm \ref{alg:2}} In what follows, we extend Algorithm \ref{alg:2} to facilitate simulation from a MVN distribution truncated on a probability simplex $\mathbb{S}^k = \{ \xv : \xv \in \Rbb^k, \bds 1^T \xv = 1, x_i \ge 0, i= 1,\cdots,k \} $. This problem frequently arises when unknown parameters can be interpreted as fractions or probabilities, for instance, in topic models \citep{blei2003latent}, admixture models \citep{pritchard2000inference,dobigeon2009bayesian,bazot2013unsupervised}, and discrete directed graphical models \citep{heckerman1998tutorial}. With Algorithm \ref{alg:2}, one may remove the equality constraint to greatly simplify the problem. More specifically, we focus on a big data setting in which the globally shared simplex-constrained model parameters could be linked to some latent counts via the multinomial likelihood. When there are tens of thousands or millions of observations in the dataset, scalable Bayesian inference for the simplex-constrained globally shared model parameters is highly desired, for example, for inferring the topics' distributions over words in latent Dirichlet allocation \citep{blei2003latent,OnlineLDA} and Poisson factor analysis \citep{zhou2012beta,GBN}. Let us denote the $\kappa$th model parameter vector constrained on a $V$-dimensional simplex by $\phiv_\kappa\in \mathbb{S}^V$, which could be linked to the latent counts $n_{vj\kappa}\in \Zbb$ of the $j$th document under a multinomial likelihood as $(n_{1j\kappa},\ldots, n_{Vj\kappa})\sim \mbox{Mult}(n_{\cdotv j\kappa}, \phiv_\kappa)$, where $\Zbb = \{ 0,1,2,\cdots\}$, $v\in\{1,\ldots,V\}$, $\kappa\in\{1,\ldots,K\}$, and $j\in\{1,\ldots,N\}$. In topic modeling, one may consider $K$ as the total number of latent topics and $n_{vj\kappa}$ as the number of words at the $v$th vocabulary term in the $j$th document that are associated with the $\kappa$th latent topic. Note that the dimension $V$ in real applications is often large, such as tens of thousands in topic modeling. Given the observed counts $n_{vj}$ for the whole dataset, in a batch-learning setting, one typically iteratively updates the latent counts $n_{vj\kappa}$ conditioning on $\phiv_\kappa$, and updates $\phiv_\kappa$ conditioning on $n_{vj\kappa}$. However, this batch-learning inference procedure would become inefficient and even impractical when the dataset size $N$ grows to a level that makes it too time consuming to finish even a single iteration of updating all local variables $n_{vj\kappa}$. To address this issue, we consider constructing a mini-batch based Bayesian inference procedure that could make substantial progress in posterior simulation while the batch-learning one may still be waiting to finish a single iteration. Without loss of generality, in the following discussion, we drop the latent factor/topic index $\kappa$ to simplify the notation, focusing on the update of a single simplex-constrained global parameter vector. More specifically, we let the latent local count vector $\nv_j =(n_{1j},\ldots,n_{Vj})^T$ be linked to the simplex-constrained global parameter vector $\phiv \in \mathbb{S}^V$ via the multinomial likelihood as $\nv_j \sim \Mult \left( n_{\cdotv j}, \phiv \right)$, and impose a Dirichlet distribution prior on $\phiv$ as $\phiv \sim \Dir \left( \eta \mathbf{1}_V\right)$. Instead of waiting for all $\nv_j$ to be updated before performing a single update of $\phiv$, we develop a mini-batch based Bayesian inference algorithm under a general framework for constructing stochastic gradient Markov chain Monte Carlo (SG-MCMC) \citep{ma2015complete}, allowing $\phiv$ to be updated every time a mini-batch of $\nv_j$ are processed. For the sake of completeness, we concisely describe the derivation for a SG-MCMC algorithm, as outlined below, for simplex-constrained globally shared model parameters. We refer the readers to \citet{TLASGR} for more details on the derivation and its application to scalable inference for topic modeling. Using the reduced-mean parameterization of the simplex constrained vector $\phiv$, namely $\varphiv = (\phi_1,\cdots,\phi_{V-1})^T$, where $\varphiv \in \mathbb{R}_{+}^{V-1}$ is constrained with $\varphi_{\cdotv} \le 1$, we develop a SG-MCMC algorithm that updates $\varphiv$ for the $t$th mini-batch as \begin{equation} \label{eq:upvarphi} \varphiv_{t + 1} \!=\! \left[ \varphiv_t \!+\! \frac{\varepsilon _t}{M} \!\left[ \left( \rho \bar \nv_{: \cdotv} \!+\! \eta \right) \!-\! \left(\rho n_{\cdotv \cdotv} \!+\! \eta V\right) \varphiv_t \right] \!+\! \Nc \left( \bds 0,\frac{2\varepsilon _t}{M} \!\left[ \diag \left( \varphiv_t \right) \!-\! \varphiv_t \varphiv_t ^T \right] \right) \right]_{\triangle} \!, \end{equation} where $\varepsilon _t$ are annealed step sizes, $\rho$ is the ratio of the dataset size $N$ to the mini-batch size, \zz $\nv_{: \cdotv}=(\nv_{1 \cdotv},\cdots,\nv_{V \cdotv})^T = \sum_{j\in I_t} \nv_{j}$, $\bar \nv_{: \cdotv} = (\nv_{1 \cdotv},\cdots,\nv_{(V-1) \cdotv})^T$, $[\cdotv]_{\triangle}$ denotes the constraint that $\varphiv \in \mathbb{R}_{+}^{V-1}$ and $\varphi_{\cdotv} \le 1$, and $M := \Ebb \left[\sum_{j=1}^N n_{\cdotv j} \right]$ is approximated along the updating using $ M = \left( 1 - \varepsilon_t \right) M + {\varepsilon_t} \rho \Ebb \left[ n_{ \cdotv \cdotv } \right] $. Alternatively, we have an equivalent update equation for $\phiv$ as \begin{equation} \label{eq:upphi} \phiv_{t + 1} \!=\! \left[ \phiv_t \!+\! \frac{\varepsilon _t}{M} \!\left[ \left( \rho \nv_{: \cdotv} \!+\! \eta \right) \!-\! \left(\rho n_{\cdotv \cdotv} \!+\! \eta V\right) \phiv_t \right] \!+\! \Nc \left( \bds 0,\frac{2\varepsilon _t}{M} \diag \left( \phiv_t \right) \right) \right]_{\angle} \!, \end{equation} where $[\cdotv]_{\angle}$ represents the constraint that $\phiv \in \Rbb_{+}^{V}$ and $\bds 1 ^T \phiv = 1$. It is clear that \eqref{eq:upvarphi} corresponds to simulation of a $V-1$ dimensional truncated MVN distribution with $V$ inequality constraints. Since the number of constraints is larger than the dimension, previously proposed iterative simulation methods such as the one in \citet{botev2016normal} are often inappropriate. Note that, by omitting the non-negative constraints, the update in \eqref{eq:upphi} corresponds to simulation of a hyperplane-truncated MVN simulation with a diagonal covariance matrix, which can be efficiently sampled as described in the following example. \noindent\textbf{Example 1:} \textit{ Simulation of a hyperplane-truncated MVN distribution as $$ \xv\sim\Nor_{\mathcal{S}}[\muv,a\,\diag(\phiv)],~~\mathcal{S}=\left\{\xv: \mathbf{1}^T \xv = 1\right\}, $$ where $\xv\in\mathbb{R}^{k}$, $\muv\in\mathbb{R}^{k}$, $\mathbf{1}^T\xv =\sum_{i=1}^k x_i$, $\phiv\in\mathbb{R}^{k}$, $a>0$, $\phi_i > 0$ for $i \in\{ 1,\cdots,k\}$, and $\mathbf{1}^T \phiv = \sum_{i=1}^k \phi_i=1$, can be realized as follows. \begin{itemize} \item Sample $\yv\sim\Nor [ \muv , a \diag(\phiv) ]$; \item Return $\xv = \yv + ( 1 - \mathbf{1}^T \yv) \phiv$. \end{itemize} } \noindent The sampling steps in Example 1 directly follow Algorithm \ref{alg:2} and Theorem~\ref{maintheorem1} with the distribution parameters specified as $\Sigmamat = a \diag (\phiv)$, $\Gmat = \bds 1^T$, and $\rv = 1$. Accordingly, we present the following fast sampling procedure for \eqref{eq:upvarphi}. \noindent\textbf{Example 2:} \textit{ Simulation from \eqref{eq:upvarphi} can be approximately but rapidly realized as \begin{itemize} \item Sample $\yv \sim \Nor \big[ \phiv_t + \frac{\varepsilon _t}{M} \left[ \left( \rho \nv_{: \cdotv} + \eta \right) - \left(\rho n_{\cdotv \cdotv} + \eta V\right) \phiv_t \right] ,\frac{2\varepsilon _t}{M} \diag \left( \phiv_t \right) \big]$; \item Calculate $\zv = \yv + ( 1 - \mathbf{1}^T \yv) \phiv_t$; \item If $\zv \in \mathbb{S}$, return $\varphiv_{t+1} = (z_1,\cdots,z_{V-1})^T$; else calculate $\dv = \max(\epsilon,\zv)$ with a small constant $\epsilon \ge 0$, let $\ev = \dv / \sum\nolimits_{i=1}^V d_i$, and return $\varphiv_{t+1} = (e_1,\cdots,e_{V-1})^T$. \end{itemize} } To verify Example 2, we conduct an experiment using multinomial-distributed data vectors of $V=2000$ dimensions, which are generated as follows: considering that the simplex-constrained vector $\phiv$ is usually sparse in a high-dimensional application, we sample a $V=2000$ dimensional vector $\fv$ whose elements are uniformly distributed between 0 and 1, randomly select 40 dimensions and reset their values to be 100, and set $\phiv = \fv / \sum\nolimits_{i=1}^V f_i$; we simulate $N=10,000$ samples, each $\nv_j$ of which is generated from the multinomial distribution $\mbox{Mult}(n_{\cdotv j}, \phiv)$, where the number of trials is random and generated as $n_{\cdotv j} \sim \Pois (50)$. We set $\varepsilon_t = t^{-0.99}$ and use mini-batches, each of which consists of 10 data samples, to stochastically update global parameters via SG-MCMC. For comparison, we choose the same SG-MCMC inference procedure but consider simulating \eqref{eq:upvarphi}, as performed every time a mini-batch of data samples are provided, either as in Example 2 or with the Gibbs sampler of \citet{rodriguez2004efficient}. Simulating \eqref{eq:upvarphi} with the Gibbs sampler of \citet{rodriguez2004efficient} is realized by updating all the $V$ dimensions, one dimension at a time, in each Gibbs sampling iteration. We set the total number of Gibbs sampling iterations for \eqref{eq:upvarphi} in each mini-batch based update as 1, 5, or 10. Note that in practice, the $\nv_j$ belonging to the current mini-batch are often latent and are updated conditioning on the data samples in the mini-batch and $\phiv$. For simplicity, all $\nv_j$ here are simulated once and then fixed. \begin{figure}[!t] \centering \subfigure[]{\label{fig:ResErrIterationComp} \includegraphics[height = 4.6 cm]{ResErrIteration10.pdf} } \subfigure[]{\label{fig:ResErrTimeComp} \includegraphics[height = 4.6 cm]{ResErrTime10.pdf} } \caption{\small Comparisons of the residual errors of the simplex-constrained parameter vector, estimated under various settings of the stochastic-gradient MCMC (SG-MCMC) algorithm, as a function of (a) the number of processed mini-batches and (b) time. The curves labeled as ``Batch posterior mean'', ``SG-MCMC-fast'', and ``SG-MCMC-Gibbs'' correspond to the batch posterior mean, SG-MCMC with \eqref{eq:upvarphi} simulated as in Example 2, and SG-MCMC with \eqref{eq:upvarphi} simulated with the Gibbs sampler of \citet{rodriguez2004efficient}, respectively. The digit following ``SG-MCMC-Gibbs'' represents the number of Gibbs sampling iterations to simulate \eqref{eq:upvarphi} for each mini-batch. } \label{fig:ResErrComp} \end{figure} Using $\phiv_{post}^* = ( \sum_{j=1}^N \nv_{j} + \eta ) / ( \sum_{j=1}^N n_{\cdotv j} + \eta V ) $, the posterior mean of $\phiv$ in a batch-learning setting, as the reference, we show in Figure \ref{fig:ResErrComp} how the residual errors for the estimated $\phiv^*$, defined as $\left\| \phiv^* - \phiv \right\|_2 $, change both as a function of the number of processed mini-batches and as a function of computation time under various settings of the mini-batch based SG-MCMC algorithm. The curves shown in Figure \ref{fig:ResErrComp} suggest that for each mini-batch, to simulate \eqref{eq:upvarphi} with the Gibbs sampler of \citet{rodriguez2004efficient}, it is necessary to have more than one Gibbs sampling iteration to achieve satisfactory results. It is clear from Figure \ref{fig:ResErrIterationComp} that the Gibbs sampler with 5 or 10 iterations for each mini-batch, even though each mini-batch has only 10 data samples, provides residual errors that quickly approach that of the batch posterior mean with a tiny gap, indicating the effectiveness of the SG-MCMC updating in \eqref{eq:upvarphi}. While simulating \eqref{eq:upvarphi} with Gibbs sampling could in theory lead to unbiased samples if the number of Gibbs sampling iterations is large enough, it is much more efficient to simulate \eqref{eq:upvarphi} with the procedure described in Example 2, which provides a performance that is undistinguishable from those of the Gibbs sampler with as many as 5 or 10 iterations for each mini-batch, but at the expense of a tiny fraction of a single Gibbs sampling iteration. \subsection{Simulation of MVNs with structured covariance matrices} To illustrate Corollary \ref{cor2}, we mimic the truncated MVN simulation in \eqref{eq:upvarphi} and present the following simulation example with a structured covariance matrix. \noindent\textbf{Example 3:} Simulation of a MVN distribution as $$ \xv_1\sim\Nor [\muv_1,a\,\diag(\phiv_1)-a\,\phiv_1\phiv_1^T], $$ where $\xv_1\in\mathbb{R}^{k-1}$, $\muv_1\in\mathbb{R}^{k-1}$, $a>0$, $\phiv_1=(\phi_1,\ldots,\phi_{k-1})^T$, $\phi_{i}>0$ for $i\in\{1,\ldots,k-1\}$, and $\sum_{i=1}^{k-1}\phi_i <1$, can be realized as follows. \begin{itemize} \item Sample $\yv_1\sim\Nor [ \mathbf{0} , a\,\diag(\phiv_1) ]$ and $\yv_2\sim\Nor ( 0 , a^{-1}\,\phi_k )$, where $\phi_k=1- \sum_{i=1}^{k-1} \phi_i$; \item Return $\xv_1 = \muv_1+\yv_1 - (\mathbf{1}^T \yv_1 + a\yv_2) \phiv_1$. \end{itemize} Denoting $\xv=(\xv_1^T,x_{k})^T$, $\phiv=(\phiv_1^T,\phi_{k})^T$, $\muv=(\muv_1^T,\mu_{k})^T$, and $\mu_k=1-\mathbf{1}^T \muv_1$, the above sampling steps can also be equivalently expressed as follows. \begin{itemize} \item Sample $\yv\sim\Nor [ \muv , a\,\diag(\phiv) ]$; \item Return $\xv_1 = \yv_1 + (1- \mathbf{1}^T \yv ) \phiv_1$. \end{itemize} Directly following Algorithm \ref{alg:3} and Corollary \ref{cor2}, the first sampling approach for the above example can be derived by specifying the distribution parameters as $\Sigmamat_{11} = a \diag (\phiv_1)$, $\Sigmamat_{12} = \phiv_1$, $\Sigmamat_{21} = \phiv_1^T$, and $\Sigmamat_{22} = a^{-1}$, while the second approach can be derived by specifying $\Sigmamat_{11} = a \diag (\phiv_1)$, $\Sigmamat_{12} = a\phiv_1$, $\Sigmamat_{21} = a\phiv_1^T$, and $\Sigmamat_{22} = a$. To illustrate the efficiency of the proposed algorithms in Example 3, we simulate from the MVN distribution $\xv_1\sim\Nor [\muv_1,a\,\diag(\phiv_1)-a\,\phiv_1\phiv_1^T]$ using both a naive implementation via Cholesky decomposition of the covariance matrix and the fast simulation algorithm for a hyperplane-truncated MVN random variable described in Example 3. We set the dimension from $k = 10^2$ up to $k=10^4$ and set $\muv = (1/ k,\ldots,1/k)$ and $a = 0.5$. For each $k$ and each simulation algorithm, we perform 100 independent random trials, in each of which $\phiv$ is sampled from the Dirichlet distribution $\Dir ( 1,\ldots,1)$ and 10,000 independent random samples are simulated using that same $\phiv$. \begin{figure}[!t] \centering \includegraphics[width = 6.3 cm]{TimeDimension.pdf} \caption{\small Comparison of the naive Cholesky decomposition based implementation and Algorithm \ref{alg:3} in terms of the average time of generating 10,000 $k$-dimensional random samples from $ \xv_1\sim\Nor [\muv_1,a\,\diag(\phiv_1)-a\,\phiv_1\phiv_1^T] $. The distribution parameters are randomly generated and computation time averaged over 100 random trials is displayed. } \label{fig:TimeVsDim} \end{figure} As shown in Figure \ref{fig:TimeVsDim}, for the proposed Algorithm \ref{alg:3}, the average time of simulating 10,000 random samples increases linearly in the dimension $k$. By contrast, for the naive Cholesky decomposition based simulation algorithm, whose computational complexity is $O(k^3)$ \citep{golub2012matrix}, the average simulation time increases at a significantly faster rate as the dimension $k$ increases. \begin{figure}[t] \centering \subfigure[]{ \includegraphics[width = 2.2cm]{NaiveContour5.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{NaiveContour6.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{NaiveContour9.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{NaiveContour15.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{NaiveContour49.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{MargCutContour5.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{MargCutContour6.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{MargCutContour9.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{MargCutContour15.pdf} } \subfigure[]{ \includegraphics[width = 2.2 cm]{MargCutContour49.pdf} } \caption{\small Comparison of the contour plots of two randomly selected dimensions of the 10,000 $k=10^4$ dimensional random samples simulated with the naive Cholesky implementation (top row) and Algorithm \ref{alg:3} (bottom row). Each of the five columns corresponds to a random trial. } \label{fig:ContourCompare} \end{figure} For explicit verification, with the 10,000 simulated $k=10^4$ dimensional random samples in a random trial, we randomly choose two dimensions and display their joint distribution using a contour plot. As in Figure \ref{fig:ContourCompare}, shown in the first row are the contour plots of five different random trials for the naive Cholesky implementation, whereas shown in the second row are the corresponding ones for the proposed Algorithm \ref{alg:3}. As expected, the contour lines of the two figures in the same column closely match each other. \begin{figure}[!th] \centering \includegraphics[width = 8 cm]{A3FixK1VaryK2DiagCov.pdf} \caption{\small Comparison of the naive Cholesky decomposition based implementation and Algorithm \ref{alg:3} in terms of the average time of generating one $k_1=4000$ dimensional sample from $\xv_1\sim\Nor(\muv_1,\Sigmamat_{11}-\Sigmamat_{12}\Sigmamat_{22}^{-1}\Sigmamat_{21})$, with diagonal $\Sigmamat_{11}$ and $\Sigmamat_{22}$. The distribution parameters are randomly generated and computation time averaged over 50 random trials is displayed. } \label{fig:A3} \end{figure} To further examine when to apply Algorithm \ref{alg:3} instead of the naive Cholesky decomposition based implementation in a general setting, we present the computational complexity analyses in Tables \ref{tab:CompNaive3} and \ref{tab:CompAlg3} of the Appendix for the naive approach and Algorithm \ref{alg:3}, respectively. In addition, we mimic the settings in Section \ref{sec:A1vsA2} to conduct a set of experiments with randomly generated $\Sigmamat_{12}$, diagonal $\Sigmamat_{11}$, and diagonal $\Sigmamat_{22}$. We fix $k_1 = 4000$ and vary $k_2$ from 1 to 8000. The computation time for one sample averaged over 50 random trials is presented in Figure \ref{fig:A3}. It is clear from Tables \ref{tab:CompNaive3} and \ref{tab:CompAlg3} and Figure \ref{fig:A3} that, as a general guideline, one may choose Algorithm \ref{alg:3} when $k_2$ is smaller than $ k_1$ and $\Sigmamat_{11}$ admits some special structure that makes it easy to invert and computationally efficient to simulate from $\Nor(\bf{0},\Sigmamat_{11})$. \subsection{Simulation of MVNs with structured precision matrices} To examine when to apply Algorithm \ref{alg:4} instead of the naive Choleskey decomposition based procedure, we first consider a series of random simulations in which the sample size $n$ is fixed while the data dimension $p$ is varying. We then show that Algorithm \ref{alg:4} can be applied for high-dimensional regression whose $p$ is often much larger than $n$. \begin{figure}[!th] \centering \includegraphics[width = 8 cm]{A4FixNVaryPDiagCov.pdf} \caption{\small Comparison of the naive Cholesky decomposition based implementation and Algorithm \ref{alg:4} in terms of the average time of generating one $p$ dimensional sample from $\betav\sim\Nor\left[\muv_{\beta}, (\Amat + \Phimat^T \Omegamat \Phimat)^{-1}\right]$, with diagonal $\Amat$ and $\Omegamat$. The distribution parameters are randomly generated and computation time averaged over 50 random trials is displayed. } \label{fig:A4} \end{figure} We fix $n = 4000$, vary $p$ from 1 to 8000, and mimic the settings in Section \ref{sec:A1vsA2} to randomly generate $\Phimat$, diagonal $\Amat$, and diagonal $\Omegamat$. As a function of dimensions $p$, the computation time for one sample averaged over 50 random trials is shown in Figure \ref{fig:A4}. It is evident that, identical to the complexity analysis in Tables \ref{tab:CompNaive4} and \ref{tab:CompAlg4}, Algorithm \ref{alg:4} has a linear complexity with respect to $p$ under these settings, which will bring significant acceleration in a high-dimensional setting with $p \gg n$. If the sample size $n$ is large enough that $n>p$, then one may directly apply the naive Cholesky decomposition based implementation. Algorithm \ref{alg:4} could be slightly modified to be applied to high-dimensional regression, where the main objective is to efficiently sample from the conditional posterior of $\betav \in \Rbb^{p \times 1}$ in the linear regression model as \begin{equation} \tv \sim\Nor(\Phimat \betav,\Omegamat^{-1}),~\betav\sim\Nor(\mathbf{0},\Amat^{-1}), \end{equation} where $\Phimat\in\mathbb{R}^{n\times p}$, $\Omegamat \in \Rbb^{n \times n}$, and different constructions on $\Amat\in\mathbb{R}^{p\times p}$ lead to a wide variety of regression models \citep{caron2008sparse,carvalho2010horseshoe,polson2014bayesian}. The conditional posterior of $\betav$ is directly derived and shown in the following example, where its simulation algorithm is summarized by further generalizing Corollary~\ref{cor3}. \noindent\textbf{Example 4:} \textit{ Simulation of the MVN distribution $$\betav\sim\Nor\left[(\Amat + \Phimat^T \Omegamat \Phimat)^{-1}\Phimat^T\Omegamat\tv, ~~(\Amat + \Phimat^T \Omegamat \Phimat)^{-1}\right]$$ can be realized as follows. \begin{itemize} \item Sample $\yv_1\sim\Nor (\mathbf{0},\Amat^{-1})$ and $\yv_2\sim\Nor (\mathbf{0},\Omegamat^{-1})$ ; \item Return $ \betav =\yv_1 + \Amat^{-1} \Phimat^T (\Omegamat^{-1}+\Phimat\Amat^{-1}\Phimat^T)^{-1} \left(\tv - \Phimat \yv_1-\yv_2\right) $, which can be realized using \begin{itemize} \item Solve $\alphav$ such that $ (\Omegamat^{-1}+\Phimat\Amat^{-1}\Phimat^T)\alphav = \tv - \Phimat \yv_1-\yv_2$; \item Return $\betav =\yv_1 + \Amat^{-1} \Phimat^T \alphav$. \end{itemize} \end{itemize} } Note that if $\Omegamat = \Imat_n$, then the simulation algorithm in Example 4 reduces to the one in Proposition 2.1 of \citet{bhattacharya2015fast}, which is shown there to be significantly more efficient than that of \citet{rue2001fast} for high-dimensional regression if $p\gg n$. \section{Conclusions} A fast and exact simulation algorithm is developed for a multivariate normal (MVN) distribution whose sample space is constrained on the intersection of a set of hyperplanes, which is shown to be inherently related to the conditional distribution of a unconstrained MVN distribution. The proposed simulation algorithm is further generalized to efficiently simulate from a MVN distribution, whose covariance (precision) matrix can be decomposed as the sum (difference) of a positive-definite matrix and a low-rank symmetric matrix, using a higher dimensional hyperplane-truncated MVN distribution whose covariance matrix is block-diagonal. \bibliographystyle{ba}
763f81307dac160ed8accf91c4bc36a35227c37a
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction}\label{sec:intro} As we enter the era of precision LHC phenomenology, where NNLO QCD calculations are becoming the standard for many processes, the influence of electroweak corrections is increasingly relevant. A complete treatment of these inevitably requires the inclusion of diagrams with initial--state photons, with corresponding photon parton distribution function (PDF) introduced in analogy to the more commonly considered PDFs of the quarks and gluons~\cite{Gluck:2002fi,Martin:2004dh,Ball:2013hta,Martin:2014nqa,Schmidt:2015zda}. As discussed recently in~\cite{Pagani:2016caq,Bourilkov:2016qum,Accomando:2016tah,Mangano:2016jyj} the photon--initiated contribution may be significant for the production of lepton, $W$ boson and top quark pairs at higher invariant masses, and hence higher parton $x$. Such processes are of much phenomenological interest, being particularly sensitive to electroweak corrections and the PDFs, as well BSM physics; high mass lepton pair production, for example, is an irreducible background to the Drell--Yan production of a new $Z'$ boson. The issue of how to extract the photon PDF, and what the uncertainties associated with this are, is therefore crucial to any complete discussion of these processes. A range of approaches to this problem have been taken by the global PDF fitting groups: the first attempt in the MRST2004QED set~\cite{Martin:2004dh} fixed the functional form of the photon PDF by taking a simple model for photon emission from the valence quarks, while CT14QED~\cite{Schmidt:2015zda} generalized this to allow freedom in the overall normalization, which can then be extracted from data. An alternative approach is taken in NNPDF2.3QED~\cite{Ball:2013hta} (and more recently NNPDF3.0QED~\cite{Bertone:2016ume}), where instead the photon is parameterised freely, as in the case of the quarks and gluons, and fitted to a selection of inclusive data. In this case, the corresponding PDF uncertainties due to the quite unconstraining data considered in these fits are very large. However, these previous approaches in fact omit an important physical distinction between the photon and the quarks and gluons. The crucial difference is that QED, in contrast to QCD, corresponds to a long range force that does not suffer from the issue of non--perturbativity at low scales. Thus, a proton may coherently emit a photon ($p\to p\gamma$): such a process is experimentally extremely well measured, being governed by the well known electric and magnetic proton form factors for coherent photon emission, and is expected to constitute the dominant component of the input photon PDF. This is accounted for in the approach of~\cite{Martin:2014nqa,Harland-Lang:2016apc} (see also~\cite{Harland-Lang:2016lhw,Harland-Lang:2016qjy}), where it is shown that the photon PDF is then determined to a relatively high degree of accuracy. It is important to emphasise that the inclusion of this effect is not a theoretical assumption: indeed, exclusive lepton and $W$ boson pair production, due to precisely this initial--state coherent photon emission have been observed by both ATLAS~\cite{Aad:2015bwa,Aaboud:2016dkv} and CMS~\cite{Chatrchyan:2011ci,Chatrchyan:2013akv,Khachatryan:2016mud} at the LHC. Such processes contribute by definition to the corresponding inclusive observables. It is therefore important to consider the consequences of these physical considerations, and the approach which derives from it for describing the photon PDF; we will consider for concreteness in this paper the cases of high mass lepton and $W$ boson pair production. Here, the $\sim \alpha^2$ suppression in the initial--state $\gamma\gamma$ luminosity may be overcome by the enhancement of the $t$--channel photon--induced process at higher masses. However, for certain sets the most significant effect is due to the PDFs themselves: the NNPDF2.3QED~\cite{Ball:2013hta,Bertone:2016ume} set in particular predicts a sharper decrease in the quark (and gluon) densities compared to the central photon value, albeit within sizeable PDF uncertainties in the latter case. In such a situation, for both processes the photon--initiated contribution is found to be potentially sizeable and even dominant at high invariant mass~\cite{Bourilkov:2016qum,Accomando:2016tah,Mangano:2016jyj}. The case of $t\overline{t}$ production has also recently been discussed in~\cite{Pagani:2016caq}, where it is again shown that the NNPDF2.3QED set is consistent with a sizeable photon--initiated contribution at larger invariance masses, as well as forward rapidities. Although for the sake of brevity we will not deal with this explicitly here, our results can be readily extended to such a process. Given these findings, it is natural to consider what the prediction is for these processes within the approach of~\cite{Martin:2014nqa,Harland-Lang:2016apc} (see also~\cite{Gluck:2002fi}). In other words, what are the consequences of this dominantly coherent photon input PDF for the size and uncertainties of the photon PDF at higher $x$, and what are the implications for these high mass production processes? This is the question we consider in this paper: we will show that this approach in fact predicts that the photon PDF is well constrained by the dominantly coherent input requirement, with at higher $x$ any unknown incoherent contribution, already expected to be small, being further kinematically suppressed. We will demonstrate that this follows from the relatively simple nature of the photon DGLAP evolution, which due to the small size of the coupling $\alpha$ may to very good approximation be solved exactly~\cite{Harland-Lang:2016apc,Harland-Lang:2016lhw}. Within this approach, we will find that the decrease in the $\gamma\gamma$ luminosity is qualitatively similar to the quark and gluon cases. We will also discuss how the recent ATLAS measurement of high mass lepton pair production~\cite{Aad:2016zzw} and the corresponding extraction of the photon PDF presented in this analysis lends qualitative support to our results. The production of lepton and $W$ boson pairs are processes of much phenomenological interest at both the LHC and at a Future Circular Collider (FCC). With this in mind, we will consider cross section predictions for both of these processes at 13 and 100 TeV. In the case of lepton pair production we will show that the DY contribution is expected to be dominant out to very high masses, with a relatively small contribution from the photon---initiated process ($\sim 10$\% for our choice of cuts). For $W$ pair production, on the other hand, the relative contribution from the photon--initiated process is more significant, and at higher masses it is comparable in size to the standard QCD--initiated process. In this case a very careful consideration of the uncertainties associated with the photon PDF is therefore essential. Finally, during the last stages of preparing this manuscript the paper~\cite{Manohar:2016nzj} appeared. While the overall approach to treating the photon PDF and the details of the analysis are quite distinct, this work also includes the dominantly coherent input component. We may therefore expect the general conclusions to be consistent with our findings. Although we will not provide a detailed investigation of this question here, we present a brief comparison to the results of this approach. As we shall see, the predicted photon PDF does indeed quite closely coincide with our results. Thus, we will expect comparable results to hold for the cases of high mass lepton and $W$ boson pair production when using the \texttt{LUXqed} set. The outline of this paper is as follows. In Section~\ref{sec:photongen} we describe our approach to modelling the photon PDF, demonstrating in Section~\ref{sec:photondglap} how the DGLAP equation for the photon PDF may be solved, and then discussing in Section~\ref{sec:input} how the input photon PDF may be described. In Section~\ref{sec:photonhighx} we discuss the implications for the photon, in particular at higher $x$, and compare our results with the other available PDFs, concentrating on the NNPDFQED3.0 set. In Section~\ref{sec:dy} we present predictions for high mass lepton and $W$ boson pair production at the LHC and $\sqrt{s}=100$ TeV FCC. Finally, in Section~\ref{sec:conc} we conclude. \section{The photon PDF: general considerations}\label{sec:photongen} \subsection{Solving the DGLAP equation}\label{sec:photondglap} The starting point for any discussion of the photon PDF is the corresponding DGLAP evolution equation for the distribution $\gamma(x,Q^2)$. At LO in $\alpha$ and $\alpha_S$ this is given by\footnote{In this section we work for simplicity at LO in $\alpha_S$, but this discussion can readily be generalised to NLO, as in~\cite{Harland-Lang:2016qjy}, using the results of~\cite{deFlorian:2015ujt} for the corresponding splitting functions; in the following sections we use the full NLO result.} \begin{equation}\label{dglap} \frac{\partial \gamma(x,Q^2)}{\partial \ln Q^2}=\frac{\alpha(Q^2)}{2\pi}\!\int_x^1\!\frac{dz}z\! \left(P_{\gamma\gamma}(z)\gamma(\frac xz,Q^2) +\sum_q e^2_qP_{\gamma q}(z)q(\frac xz,Q^2)\right)\;. \end{equation} Here $P_{\gamma q}(z)$ is the $q\to \gamma$ splitting function, and $P_{\gamma\gamma}$ corresponds to the virtual self--energy correction to the photon propagator, given by \begin{equation}\label{pgg} P_{\gamma\gamma}(z)=-\frac{2}{3}\left[N_c\sum_q e^2_q +\sum_l e^2_l\right]\delta(1-z)\;, \end{equation} where $q$ and $l$ denote the active quark and lepton flavours in the fermion loop. As the virtual correction (\ref{pgg}) is proportional to an overall delta function the corresponding contribution to (\ref{dglap}) is proportional to the photon PDF evaluated at $x$. Therefore, if we ignore the small effect that the photon PDF has on the evolution of the quark and gluons (as discussed in~\cite{Harland-Lang:2016apc}, these generally give less than a 0.1\% correction to the photon), which enter at higher orders in $\alpha$, then (\ref{dglap}) can be solved exactly, giving~\cite{Harland-Lang:2016apc,Harland-Lang:2016lhw} \begin{align} \nonumber \gamma(x,\mu_F^2)&=\gamma(x,Q_0^2)\,S_{\gamma}(Q_0^2,\mu_F^2)+\int_{Q_0^2}^{\mu_F^2}\frac{\alpha(Q^2)}{2\pi}\frac{dQ^2} {Q^2}\int_x^1\frac{dz}z \;\sum_q e^2_qP_{\gamma q}(z)q(\frac xz,Q^2)\,S_{\gamma}(Q^2,\mu_F^2)\;,\\ \label{gampdf} &\equiv \gamma^{{\rm in}}(x,\mu^2)+\gamma^{\rm evol}(x,\mu^2)\;, \end{align} where $\gamma(x,Q_0^2)$ is the input PDF at the scale $Q_0$, and we have introduced the photon Sudakov factor \begin{equation}\label{sudgam} S_{\gamma}(Q_0^2,\mu_F^2)=\exp\left(-\frac{1}{2}\int_{Q_0^2}^{\mu_F^2}\frac{{\rm d}Q^2}{Q^2}\frac{\alpha(Q^2)}{2\pi}\int_0^1 {\rm d} z\sum_{a}\,P_{a\gamma}(z)\right)\;. \end{equation} Here $P_{q(l)\gamma}(z)$ is the $\gamma$ to quark (lepton) splitting function, given by \begin{equation} P_{a\gamma}(z)=N_a\left[z^2+(1-z)^2\right]\;, \end{equation} where $N_a=N_c e_q^2$ for quarks and $N_a=e_l^2$ for leptons, while the factor of $1/2$ in (\ref{sudgam}) is present to avoid double counting over the quark/anti--quarks (lepton/anti--leptons). Written in this form, the physical interpretation of the Sudakov factor is clear: it represents the Poissonian probability for no parton emission from the photon during its evolution from the low scale $Q_0$ up to the hard scale $\mu_F$. Thus, the photon PDF (\ref{gampdf}) at $\mu_F$ can be written as the sum of a contribution from low--scale emission of a photon, with no further branching, and a term due to higher scale DGLAP emission from quarks\footnote{This separation was used in~\cite{Harland-Lang:2016apc} to demonstrate how a rapidity gap veto can be included in photon--initiated processes.}. For the purposes of the discussion in this paper, the crucial point is that when considering the photon PDF and its corresponding uncertainty at some given $x$ and $\mu_F^2$ value, the contributions to this from the input photon distribution, at the starting scale $Q_0$, and from the DGLAP evolution term due to high scale emission from the quarks, are completely separated; this will greatly simplify the discussion which follows, and allow some fairly simple and robust conclusions to be drawn. \subsection{The input distribution}\label{sec:input} The photon PDF has been separated in (\ref{gampdf}) into an input component at $Q_0$ and an evolution component, due to high scale $q\to q\gamma$ emission. While the latter quantity is given in terms of the generally well determined quark PDFs, the former quantity is on the face of it completely unknown. Thus the uncertainty on the photon distribution at some scale $\mu_F$ is given quite directly in terms of the uncertainty on the starting distribution $\gamma^{\rm in}$, and it is this object which we are principally interested in. It is perfectly possible to simply treat this as an unknown quantity in a global fit, i.e. in the same way as the quarks and gluons. This is the approach taken in the latest NNPDFQED fit~\cite{Ball:2013hta,Bertone:2016ume}, where the freely parameterised photon is fitted to DIS and a small set of LHC data, namely $W,Z$ and high/low--mass Drell--Yan production (more precisely this is achieved by Bayesian reweighting, see~\cite{Ball:2013hta} for full details). Due to the generally small contribution from photon--initiated processes, the constraining power of this data is quite limited, and the corresponding PDF uncertainties are large. However, by treating the photon PDF identically to the quark and gluons, a significant part of the available experimental information is in fact being thrown away~\cite{Gluck:2002fi,Martin:2014nqa,Harland-Lang:2016apc}. The crucial difference is that QED, in contrast to QCD, corresponds to a long range force that does not suffer from the issue of non--perturbativity at low scales. Thus a proton may coherently emit a photon ($p\to p\gamma$) at low scale $Q<Q_0\sim $ 1 GeV, and this will contribute directly to the input component in (\ref{gampdf}). Such a process is experimentally extremely well measured, being governed by the well known electric and magnetic proton form factors for coherent photon emission. In particular we have \begin{equation}\label{gamcoh} \gamma_{\rm coh}(x,Q_0^2)=\frac{1}{x}\frac{\alpha}{\pi}\int_0^{Q^2<Q_0^2}\!\!\frac{{\rm d}q_t^2 }{q_t^2+x^2 m_p^2}\left(\frac{q_t^2}{q_t^2+x^2 m_p^2}(1-x)F_E(Q^2)+\frac{x^2}{2}F_M(Q^2)\right)\;, \end{equation} where $q_t$ is the transverse momentum of the emitted photon, and $Q^2$ is the modulus of the photon virtuality, given by \begin{equation}\label{qmincoh} Q^2=\frac{q_t^2+x^2m_p^2}{1-x}\;, \end{equation} The functions $F_E$ and $F_M$ are \begin{equation}\label{form1} F_M(Q^2)=G_M^2(Q^2)\;,\qquad F_E(Q^2)=\frac{4m_p^2 G_E^2(Q^2)+Q^2 G_M^2(Q^2)}{4m_p^2+Q^2}\;, \end{equation} with \begin{equation}\label{form2} G_E^2(Q^2)=\frac{G_M^2(Q^2)}{7.78}=\frac{1}{\left(1+Q^2/0.71 {\rm GeV}^2\right)^4}\;, \end{equation} in the dipole approximation, where $G_E$ and $G_M$ are the `Sachs' form factors, which have been very precisely measured in a range of elastic $ep$ scattering experiments\footnote{The dipole approximation describes the available data to within a few percent in the lower $Q^2$ region most relevant to our study, however this description is not perfect, and a completely precise calculation would go beyond this and in addition should consider the uncertainties associated with the available form factor data. For the purposes of this paper, however, such a high level of precision is not necessary.}. As the contribution to $ep$ scattering for low photon $Q^2$ is dominantly coherent, we expect (\ref{gamcoh}) to give the dominant contribution to the input photon PDF. Thus the input photon distribution is in fact already well determined. More precisely, in general there will also be some contribution from incoherent emission ($\gamma p \to \gamma X$), where the proton dissociates after the scattering process. That is, we have \begin{equation}\label{inputdef} \gamma(x,Q_0^2)=\gamma_{\rm coh}(x,Q_0^2)+\gamma_{\rm incoh}(x,Q_0^2)\;, \end{equation} where the second term corresponds to this incoherent input; it is this combined input PDF, including both coherent and incoherent components, which corresponds to the freely parameterised NNPDF distribution described above. In general, as recently discussed in~\cite{Manohar:2016nzj} this incoherent contribution may be constrained from experimental data on $F_2$ and $F_L$, however for our considerations it is sufficient to use a simplified model which gives an upper bound on such a contribution. Thus, following~\cite{Gluck:2002fi,Martin:2014nqa} we model this emission process as being due to one photon emission from the valence quarks in the leading--logarithmic approximation; such an approach is also taken in~\cite{Martin:2004dh,Schmidt:2015zda} to model the photon PDFs, although in these cases no coherent component is included. We take\footnote{In fact, we take the slightly different form described in footnote 3 of~\cite{Martin:2014nqa}, with as in (\ref{gamincoh}) the replacement $F_1(Q^2)\to G_E(Q^2)$ made to give a more precise evaluation for the probability of coherent emission.} \begin{equation}\label{gamincoh} \gamma_{\rm incoh}(x,Q_0^2)=\frac{\alpha}{2\pi}\int_x^1\frac{{\rm d}z}{z}\left[\frac{4}{9}u_0\left(\frac{x}{z}\right)+\frac{1}{9}d_0\left(\frac{x}{z}\right)\right]\frac{1+(1-z)^2}{z}\int^{Q_0^2}_{Q^2_{\rm min}}\frac{{\rm d}Q^2}{Q^2+m_q^2}\left(1-G_E^2(Q^2)\right)\;, \end{equation} where \begin{equation}\label{qminincoh} Q^2_{\rm min}=\frac{x}{1-x}\left(m_\Delta^2-(1-x)m_p^2\right)\;, \end{equation} accounts for the fact that the lowest proton excitation is the $\Delta$--isobar, and the final factor $(1-G_E^2(Q^2))$ corresponds to the probability to have no intact proton in the final state (which is already included in the coherent component). Here $m_q=m_d$($m_u$) when convoluted with $d_0$($u_0$), and the current quark masses are taken. Crucially, as the quark distributions are frozen for $Q<Q_0$, this represents an upper bound on the incoherent contribution. If we consider the momentum fraction \begin{equation} p_\gamma = \int {\rm d}x\, x\gamma(x,Q_0^2)\;, \end{equation} carried by the photon at the starting scale $Q_0^2=2\,{\rm GeV}^2$, then even for this upper bound we find \begin{equation}\label{pgam} p_\gamma^{\rm coh} =0.15\% \quad\quad p_\gamma^{\rm incoh.}=0.05\%\;, \end{equation} that is we expect $p_\gamma^{\rm incoh.} \ll p_\gamma^{\rm coh.}$, consistent with the general expectation that the emission process for low $Q^2$ photons should be dominantly coherent. As the coherent input is quite precisely determined, the uncertainty on the input photon PDF is essentially purely due to the incoherent term. Being maximally conservative, we can consider a range of incoherent inputs, with the lower bound being simply setting $\gamma^{\rm incoh.}(x,Q_0^2)=0$, and the upper bound calculated as described above. Taking this as our uncertainty band, we then expect from (\ref{pgam}) a $\sim \pm 10-15\%$ uncertainty on the photon PDF at the starting scale $Q_0$. On the other hand, at higher scales as the contribution from the evolution term in (\ref{gampdf}) becomes more significant, this uncertainty will be smaller; we will show this explicitly in the following section, where we will see that for phenomenologically relevant scales the uncertainty due to the incoherent input shrinks to $\sim \pm 5\%$. It should be emphasised that this is a relatively conservative estimate of the uncertainty on the photon PDF due to the incoherent input component. In particular, it is possible and desirable to further constrain this incoherent input in a global fit, while a more complete treatment accounting for example for the ($\Delta$...) resonant contribution to the incoherent input, and more generally constraining this with the existing $ep$ scattering data will also further reduce this uncertainty. Nonetheless, even within this simplified and conservative approach, we can see that the corresponding uncertainty on the photon PDF is already under relatively good control. In contrast, the NNPDF3.0QED set~\cite{Bertone:2016ume} gives \begin{equation} p_{\gamma}=(1.26\pm 1.26)\%\;, \end{equation} that is, a $\sim 100\%$ uncertainty, with a central value which lies much higher than that expected from the simple physical considerations above; we will see the impact of this is in the following sections. \section{The photon PDF: results}\label{sec:photonhighx} In the previous section we demonstrated how the photon PDF at the starting scale is already quite precisely determined in terms of the form factors for coherent $p\to p \gamma$ emission. We will now demonstrate the impact this result has on the photon PDF at higher $x$ values. We show in Fig.~\ref{fig:gam2q} (left) the contributions from the coherent (\ref{gamcoh}) and incoherent (\ref{gamincoh}) components of the photon PDF at the starting scale $Q_0^2=2\, {\rm GeV}^2$, where as described above the incoherent term corresponds to an upper bound on this contribution. Here, and in all results which follow, we take treat the evolution (\ref{dglap}) at NLO in $\alpha_s$. We make use of the MMHT2014NLO~\cite{Harland-Lang:2014zoa} set for the quark PDFs in the incoherent component, as well as for the PDFs in the evolution of the photon. \begin{figure} \begin{center} \includegraphics[scale=0.75]{gamq2r1}\quad \includegraphics[scale=0.66]{tmin} \caption{(Left) The coherent (\ref{gamcoh}) and incoherent (\ref{gamincoh}) components of the photon PDF at the starting scale $Q_0^2=2\, {\rm GeV}^2$. (Right) The minimum photon $Q^2$ vs. the momentum fraction $x$ carried for the coherent (\ref{gamcoh}) and incoherent (\ref{gamincoh}) emission processes.} \label{fig:gam2q} \end{center} \end{figure} As expected from the previous discussion, the incoherent component is smaller than the well determined coherent component, and constitutes $\sim 25\%$ of the total photon PDF at intermediate values of $x$, consistent with (\ref{pgam}). However, interestingly the ratio of incoherent to coherent is found to decrease with increasing $x$, such that in the higher $x\gtrsim 0.1$ region, the coherent component is particularly dominant. This is due in part to the decreasing phase space for photon emission from the individual quarks, with the range of the $z$ integral in (\ref{gamincoh}) decreasing with increasing $x$. In addition to this, another physical effect is playing a role, due to the minimum photon virtuality $Q^2_{\rm min}$, given by (\ref{qmincoh}) and (\ref{qminincoh}). The kinematic minimum (\ref{qmincoh}) follows simply from the on--shellness requirement for the outgoing proton, and similarly in (\ref{qminincoh}) for the outgoing $\Delta$ resonance. In the latter case this corresponds to the mass of the lowest lying resonance above the proton: for higher mass resonance production, the kinematic minimum will be larger still. The effect of this is shown in Fig.~\ref{fig:gam2q} (right). Due to the higher mass of the dissociating state in the latter incoherent case, the minimum photon virtuality can be quite large for higher $x$, with the effect that contribution from low--scale incoherent photon emission becomes kinematically limited; by construction we must have $Q^2<Q_0^2\sim 1\,{\rm GeV^2}$, while the contribution for photon $Q^2>Q_0^2$ is given by the evolution component in (\ref{gampdf}), in terms of the relatively well constrained quark PDFs. This effect can be seen in Fig.~\ref{fig:gam2q} (left) in the high $x$ region, where the coherent component becomes completely dominant. \begin{figure} \begin{center} \includegraphics[scale=0.85]{gammu100rn} \includegraphics[scale=0.85]{gammu2rn} \caption{The photon PDF at scale $\mu_F=100$ and 2000 GeV, with the breakdown between coherent, incoherent and evolution components, defined as in (\ref{gampdf}) and (\ref{inputdef}) given. Also shown is the NNPDF3.0~\cite{Bertone:2016ume} result, with the corresponding $68\%$ C.L. uncertainty bands, and the LUXqed~\cite{Manohar:2016nzj} prediction. In the lower plots the ratios of the different components to the total photon PDF are shown.} \label{fig:pdfcomp} \end{center} \end{figure} In Fig.~\ref{fig:pdfcomp} (right) we show the photon PDF at $\mu_F=2$ TeV, corresponding to $x\sim 0.2$ at the LHC, with the contributions from the input coherent and incoherent, and evolution components shown explicitly. We also show in Fig.~\ref{fig:pdfcomp} (left) the corresponding PDF for the lower scale $\mu_F=10$ GeV, to give an indication of the influence of the photon evolution on these different components. We can see that the effect of evolution is as expected to further decrease the contribution from the incoherent input, which is already $\lesssim 10\%$ of the total photon at $\mu_F=100$ GeV, and even lower for $\mu_F=2$ TeV. Thus in the cross section results which follow we expect a conservative $\sim \pm 5\%$ uncertainty due to this effect. The coherent contribution, even at the higher scale $\mu_F=2$ TeV is $\sim 20\%$ at $x\sim 0.01$, and increases to $\sim 50\%$ at higher $x$; for $\mu_F=100$ GeV it is larger still. The \texttt{LUXqed} prediction is shown and is found to coincide quite closely with our result, although some deviation is visible, in particular at higher $x$. The NNPDF3.0 distribution, with the corresponding $68\%$ C.L. uncertainty bands, is also shown in Fig.~\ref{fig:pdfcomp}: here, and in all results which follow, we take the NLO set with $\alpha_s(M_Z)=0.118$. For the lower $x$ region the contribution from the high scale $q\to q\gamma$ evolution component in (\ref{gampdf}) is dominant, and as a result the corresponding uncertainties are under reasonable control\footnote{The slight deviation between our results and the NNPDF sets, even accounting for the PDF uncertainties, at lower $x$ is due to the differing `truncated' solution to the DGLAP equation applied in the latter case.}. As $x$ increases, however, the phase space for the DGLAP $q\to q\gamma$ emission process decreases, and the contribution from the coherent photon input becomes more important. This effect is evident in the NNPDF set, where the increasing contribution from the poorly determined input photon leads to a rapidly increasing uncertainty as $x$ increases. In Fig.~\ref{fig:pdflumi} we show the corresponding PDF luminosities, defined as \begin{equation}\label{lumi} \frac{{\rm d}\mathcal{L}_{ij}}{{\rm d} \ln M_X^2}= \frac{M_X^2}{s}\int^1_{\tau} \frac{{\rm d}x_1}{x_1} \,f_i(x_1,M_X^2)f_j(\tau/x_1,M_X^2)\;, \end{equation} where $\tau=M_X^2/s$ and $f_i$ is the corresponding PDF for parton $i$. As well as the $\gamma\gamma$ case discussed above, we also show for comparison the $qq$, $q\overline{q}$ (defined in both cases as a uniform sum over the 5 corresponding quark flavours) and $gg$ cases, using the same NNPDF set. For our prediction, we now for illustration include an uncertainty band due to varying the incoherent component between $x\gamma(x,Q_0)=0$ and the upper bound of (\ref{gamincoh}), although in the plots this is essentially invisible within the width of the central curves. Other uncertainties, due for example to the quark (and at higher orders, gluon) PDFs entering the photon evolution in (\ref{dglap}), the use of the dipole approximation (\ref{form2}) for the elastic form factor and the choice of $Q_0$ in (\ref{gampdf}) are not included here. These effects are expected to be generally subleading in comparison to that due to the incoherent input, and will be omitted in the results which follow. Nonetheless, it is worth bearing in mind that the effect of these will be to increase the total uncertainty on the photon PDF somewhat, which should be accounted for in a complete analysis; for the current purposes, however, this is not necessary. The \texttt{LUXqed} prediction is shown and is again found to coincide quite closely with our result, with some deviation at higher $M_X$. The same increase in Fig.~\ref{fig:pdflumi} in the NNPDF uncertainty band at high $M_X$ for the $\gamma\gamma$ case is clear. However, interestingly we can see that the trend in the central value of the NNPDF $\gamma\gamma$ luminosity is remarkably different compared to the other partons, with the former decreasing much less rapidly at high $M_X$, i.e. high $x$. On the other hand, our prediction shows no such significant difference, and roughy follows the same trend as in the quarks. As discussed in~\cite{Mangano:2016jyj} some steepening of the PDF luminosities for the QCD partons may be expected due to the differing behaviours of $\alpha_{\rm QED}$ and $\alpha_s$ at higher scales. However this effect, which is indeed observable in particular upon comparison of our result for the $\gamma\gamma$ and the $gg$ luminosity, is relatively small and cannot explain the difference seen in the NNPDF case. We are therefore led to conclude that this potentially significant difference is an artefact of the large uncertainties in the NNPDF photon PDF; the physically motivated photon PDF of our approach, which lies towards the lower end of the NNPDF uncertainty band, displays no significant difference in behaviour at higher $x$ compared to the quarks and gluons. \begin{figure} \begin{center} \includegraphics[scale=0.65]{lumi13n} \includegraphics[scale=0.65]{lumi100n} \caption{The $\gamma\gamma$, $gg$, $q\overline{q}$ and $qq$ PDF luminosities. The $\gamma\gamma$ case is shown for the NNPDF3.0~\cite{Bertone:2016ume} set and following the approach of Section~\ref{sec:input} , while all other luminosities correspond to the NNPDF set. The corresponding $68\%$ C.L. uncertainty bands are shown in the NNPDF cases, while an uncertainty band due to varying the incoherent component between $x\gamma(x,Q_0)=0$ and the upper bound of (\ref{gamincoh}) is shown, although barely visible, for our prediction. The LUXqed~\cite{Manohar:2016nzj} prediction is also shown.} \label{fig:pdflumi} \end{center} \end{figure} It is therefore in this higher $x$ region that the importance of including all available information about the photon PDF is clearest; by excluding the additional input which comes from considering the physics of the dominantly coherent photon emission process at the starting scale $Q_0$, the corresponding PDF uncertainties are dramatically over--inflated. By including this information, as in Section~\ref{sec:input}, the predicted photon PDF at higher $x$ is determined quite precisely to lie close to the lower edge of the NNPDF uncertainty band. It has for example been pointed out in~\cite{Bourilkov:2016qum,Accomando:2016tah} that the upper limits on the NNPDF photon PDF predict potentially sizeable photon--initiated contributions to the Drell--Yan cross section at high mass, with large corresponding PDF uncertainties. From the above considerations, however, we expect this not to be the case: we will consider this in more detail in the following section. \section{Cross section predictions}\label{sec:dy} As discussed in the introduction, the photon--initiated contribution to lepton and $W$ boson pair production may be particularly significant at higher mass, where the production cross sections are relatively enhanced due to the $t$--channel nature of the corresponding Feynman diagrams. We therefore consider predictions for both these processes at the LHC and FCC in this section. We use our own implementation of these processes, with the corresponding LO cross sections given by (see e.g.~\cite{Brodsky:1970vk,Tupper:1980bw}) \begin{align}\nonumber \frac{{\rm d}\sigma}{{\rm d}\cos\theta^*}(\gamma\gamma\to W^+W^-) &= \frac{\pi\alpha^2\beta}{\hat{s}}\frac{19-6\beta^2(1-\beta^2)+2(8-3\beta^2)\beta^2\cos^2\theta^*+3\beta^4\cos^4\theta^*}{(1-\beta^2\cos^2\theta^*)^2} \,, \\ \frac{{\rm d}\sigma}{{\rm d}\cos\theta^*}(\gamma\gamma\to l^+l^-) &= \frac{2\pi\alpha^2\beta}{\hat{s}}\frac{1+2\beta^2(1-\beta^2)(1-\cos^2\theta^*)-\beta^4\cos^4\theta^*}{(1-\beta^2\cos^2\theta^*)^2}\,, \end{align} where $\beta=(1-4m^2/\hat{s})^{1/2}$, with $m=m_{W},m_{l}$, and $\theta^*$ is the angle of the outgoing particles with respect to the photons in the $\gamma\gamma$ C.M. frame. As we are interested in the high mass regime, in the lepton case we will for concreteness take the massless limit, $m_l=0$, in what follows. In Fig.~\ref{fig:cross13} we show the lepton pair production cross section via the photon--initiated production and Drell--Yan production mechanisms, for lepton $|\eta|<2.5$ and $p_\perp> 20$ GeV. The former is shown using both the approach of Section~\ref{sec:input} and with the NNPDF3.0QED set, while the latter is calculated at NLO in $\alpha_s$ with \texttt{MCFM}~\cite{Campbell:2011bn} using MMHT2014NLO~\cite{Harland-Lang:2014zoa} PDFs; the results which follow are not affected significantly by NNLO corrections. For the curve corresponding to our approach, we show an uncertainty band due to varying the incoherent component between $x\gamma(x,Q_0)=0$ and the upper bound of (\ref{gamincoh}). The 68\% PDF uncertainty bands are shown for the \texttt{MCFM} predictions in all cases which follow, although for certain distributions this is sufficiently small that it is not visible on the plots. \begin{figure} \begin{center} \includegraphics[scale=0.65]{crosslhc} \includegraphics[scale=0.65]{crossfcc} \caption{The differential lepton pair production cross sections at $\sqrt{s}=13$ TeV and 100 TeV with respect to the invariant mass of the pair $M_{ll}$, for lepton $|\eta|<2.5$ and $p_\perp> 20$ GeV. The photon--initiated contributions predicted following the approach of Section~\ref{sec:input} and the NNPDF3.0QED~\cite{Bertone:2016ume} set, including the 68\% C.L. uncertainty bands are shown, in addition to the NLO Drell--Yan cross section, calculated with \texttt{MCFM}~\cite{Campbell:2011bn}. An uncertainty band due to varying the incoherent component between $x\gamma(x,Q_0)=0$ and the upper bound of (\ref{gamincoh}) is shown for our prediction.} \label{fig:cross13} \end{center} \end{figure} For the LHC predictions in Fig.~\ref{fig:cross13} (left), we can see that for higher $M_{ll}$ the photon--initiated cross section predicted by the NNPDF set may be comparable in size and even larger than the Drell--Yan cross section, within the increasingly large PDF uncertainty bands. This was recently discussed in~\cite{Accomando:2016tah}, where it was pointed out that the potential dominance of the photon--initiated NNPDF prediction induced a large uncertainty in the predicted cross section for high mass lepton pair production; this could, for example, have an impact on searches for new heavy particles decaying to lepton pairs. However, it is our finding that this is not the case. In particular, we can see from Fig.~\ref{fig:cross13} that even up to the highest $M_{ll}$ values the predicted contribution from the photon--initiated process is fairly small, $\sim 10\%$ of the Drell--Yan. This result is entirely consistent with the expectations from Fig.~\ref{fig:pdfcomp}. Thus we expect no significant contamination from the photon--initiated process. For the FCC case shown in Fig.~\ref{fig:cross13} (right), which was recently discussed in~\cite{Mangano:2016jyj}, a similar trend is seen. Moreover, it is worth emphasising that for both the LHC and FCC cases, tighter cuts on the lepton transverse momentum $p_\perp$ and pseudorapidity $\eta$ will further decrease the relative contribution from the photon--initiated process, which being due to the $t$ and $u$ channel diagrams is more strongly peaked towards low $p_\perp$ and high $\eta$. \begin{figure} \begin{center} \includegraphics[scale=0.65]{crosswlhc} \includegraphics[scale=0.65]{crosswfcc} \caption{The differential $W$ boson pair production cross sections at $\sqrt{s}=13$ TeV and 100 TeV with respect to the invariant mass of the pair $M_{WW}$, for $W$ pseudorapidity $|\eta|<4$. The photon--initiated contributions predicted following the approach of Section~\ref{sec:input} and the NNPDF3.0QED~\cite{Bertone:2016ume}, including the 68\% C.L. uncertainty bands are shown, in addition to the NLO QCD cross section, calculated with \texttt{MCFM}~\cite{Campbell:2011bn}, and including the gluon--initiated box contribution. An uncertainty band due to varying the incoherent component between $x\gamma(x,Q_0)=0$ and the upper bound of (\ref{gamincoh}) is shown for our prediction.} \label{fig:crossw} \end{center} \end{figure} In Fig.~\ref{fig:crossw} we show predictions for the $W$ boson pair production cross sections, again at the LHC and FCC. We impose the same cuts on the $W$ boson pseudorapidities, and include no further decays, as in~\cite{Mangano:2016jyj}, for the sake of comparison. \texttt{MCFM}~\cite{Campbell:2011bn} with MMHT2014NLO~\cite{Harland-Lang:2014zoa} PDFs is used to generate the the QCD $WW$ production process at NLO in $\alpha_s$, with the $gg$--initiated box contribution also included. Again a similar trend is clear, with the NNPDF set predicting potentially a completely dominant photon--initiated contribution at higher masses, within very large uncertainties. However, for the LHC our approach predicts instead that the standard QCD--initiated is dominant, apart from at the very highest masses. On the other hand for the FCC this is no longer the case: over the mass range considered the $\gamma\gamma$ and QCD--initiated contributions are generally expected to be comparable in size. In this case a very careful treatment of the photon PDF uncertainties will be essential. It should be emphasised that the predicted cross sections within our approach lie entirely within the NNPDF uncertainty bands, and are therefore completely consistent with these. The issue is simply that the NNPDF approach, by omitting the physical constraints on the photon input described in Section~\ref{sec:input}, allows in principle unphysically large input photon distributions, the effect of which becomes increasingly dominant at higher $x$, where the contribution from the DGLAP $q\to q\gamma$ emission from the quarks becomes smaller. On the other hand, the NNPDF starting distribution effectively parameterises the contribution from both the coherent and incoherent input components as in (\ref{inputdef}), but without any further constraints, and therefore we fully expect consistency within both approaches, once the PDF uncertainties have been properly included; it is encouraging to find that this is indeed the case. Moreover, as further data from the LHC is included in the NNPDF fit, we fully expect this consistency to continue as the PDF uncertainties decrease. \begin{figure} \begin{center} \includegraphics[scale=0.65]{crosslhc8} \caption{The differential lepton pair production cross section at $\sqrt{s}=8$ TeV with respect to the invariant mass of the pair $M_{ll}$, calculated as in Fig.~\ref{fig:cross13}.} \label{fig:cross8} \end{center} \end{figure} Indeed, this expectation is supported by the recent ATLAS measurement~\cite{Aad:2016zzw} of high mass lepton pair production in the $116<M_{ll}<1500$ GeV region at $\sqrt{s}=8$ TeV, corresponding to $0.015\lesssim x \lesssim 0.2$. We show the predicted cross section for this mass region in Fig.~\ref{fig:cross8}, and we can see that for larger $M_{ll}$ quite significant photon--initiated contributions are allowed within the NNPDF uncertainty bands. However, in~\cite{Aad:2016zzw} a Bayesian reweighting exercise is performed, and it is found that a photon PDF which lies systematically on the lower end of the NNPDF2.3 uncertainty band is preferred, with greatly reduced uncertainties, see in particular Fig.~13 of~\cite{Aad:2016zzw}\footnote{It is worth emphasising that in reality it is only for $x \gtrsim 0.05$, i.e. $M_{ll}\gtrsim 400$ GeV, that the ATLAS data begin to place any significant constraint on the photon. Thus the form of the reweighted photon distribution below $x\sim 0.05$ should not be taken too literally.}. While a full refit would be required to make a completely firm conclusion, this nonetheless provides a strong indication that the larger photon--initiated cross section predictions allowed by the higher $x$ NNPDF photon uncertainty band are already disfavoured by this ATLAS data, while our result, which predict lower photon--initiated cross sections, are qualitatively consistent. Further experimental support for this result at the LHC can be found in the ATLAS~\cite{Aaboud:2016dkv} and CMS~\cite{Chatrchyan:2013akv,Khachatryan:2016mud} measurements of exclusive two--photon induced $W^+W^-$ production. Here, events are selected by demanding a dilepton vertex with no additional associated charged tracks within the tracker acceptance ($\eta<$ 2.4). As part of these measurements, a larger data sample of $\mu^+\mu^-$ events is also selected with the same track veto applied. Crucially, such a veto effectively isolates the photon--initated contribution; the standard Drell--Yan production process will dominantly produce additional tracks centrally, and the remaining contribution can be suppressed with further cuts and then subtracted using MC simulation. The photon--initiated cross section for this semi--exclusive case was considered in~\cite{Harland-Lang:2016apc}, where it was demonstrated how (\ref{gampdf}) could be relatively simply modified to account for such a rapidity gap veto. Predictions for semi--exclusive $\mu^+\mu^-$ production within the ATLAS/CMS event selection were presented, and it was shown that the data are completely consistent with the approach described in this paper, but are in strong tension with the higher cross sections allowed by the NNPDF2.3 set. \section{Conclusion}\label{sec:conc} Processes with initial--state photons are becoming increasingly relevant for phenomenology at the LHC, in particular as the requirement for high precision becomes standard. Given this, it is becoming increasingly important to constrain as precisely as possible the corresponding photon PDF. This is particularly so for processes such as high mass lepton and $W$ boson production, where the contribution from the photon--initiated process may be significant. In this paper we have described how the photon PDF may we be quite precisely constrained by relatively simple considerations about the nature of the initial--state photon, see~\cite{Martin:2014nqa,Harland-Lang:2016apc}. In particular, QED, in contrast to QCD, corresponds to a long range force that does not suffer from the issue of non--perturbativity at low scales, and thus a proton may coherently emit a photon ($p\to p\gamma$) at low scales. This process is already experimentally very well measured in $ep$ collisions, and is governed by the well known electric and magnetic proton form factors for coherent photon emission. Taking these as input, we have shown how the photon PDF can already be well constrained with even a quite conservative model estimation for an upper limit on the remaining incoherent contribution to the photon at the starting scale: the uncertainty due to this is found to be generally $\sim 10\%$ or less, with the precise value depending on the scale and parton $x$. A more detailed treatment of the (resonant and non--resonant) incoherent contribution, and the inclusion of further constraints within the context of a global fit will reduce this even further. Indeed, as discussed in the recent analysis of~\cite{Manohar:2016nzj}, it is possible to place quite precise constraints by using low $Q^2$ inelastic structure function data. In this work, a dominantly coherent component is also included, and we have seen that the resultant photon PDF lies quite close to our prediction. These constraints are not included in the currently available global fits~\cite{Martin:2004dh,Ball:2013hta,Schmidt:2015zda} and may lead to a signifiant overestimation in the uncertainty in the photon PDF. This can have important phenomenological implications, and indeed this is the case for high mass lepton and $W$ boson pair production, which have recently been discussed in~\cite{Bourilkov:2016qum,Accomando:2016tah,Mangano:2016jyj}. Here, it was found that the NNPDF2.3QED~\cite{Ball:2013hta} set, which freely parameterises the input photon and fits to a selection of DIS and LHC data, is consistent with strongly dominant photon--initiated contributions to both of these processes at higher masses, at the LHC and FCC. In this paper we have shown that this is not expected to be the case. We have found that the DY contribution to lepton pair production is dominant out to very high masses, while for $W$ boson pair production the QCD process is generally larger, although for higher masses the $\gamma\gamma$--initiated contribution is comparable in size. In this latter case, therefore, a precise treatment of the photon PDF will be essential. \newline \section*{Acknowledgements} We thank Valerio Bertone, Amanda Cooper--Sarkar, David D'Enterria and Juan Rojo for useful discussions. VAK thanks the Leverhulme Trust for an Emeritus Fellowship. The work of MGR was supported by the RSCF grant 14-22-00281. LHL thanks the Science and Technology Facilities Council (STFC) for support via the grant award ST/L000377/1. MGR thanks the IPPP at the University of Durham for hospitality.
192b355180b05a1fe9ebfa1db5f1ef34cec8d5f9
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Jovian Trojans are actually large populations of minor bodies in the 1:1 mean motion resonance (MMR) with Jupiter, librating around $L_4$ and $L_5$ Lagrangian points. In general, there are two classes of theories explaining their origin: i) a theory in the framework of accretion model (e.g. Goldreich 2004, Lyra et al. 2009) and ii) a capture of bodies located in libration zones during a migration of giant planets (Morbidelli et al. 2005, Morbidelli et al. 2010, Nesvorn\'y et al. 2013), which is preferred in our solar system. Since the librating regions are very stable in the current configuration of planets and they are surrounded by strongly chaotic separatrices, bodies from other source regions (e.g. Main belt, Centaurs, Jupiter family comets) cannot otherwise enter the libration zones and Jupiter Trojans thus represent a rather primitive and isolated population. Several recent analyses confirmed the presence of several families among Trojans (e.g. Nesvorný et al. 2015, Vinogradova, 2015). The Trojan region as such is very favourable for dynamical studies of asteroid families, because there is no significant systematic Yarkovsky drift in semimajor axis due to the resonant dynamics. On the other hand, we have to be aware of boundaries of the libration zone, because ballistic transport can cause a partial depletion of family members. At the same time, as we have already shown in Bro\v{z} \& Rozehnal (2011), no family can survive either late phases of a slow migration of Jupiter, or Jupiter ``jump'', that results from relevant scenarios of the Nice model (Morbidelli et al. 2010). We thus focus on post-migration phase in this paper. We feel the need to evaluate again our previous conclusions on even larger datasets, that should also allow us to reveal as-of-yet unknown structures in the space of proper elements or unveil possible relations between orbital and physical properties (e.g. albedos, colours, diameters) of Jovian Trojans. In Section \ref{sec:obsdata} we use new observational data to compute appropriate resonant elements. In Section \ref{sec:physchar} we use albedos obtained by Grav et al. (2012) to derive size-frequency distributions and distribution of albedos, which seem to be slightly dependent on the proper inclination $I_{\rm p}$. In Section \ref{sec:groups} we identify families among Trojans with our new ``randombox'' method. We discuss properties of statistically significant families in Section \ref{sec:props}. Then we focus mainly on the Hektor family because of its unique D--type taxonomical classification, which is the first of its kind. We also discuss its long-term dynamical evolution. In Section \ref{sec:coll_model} we simulate collisional evolution of Trojans and we estimate the number of observable families among Trojans. Finally, in Section \ref{sec:SPH} we simulate an origin of the Hektor family using smoothed-particle hydrodynamics and we compare results for single and bilobed targets. Section \ref{sec:conclusions} is devoted to Conclusions. \section{New observational data}\label{sec:obsdata} \begin{figure*} \centering \renewcommand{\tabcolsep}{0pt} \begin{tabular}{cc} $L_4$ Trojans & $L_5$ Trojans \\ \includegraphics[width=9cm]{figures/aiL4v3color.eps} & \includegraphics[width=9cm]{figures/aiL5v3color.eps} \\ \includegraphics[width=9cm]{figures/eiL4v3color.eps} & \includegraphics[width=9cm]{figures/eiL5v3color.eps} \\ \end{tabular} \caption{ The resonant semi-major axis vs inclination $(a_{\rm{p}}, I_{\rm{p}})$ (top) and eccentricity vs inclination $(e_{\rm{p}}, I_{\rm{p}})$ (bottom) for $L_4$ (left) and $L_5$ Trojans (right). The circles indicate relative diameters of bodies, as determined by WISE (Grav et al. 2011), or when unavailable, computed from the absolute magnitude~$H$ and geometric albedo~$p_{\rm V}$, which we assumed to be $p_{\rm V} = 0.07$ for both the $L_4$ and $L_5$ Trojans (WISE median value is $p_{\rm V} = 0.072$ for $L_4$ and $p_{\rm V} = 0.069$ for $L_5$ Trojans). Colours correspond to the values of $p_{\rm V}$, blue are dark ($p_{\rm V} \simeq 0.05$) and yellow are bright ($p_{\rm V} \simeq 0.25$). One can see clearly all asteroid families on this plot, especially in $(a_{\rm{p}}, I_{\rm{p}})$, because they tend to be confined in inclinations.} \label{trojans_L4_arer_sizes} \end{figure*} \subsection{Resonant elements} We computed resonant elements, i.e. the averaged semimajor axis $\bar{a}$, libration amplitude $\Delta a_{\rm p}$, eccentricity $e_{\rm p}$ and inclination $I_{\rm p}$ of 3907 Trojans in $L_4$ cloud and 1945 Trojans in $L_5$ cloud. As an input, we used osculating elements listed in AstOrb catalogue (Bowell et al. 2002), released in July 2014. A detailed description of the resonant elements computation can be found in Bro\v{z} \& Rozehnal (2011). Positions of Trojans in the space of proper elements ($a_{\rm p}$, $I_{\rm p}$), where $a_{\rm p} = \bar{a} + \Delta a_{\rm p}$, and ($e_{\rm p}$, $I_{\rm p}$), calculated with a suitably modified version of the SWIFT integrator (Levison \& Duncan, 1994), are presented graphically in Figure \ref{trojans_L4_arer_sizes}, together with their sizes and albedos. \footnote{The table of resonant elements is listed online at \tt{http://sirrah.troja.mff.cuni.cz/$\sim$mira/mp/trojans/}.} \subsection{WISE and AKARI albedos and diameters} To construct size-frequency distributions of the whole $L_4$ and $L_5$ Trojan populations and later of individual families, we mostly used WISE albedos and diameters derived by Grav et al. (2012). We also compared the respective values to AKARI, as reported by Usui et al. (2011).\footnote{While there are some differences between individual values even at $3\sigma$ level, they do not seem to be important for population studies like ours.} We used albedo values of 1609 Trojans in both $L_4$ and $L_5$ clouds obtained by WISE; about one third of these albedos were obtained during cryo-phase, the rest were measured in post-cryo-phase (see Grav et al., 2011). \section{Physical characterisation of Trojan populations}\label{sec:physchar} \subsection{Albedo distribution and taxonomy} The values of visible albedos $p_{\rm V}$ of Trojans derived by Grav et al. (2012) vary in the range from $p_{\rm V} = 0.025$ to $p_{\rm V} \simeq 0.2$. Distributions of albedos are qualitatively the same for both $L_4$ and $L_5$ populations. The median albedo of WISE sample is $\widetilde{p}_{\rm v}=0.072 \pm 0.017$ for $L_4$ and $\widetilde{p}_{\rm v}=0.069 \pm 0.015$ for $L_5$. These values of visible albedos mostly correspond to C or D taxonomical classes in Tholen taxonomic classification scheme (Mainzer et al., 2011). However, there is a significant presence of small asteroids $(D<15\,\rm km)$ with apparently high albedo --- almost 20\,\% of asteroids in $L_4$ and 13\,\% of asteroids in $L_5$ have albedo $p_{\rm V} > 0.10$. As stated in Grav et al. (2012), this is probably not a physical phenomenon, it is rather due to the fact that for small diameters the photon noise contribution becomes too significant. When we compute the median albedo from AKARI data, we realize that its value is slightly lower ($\widetilde{p}_{\rm v}=0.054 \pm 0.005$) than that from WISE, but when we compute the median from WISE values for the same asteroids which are listed in AKARI catalogue, we obtain a similar value ($\widetilde{p}_{\rm v}=0.061 \pm 0.012$). What is more serious, AKARI and WISE data differ considerably for large asteroids with $D>100\,\rm km$ --- the average difference between albedos is $|p_{\rm V_{AKARI}} - p_{\rm V_{WISE}}| = 0.02$. The same difference we see in derived diameters. These discrepancies may be caused for example by limitations of the thermal model (cf. spheres in NEATM models). Hereinafter, we prefer to use the WISE data when available, because they represent orders of magnitude larger sample than AKARI. When we split Trojan asteroids according to their albedo into two rather artificial subpopulations with $p_{\rm V} < 0.08$ and $p_{\rm V} > 0.08$ respectively, and then we compute distributions of these subpopulations with respect to the resonant inclination $I_{\rm p}$, we get two different pictures. As can be seen in Figure \ref{albedo_separation_L4_hist}, most bodies have resonant inclinations $I_{\rm p} < 15^{\circ}$, but there are 77\,\% of bodies with higher albedo with $I_{\rm p} < 15^{\circ}$, while only 55\,\% of the population with lower albedo is located in the same range of inclinations. This is a similar phenomenon as described by Vinogradova (2015), who reported different upper limits in inclinations for different taxonomical types obtained mostly from SDSS colour data. \begin{figure} \centering \includegraphics[width=8.5cm]{figures/albedo_separation_L4_hist.eps} \caption{The differential histogram of the resonant inclination $I_{\rm p}$ for $L_4$ Trojans with a lower albedo ($p_{\rm V} < 0.08$, red) and a higher one ($p_{\rm V} >0.08$, green). Eurybates family was removed from the dataset.} \label{albedo_separation_L4_hist} \end{figure} \subsection{Size-frequency distributions}\label{sec:SFDs} The WISE data (Grav et al. 2011, 2012) provide very useful source of information on diameters we need to construct size-frequency distributions (SFDs) of Trojan populations in $L_4$ and $L_5$. However, the sample measured by WISE is not complete. In our previous work (Bro\v{z} and Rozehnal, 2011), we constructed the SFDs assuming a constant albedo which we set to be equal to the median albedo of Trojans that was measured back then. Since the number of measurements was very low (several tens), this was the only reasonable way. Now we choose another method to construct more reliable SFDs. As we calculated resonant elements for more than 5800 Trojans and we have more than one quarter of appropriate albedos, we constructed the SFDs by assigning albedos randomly from the observed WISE distribution to the remaining Trojans, whose albedo was not measured. To avoid a bias, we compared different SFDs constructed with different random generator seeds and we realized that the overall shape of SFDs does not change noticeably, the slope $\gamma$ varies in the range of $\pm 0.1$ at most. The SFDs we constructed this way are shown in Figure \ref{SFD_WISE}. The SFDs for the $L_4$ and $L_5$ clouds look slightly different, especially in the size range from 60\,km to 100\,km. This part of the SFD is not influenced by the Eurybates family, the largest family among Trojans, because all its members have diameters $D<50\,\rm{km}$. We used these SFDs to determine the ratio of the number of asteroids in $L_4$ and $L_5$ clouds. There are 2746 asteroids with diameter $D>8\,\rm km$ in $L_4$ and 1518 asteroids in $L_5$. When we remove all family members with diameters $D>8\,\rm km$, we have 2436 asteroids in~$L_4$ and 1399 in~$L_5$. However, this sample may be still influenced by debris produced by catastrophic disruptions of small bodies ($D\geq 50\,\rm km$), which need not to be seen as families. Counting only asteroids with diameter $D>20\,\rm km$, which corresponds to the absolute magnitude $H\simeq12$, and removing family members, we get the ratio $N_{L_4}/N_{L_5}=1.3 \pm 0.1$. As this is entirely consistent with value of Nesvorn\'{y} et al. (2013), which was derived for Trojans with $H>12$, and with Grav et al. (2012), whose estimate is $N_{L_4}/N_{L_5} = 1.4 \pm 0.2$, we can confirm a persisting asymmetry between the number of $L_4$ and $L_5$ Trojans in new data. Although for bodies with diameter $D>100\,\rm km$, the $L_5$ cloud has more asteroids than $L_4$, the total number of these bodies is of the order of 10, so this is just an effect of small-number statistics and does not affect the $N_{L_4}/N_{L_5}$ ratio much. \begin{figure*} \centering \renewcommand{\tabcolsep}{0pt} \begin{tabular}{cc} $L_4$ Trojans & $L_5$ Trojans \\ \includegraphics[width=8.5cm]{figures/L4_SFD.eps} & \includegraphics[width=8.5cm]{figures/L5_SFD.eps} \\ \end{tabular} \caption{Size-frequency distributions for both $L4$ and $L_5$ Trojans, constructed using the albedos measured by WISE satellite (Grav et al. 2012). Since WISE data cover just about 18\,\% of $L_4$ and 29\,\% of $L_5$ Trojans known today, we assigned albedos randomly from the WISE distribution to the remaining Trojans. We also present SFDs of individual asteroid families discussed in the main text. There are also our fits of each SFD in the range $D=12$ to 30 km by the power law $N(>D) = C D^\gamma$. As we can see, both clouds seem to be near the collisional equilibrium ($\gamma\simeq-2.5$, Dohnanyi 1969), while most families have slope $\gamma$ significantly steeper. Of course, we can expect the slopes of the SFDs become shallower for smaller $D$ due to observational incompleteness.} \label{SFD_WISE} \end{figure*} \begin{table*}\vspace{-5mm}\centering \caption{Physical properties of Trojan families identified by both the ``randombox'' and the HCM methods. We list Family Identification Number (FIN, as in Nesvorn\'{y} et al., 2015), the designation of the family, the designation of the asteroid with which the family is associated (i.e. usually the largest remnant of the parent body), the cutoff velocity $v_{\rm cutoff}$, for which family is still clearly detached from the background, and the number of members of the family $N_{\rm{memb}}$ corresponding to the respective $v_{\rm cutoff}$. Next we list optical albedos $p_{\rm V}$ of associated bodies determined by Grav et al. (2012) from WISE observations, and their taxonomical classification.} \small \begin{tabular}{c|c|c|c|c|c|c|c|} \hline\ FIN & family designation & cloud & asteroid designation & $v_{\rm cutoff} \rm [m\,s^{-1}]$ & $N_{\rm{memb}}$ & $p_{\rm V} \rm (WISE)$ & tax. type \\ \hline\hline 004 & Hektor & $L_4$ & (624) Hektor & 110 & 90 & $0.087 \pm 0.016$ & D \\ 005 & Eurybates & $L_4$ & (3548) Eurybates & 60 & 310 & $0.060 \pm 0.016$ & C/P \\ 006 & $1996\,\rm RJ$ & $L_4$ & (9799) $1996\,\rm RJ$ & 140 & 17 & $0.082 \pm 0.014$ & -- \\ 008 & Arkesilaos & $L_4$ & (20961) Arkesilaos & 55 & 35 & n/a & -- \\ 009 & Ennomos & $L_5$ & (17492) Hippasos & 100 & 104 & $0.064 \pm 0.012$ & -- \\ 010 & $2001\,\rm UV_{209}$ & $L_5$ & (247341) $2001\,\rm UV_{209}$ & 120 & 30 & $0.088 \pm 0.023$ & -- \\ \hline \vspace{2mm} \end{tabular} \label{fams_list} \end{table*} \begin{table*}\vspace{-5mm}\centering \caption{Derived properties of Trojan families. We list here the family designation, the diameter of the largest remnant $D_{\rm LR}$, the minimal diameter of the parent body min $D_{\rm PB}$, obtained as the sum of all observed family members, the diameter of the parent body $D_{\rm PB(SPH)}$ and the mass ratio $M_{\rm{LR}}/M_{\rm{PB}}$ of the largest fragment and the parent body, both derived from our fits by scaled SPH simulations performed by Durda et al. (2007). We use this ratio to distinguish between the catastrophic disruption ($M_{\rm{LR}}/M_{\rm{PB}} < 0.5$) and the cratering ($M_{\rm{LR}}/M_{\rm{PB}}>0.5$). Finally, there is the escape velocity $v_{\rm esc}$ from the parent body and estimated age of the family derived in this and our previous work (Bro\v{z} and Rozehnal, 2011).} \small \begin{tabular}{|c|c|c|c|c|c|c|c|} \hline\ family desig. & $D_{\rm LR}$ [km] & min $D_{\rm PB}$ & $D_{\rm PB(SPH)}$ & $M_{\rm{LR}}/M_{\rm{PB}}$ & $v_{\rm esc} \rm [m\,s^{-1}]$ & age [Gyr] & notes, references\\ \hline\hline Hektor & $250 \pm 26$ & 250 & 257 & 0.92 & 73 & 0.3 or 3 & 1, 3\\ Eurybates & $59.4 \pm 1.5$ & 100 & 155 & 0.06 & 46 & 1.0 to 3.8 & 2\\ $1996\,\rm RJ$ & $58.3 \pm 0.9$ & 61 & 88 & 0.29 & 26 & -- & 2,4\\ Arkesilaos & $24 \pm 5$ & 37 & 87 & 0.02 & 16 & -- & 2\\ Ennomos & $55.2 \pm 0.9$ & 67 to 154 & 95 to 168 & 0.04 to 0.19 & 29 to 66 & 1 to 2 & 2, 5\\ $2001\,\rm UV_{209}$ & $16.3 \pm 1.1$ & 32 & 80 & 0.01 & 14 & -- & 2\\ \hline \end{tabular} \scriptsize{$^1 D_{\rm LR}$~derived~by~Marchis~et~al.~(2014)}, \scriptsize{$^2 D_{\rm LR}$ derived by Grav et al. (2012)}, \scriptsize{$^3$ bilobe, satellite (Marchis et al. 2014)}, \scriptsize{$^4$ very compact, Bro\v{z} and Rozehnal (2011)}, \scriptsize{$^5 D_{\rm PB}$ strongly influenced by interlopers}, \scriptsize{$^6$ The largest fragment of Ennomos family is (17492) Hippasos}. \label{fams_prop} \end{table*} \section{Families detection methods}\label{sec:groups} A brief inspection of the resonant-element space ($a_{\rm p}$, $e_{\rm p}$, $I_{\rm p}$) (see Figure \ref{trojans_L4_arer_sizes}), reveals several locations with higher concentrations of bodies. These could be collisional families, created by a disruption of a parent body during a random collision, but they could also originate randomly by chaotic diffusion and due to effects of secular and high-order resonances. To be regarded as a family, the cluster must comply with, inter alia, the following criteria: i) it must be concentrated in the space of proper elements; ii) the cluster must have the SFD different from that of the whole $L_4$ and $L_5$ population; iii) the last criterion is usually spectral, or at least, albedo homogeneity of family members, but so far, there are not enough sufficiently accurate data for Trojans, especially for bodies with diameters $D< 50\,\rm km$, which usually form a substantial part of Trojan families. Therefore we cannot perform any detailed spectral analysis in this work. We analyzed the space of resonant elements both in terms of mutual distances among bodies and in terms of statistical probability that clusters are \textit{not} random. \subsection{Randombox method} Besides the commonly used hierarchical clustering method (HCM, Zappal\`{a} et al., 1990), we applied a ``randombox'' method, based on numerical Monte-Carlo simulations. This method allows us to compute the statistical significance of the clusters, i.e. the probability that the cluster is a random concentration of bodies in the space of proper elements ($a_{\rm p}, e_{\rm p}, \sin{I_{\rm p}}$). We divided the space of proper elements into equally sized ``boxes'' with dimensions $\Delta a_{\rm p} = 0.025\, \rm au$, $\Delta e_{\rm p} = 0.2$ and $\Delta \sin{I_{\rm p}} = 0.025$. Then we created $N = 100,000$ random distributions of the same number of bodies which are observed together in the given box and two adjacent boxes (in the direction of the $y$-axis, cf. Figure \ref{Randombox}), and we counted number of positive trials $N^+$, for which the randomly generated number of bodies in the central box was larger than the observed one. From here we can calculate the probability $P_{\rm rnd}$, that the observed number of bodies in the box is random: $P_{\rm rnd} = N^+ / N$. Alternatively, one can also use our analytical formula: \begin{equation}\label{randombox_p} p_{\rm rnd} = \frac{\sum _{k=n_2} ^n C(n,k)V^\prime (n_{\rm box}-1,n-k)}{V^\prime(n_{\rm box},n)}\,, \end{equation} where $n$ denotes the total number of bodies, $n_{\rm box}$ is the total number of boxes (3 in our case), $n_2$ is the observed number of bodies in the middle box, $k$ is the number of observed bodies in the current box, $C(n,k)$ are combinations without repetitions, i.e. the total number of trials to select $k$ bodies observed in the current box from the total number of $n$ bodies; $V^\prime (n_{\rm box}-1,n-k)$ are variations with repetitions, i.e. the total number of trials to distribute the remaining bodies into the remaining boxes; and $V^\prime(n_{\rm box},n)$ are also variations with repetitions, i.e. the total number of trials to distribute all $n$ bodies into all $n_{\rm box}$ boxes. We verified the results of the analytical formula (\ref{randombox_p}) by the MC method. We plot the results in Figure \ref{Randombox} for both the $L_4$ and $L_5$ clouds. In comparison with Figure \ref{trojans_L4_arer_sizes}, one can see that for all clusters we identified as families the probability $P_{\rm rnd}$ varies between $2\cdot 10^{-3}$ and $5\cdot 10^{-5}$, i.e. the probability that clusters are random fluctuations is indeed very low. We also re-evaluated all families identified by the hierarchical clustering method using the ``randombox'' method, which makes our decision whether the cluster is a real family much more quantitative. \begin{figure*} \renewcommand{\tabcolsep}{0pt} \renewcommand{\arraystretch}{0.2} \centering \begin{tabular}{cc} $L_4$ Trojans & $L_5$ Trojans \\ \\ \includegraphics[width=8.9cm]{figures/RandomboxL4.eps} & \includegraphics[width=8.9cm]{figures/RandomboxL5.eps} \\ \end{tabular} \caption{The statistical significance $p$ expressed as colour on the logarithmic scale for observed asteroids in the proper semimajor axis vs proper inclination plane ($a_{\rm{p}},\sin I_{\rm{p}}$) (i.e. the same data as in Figure \ref{trojans_L4_arer_sizes}). $L_4$ Trojans are on the left, $L_5$ Trojans on the right. We computed the values of $p$ for 7 times 18 boxes using our ``randombox'' method The range in proper eccentricity is 0.00 to 0.20. Statistically significant groups appear as orange boxes and they correspond to the families reported in Table \ref{fams_list}.} \label{Randombox} \end{figure*} \subsection{Hierarchical clustering method}\label{sec:HCM} We also used the HCM independently to extract significant clusters. Families identified by both the ``randombox'' and HCM methods are listed in Table \ref{fams_list}. For each family, we constructed a dependence of the number of members of the cluster $N_{\rm{memb}}$ on the cutoff velocity $v_{\rm{cutoff}}$. Because the number of members of a real collisional family rises first slowly with rising $v_{\rm{cutoff}}$ (Bro\v{z} and Rozehnal, 2011) --- in contrast with random clusters which are merging very quickly with the background --- the constructed dependence allows us to guess a realistic number of family members $N_{\rm{memb}}$. For all families listed in Table \ref{fams_list} we were convinced that they fulfill this criterion. However, we cannot distinguish possible interlopers this way, and it is also possible that some fraction of family members with high $v_{\rm{cutoff}}$ (so called halo, as in Bro\v{z} and Morbidelli, 2013) remains unidentified in the surrounding background. \section{Properties of statistically significant families}\label{sec:props} \subsection{Eurybates} As we have already demonstrated in Bro\v{z} and Rozehnal (2011), the family associated with asteroid (3548) Eurybates is the largest collisional family, and it is the only family among Trojans with the parent body size $D_{\rm{PB}} > 100\,\rm{km}$, which originated by a catastrophic disruption (this means that the mass ratio of the largest remnant to the parent body $M_{\rm{LR}}/M_{\rm{PB}}< 0.5$). Using new albedos derived by Grav et al. (2012), we recalculated the overall SFD slope of the family to be $\gamma = -3.4 \pm 0.1$. As the WISE sample provides albedos for only about 1/5 of the family members, we calculated two values of~$\gamma$: the first one assuming that remaining asteroids have a constant albedo $p_{\rm V} = 0.06$, the second one by assigning albedos randomly from the WISE distribution, as described in Section \ref{sec:SFDs}. Both values are equal within their errorbars. The new slope $\gamma$ is significantly steeper than our previous calculation ($\gamma = -2.5 \pm 0.1$), derived with the assumption of a constant albedo of all members of the family. The lower value was most likely caused by a significant observational incompleteness in the size range from $D=12\,\rm km$ to $D=30\,\rm km$. We also derived the new value of the parent body diameter, which is still above the limit of 100 km. An extrapolation of the SFD by a power law gives the value $D_{\rm{PB}}\simeq 140\,\rm{km}$. By fitting the synthetic SFDs from SPH simulations (Durda et al., 2007), we obtained the value $D_{\rm{PB(SPH)}}\simeq155\,\rm{km}$. \subsection{Hektor --- the first D--type family}\label{subsec:Hektor} Since asteroid (624) Hektor is a close binary with a satellite (Marchis et al. 2014), i.e. an exceptional object, we want to address its association with the family. The cluster around the largest Trojan asteroid appears in the space of proper elements as a relatively compact group, which is limited particularly in proper inclinations, $I_{\rm{p}}\in\langle 18.13^\circ;19.77^\circ\rangle$, and with resonant semimajor axes located in the interval $a_{\rm{p}}\in\langle 5.234;5.336\rangle$ au. The number of members of this group slowly increases with increasing cutoff velocity up to $v_{\rm{cutoff}} \simeq 110\,\rm{m}\,\rm{s^{-1}}$, above which it quickly joins the background. With our randombox method, we estimated the probability that the family is just a random fluke to be as low as $P_{\rm{rnd}}\simeq2\cdot10^{-3}$. The nominal diameter of asteroid (624) Hektor derived from its albedo is 164 km (Grav et al., 2012), but the albedo measured by AKARI $p_V = 0.034 \pm 0.001$ (Usui et al., 2011) totally differs from that measured by WISE, $p_V = 0.087 \pm 0.016$. and these values do not match even within the error limits. This may be caused by applying a thermal model assuming spheres to the bilobed shape of the asteroid (Marchis et al., 2014). We hence do not determine Hektor's diameter from its albedo, but from fits of Marchis et al. (2014), which effective value $D = (250 \pm 26)\,\rm km$ is suitable within its uncertainty for all possible geometries (convex, bilobe and binary). For other bodies in family we use a nominal value $p_{\rm V}=0.072$, which is the median of WISE measurements. Asteroid (624) Hektor is often classified as D-type (e.g. Cruikshank et al., 2001, Emery et al., 2006, Emery et al., 2011). We tried to evaluate taxonomical classification of other family members and we have found colours for two more expected family members in SDSS-MOC vers. 4 (Ivezic et al., 2002): asteroids (65000) 2002 AV63 and (163702) 2003 FR72. Even though the photometric noise in individual bands is not negligible ($\sigma_i = 0.02\,{\rm mag}$ up to $\sigma_u = 0.12\,{\rm mag}$) both of them are D-types, with principal components (aka slopes) ${\rm PC}_1 > 0.3$. This seems to support the D-type classification of the whole family. We also tried to constrain the taxonomic classification of the family members by comparing their infrared albedos $p_{\rm IR}$ and visual albedos $p_{\rm V}$ as described in Mainzer et al. (2011), but there are no data for family members in the W1 or W2 band of the WISE sample, which are dominated by reflected radiation. The fact that we observe a collisional family associated with a D--type asteroid is the main reason we use word ``exceptional'' in connection with the Hektor family. As we claimed in Bro\v{z} et al. (2013), in all regions containing a mixture of C--type and D--type asteroids (e.g. Trojans, Hildas, Cybeles), there have been only C--type families observed so far, which could indicate that disruptions of D--type asteroids leave no family behind, as suggested by Levison et al. (2009). Nevertheless, our classification of the Hektor family as D--type is not in direct contradiction with this conclusion, because Levison et al. (2009) were concerned with catastrophic disruptions, while we conclude below that the Hektor family originated from a cratering event, i.e. by an impactor with kinetic energy too small to disrupt the parent body. \subsubsection{Simulations of long-term dynamical evolution}\label{sub:dyn_evol} To get an upper limit of the age of the Hektor family, we simulated a long-term evolution of \textit{seven synthetic families} created for different breakup geometries. Our model included four giant planets on current orbits, integrated by the symplectic integrator SWIFT (Levison and Duncan, 1994), modified according to Laskar and Robutel (2001), with the time step of $\Delta t = 91\,\rm days$ and time span 4 Gyr. We also accounted for the Yarkovsky effect in our simulations. Although in a first-order theory, it is not effective in zero-order resonances (it could just shift libration centre, but there is no systematic drift in semimajor axis) and the observed evolution of proper elements is mainly due to chaotic diffusion, in higher-order theories the Yarkovsky effect can play some role. In our model, we assumed a random distribution of spins and rotation periods (typically several hours), the bulk and surface density $\rho_{\rm{bulk}} = \rho_{\rm{surf}} = 1.3\,\,\rm{g}\,\rm{cm^{-3}}$, the thermal conductivity $K = 0.01\,\,\rm{W}\,\rm{m^{-1}}\,\rm{K^{-1}}$, the specific heat capacity $C = 680\,\,\rm{J}\,\rm{kg^{-1}}\,\rm{K^{-1}}$, the Bond albedo $A_{\rm B} = 0.02$ and the IR emissivity $\epsilon = 0.95$. We created each synthetic family by assigning random velocities to 234 bodies (i.e. 3 times more than the number of the observed family members), assuming an isotropic velocity field with a typical velocity of $70\,\,\rm{m}\,\rm{s^{-1}}$, corresponding to the escape velocity from parent body (Farinella et al., 1993). Here we assumed the velocity of fragments to be size independent. Possible trends in the ejection velocity field cannot be easily revealed in the ($a, H$) space in the case of the Hektor family, because of its origin by a cratering event -- there is a large gap in the range between absolute magnitude of (624) Hektor ($H = 7.20$) and other bodies ($H > 11.9$), so we are not able to distinguish a simple Gaussian dispersion from the physical dependence (cf. Carruba et al. 2016). Either way, we are interested in the orbital distribution of \textit{mostly} small bodies. Our assumption of size-independent ejection velocity is also in good agreement with results of SPH models (see Subsection \ref{subsec:velocity_fields} and Figure \ref{fig:SPH_v_evol}). To create a synthetic family in the same position as occupied by the observed Hektor family, we integrated the orbit of asteroid (624) Hektor with osculating elements taken from AstOrb catalogue (Bowell et al., 2002), until we got appropriate values of the true anomaly $f$ and the argument of pericentre $\omega$. We tried values of $f$ ranging from $0^\circ$ to $180^\circ$ with the step of $30^\circ$ and $\omega$ always satisfying the condition $f + \omega = 60^\circ$, i.e. we fixed the angular distance from the node to ensure a comparably large perturbations in inclinations. Initial positions of synthetic families members just after the disruption, compared to the observed Hektor family, are shown in Figure \ref{Hektor_sim_init}. To make a quantitative comparison of the distribution in the space of proper elements, we used a two-dimensional Kolmogorov--Smirnov test to compute KS distance of the synthetic family to the observed one with the output timestep of 1 Myr. The results for different initial geometries are shown in Figure \ref{KS}. \begin{figure*} \centering \renewcommand{\tabcolsep}{0pt} \begin{tabular}{ccc} \\ \includegraphics[width=6cm]{figures/H_obs_ae.eps} & \includegraphics[width=6cm]{figures/H2_ae.eps} & \includegraphics[width=6cm]{figures/H4_ae.eps} \\ \includegraphics[width=6cm]{figures/H5_ae.eps} & \includegraphics[width=6cm]{figures/H6_ae.eps} & \includegraphics[width=6cm]{figures/H8_ae.eps} \\ \end{tabular} \caption{Initial conditions for simulations of long-term evolution of synthetic families (red), compared to the observed Hektor family (blue) in the space of proper elements $(a_{\rm p},e_{\rm p})$. Each figure shows a different disruption geometry with different values of the true anomaly $f$ and the argument of pericentre $\omega$. Note the shaded area in the top left figure -- there are only two observed asteroids with proper semimajor axis $a_{\rm p} > 5.32\,\rm au$. This is due to the proximity to the border of the stable librating region. As there are many more synthetic asteroids in this region in all cases of initial distributions, we need to simulate a dynamical evolution of the family.} \label{Hektor_sim_init} \end{figure*} Our two best fits corresponding to the lowest KS distance are displayed in Figure \ref{Hektor_best_fit}. As we can see from the image of the whole Trojan L4 population, Hektor seems to be near the outskirts of the librating region (cf. Figure \ref{trojans_L4_arer_sizes}). In Figure~\ref{Hektor_sim_init}, we can note, that there are almost no observed asteroids in the shaded area with $a_{\rm p}> 5.32\,\rm{au}$, but we can see some synthetic family members in the left panel of Figure \ref{Hektor_best_fit} (initial geometry $f=0^\circ$, $\omega = 60^\circ$). On the other hand, when we look at right panel of Figure \ref{Hektor_best_fit} (initial geometry $f=150^\circ$, $\omega = 270^\circ$), we can see that there are many fewer bodies in the proximity of the border of the stable librating region. One can also see the initial ``fibre-like'' structure is still visible on the left, but is almost dispersed on the right. Hence, we conclude that the geometry at which the disruption occurred is rather $f=150^\circ$, $\omega = 270^\circ$ and the corresponding age is between 1 and 4 Gyr. The second but less likely possibility is that the disruption could have occured more recently (0.1 to 2.5 Gyr) at $f=0^\circ$, $\omega = 60^\circ$. \subsubsection{Parent body size from SPH simulations}\label{subsec:Durda_fit} We tried to estimate the parent body size of Hektor family and other families by the method described in Durda et al. (2007). To this point, we calculated a pseudo$-\chi^2$ for the whole set of synthetic size-frequency distributions as given by the SPH simulations results (see Figure \ref{Hektor_SPH_fit}). Parent body sizes $D_{\rm PB(SPH)}$ and mass ratios of the largest fragment and parent body $M_{\rm{LF}}/M_{\rm{PB}}$ estimated by this method are listed in Table \ref{fams_prop}. The parent body size for Hektor family we derived from SPH simulations is $D_{\rm PB(SPH)} = (260 \pm 10)\,\rm km$, the impactor diameter $D_{\rm imp} = (24 \pm 2)\,\rm km$, the impactor velocity $v_{\rm imp} = (4 \pm 1)\,\rm{km}\,\rm{s^{-1}}$ and the impact angle $\varphi_{\rm imp} = (60^\circ \pm 15^\circ)$. We will use these values as initial conditions for simulations of collisional evolution below. \begin{figure*} \centering \renewcommand{\tabcolsep}{0pt} \begin{tabular}{cc} $(a_{\rm p},e_{\rm p})$ & $(a_{\rm p},I_{\rm p})$ \\ \includegraphics[width=9cm]{figures/KS_mean_d_ae.eps} & \includegraphics[width=9cm]{figures/KS_mean_d_ai.eps} \\ \end{tabular} \caption{Kolmogorov--Smirnov distance $D_{\rm KS}$ vs time for 7 different synthetic families compared with the observed Hektor family. Two-dimensional KS test was computed for the distributions of synthetic and observed families in the space of proper elements $(a_{\rm p},e_{\rm p})$ (left) and $(a_{\rm p},I_{\rm p})$ (right). The synthetic families were created assuming different impact geometries, namely the true anomaly $f=0^\circ,30^\circ,60^\circ, 90^\circ, 120^\circ, 150^\circ, 180^\circ $ and the argument of pericenter $\omega = 60^\circ, 30^\circ, 0^\circ, 330^\circ, 300^\circ, 270^\circ, 240^\circ$, which were combined so that the sum $f + \omega = 60^\circ$. The averaged distance $D_{\rm KS}$ changes in the course of dynamical evolution and we can see two minima: for $f=0^\circ$ and $\omega=60^\circ$ (red curve) it is at about ($350 \pm 100$) Myr; for $f=150^\circ$ and $\omega=270^\circ$ (yellow curve) there is a flat minimum at ($2800 \pm 1500$) Myr. Since the red and yellow curves are overlapping in the range from 1800~Myr to 2500~Myr, we adopt the values of possible ages as 100 to 2500~Myr for the $f=0^\circ$ and $\omega=60^\circ$ geometry (red curve) and 1000 to 4000~Myr for the $f=150^\circ$ and $\omega=270^\circ$ geometry (yellow curve).} \label{KS} \end{figure*} \begin{figure*} \centering \begin{tabular}{cc} \includegraphics[width=8.5cm]{figures/H2_obs_364.eps} & \includegraphics[width=8.5cm]{figures/H8_obs_3100.eps} \\ \end{tabular} \caption{Two evolved synthetic families in the space of proper elements $(a_{\rm p},e_{\rm p})$, which correspond to the minima of KS distance in Figure \ref{KS}. Left picture shows the synthetic family (red) with $f=0^\circ$ and $\omega=60^\circ$ after 364 Myr of evolution in comparison with the observed Hektor family (blue). Right picture corresponds to the synthetic family with $f=150^\circ$ and $\omega=270^\circ$ after 3100 Myr of evolution. These two pictures differ in fine details, which cannot be accounted for in the KS statistics: i) the ``fibre-like'' structure of the relatively young family is still visible in the left picture; ii) there are many fewer synthetic bodies in the shaded area of the right picture $(a_{\rm p}>5.32\,\rm au)$ than on the left, which is closer to the observed reality.} \label{Hektor_best_fit} \end{figure*} \subsection{1996\,RJ --- extremely compact family} In our previous work, we mentioned a small cluster associated with asteroid (9799) 1996~RJ, which consisted of just 9~bodies. With the contemporary sample of resonant elements we can confirm that this cluster is indeed visible. It is composed of 18 bodies situated near the edge of the librating zone on high inclinations, within the ranges $I_{\rm p} \in \langle 31.38^\circ ; 32.27^\circ \rangle$ and $a_{\rm{p}}\in\langle 5.225\,;\,5.238\rangle$~au. As it is detached from the background in the space of proper elements, it remains isolated even at high cutoff velocity $v_{\rm{cutoff}}=160\, \rm{m\,s^{-1}}$. Unfortunately, we have albedos measured by WISE for just 4 members of this family. These albedos are not much dispersed. They range from $p_{\rm{V}} = 0.079 \pm 0.019$ to $p_{\rm{V}} = 0.109 \pm 0.029$ and, compared to the median albedo of the whole $L_4$ population $\widetilde {p_{\rm{V}}} = 0.072 \pm 0.017$, they seem to be a bit brighter, but this statement is a bit inconclusive. \begin{figure} \centering \includegraphics[width=7.0cm]{figures/Hektor_002_Basalt_4_60_22sfd.eps} \caption{Our best-fit size-frequency distribution of Hektor family by scaled SFDs from SPH simulations of Durda et al. (2007). In this particular case $D_{\rm PB(SPH)} = 257\,\rm km$, impactor diameter $D_{\rm imp} = 48\,\rm km$, impactor velocity $v_{\rm imp} = 4\,\rm{km}\,\rm{s^{-1}}$ and impact angle $\varphi_{\rm imp} = 60^\circ$. However, other fits with similar pseudo-$\chi^2$ suggest the uncertainties are as follows: $\Delta D_{\rm PB(SPH)} = 10\,\rm km$, $\Delta D_{\rm imp} = 2\,\rm km$, $\Delta v_{\rm imp} = 1\,\rm{km}\,\rm{s^{-1}}$ and $\Delta \varphi_{\rm imp} = 15^\circ$. SFD shape seems to be more dependent on impact geometry than on impact velocity.} \label{Hektor_SPH_fit} \end{figure} \subsection{Arkesilaos} This family is located on low inclinations $I_{\rm p} \in \langle 8.52^\circ ; 9.20^\circ \rangle$, in the range of $a_{\rm{p}}\in\langle 5.230\,;\,5.304\rangle$~au. It is clearly visible in the space of proper elements, although this area of $L_4$ cloud is very dense. Still, it is difficult to find the largest remnant of the parent body, because this region is populated mainly by small asteroids with absolute magnitudes $H>12$. The only four asteroids with $H<12$ are (2148) Epeios with $H=10.7$, (19725)~$1999\,\rm WT_4$ with $H=10.7$, (38600)~$1999\,\rm XR_{213}$ with $H=11.7$ and (20961)~Arkesilaos with $H=11.8$. The only diameter derived from measured albedo is that of (2148) Epeios, which is $D = (39.02 \pm 0.65)\,\rm km$. Diameters of remaining bodies were calculated from their absolute magnitude assuming albedo $p_{\rm V} = 0.072$, which is the median of $L_4$ Trojans. Although (20961) Arkesilaos has the diameter only $D = (24 \pm 5)\,\rm km$, it is the only asteroid with $H<12$, for which the associated family has a reasonable number of members $N_{\rm memb}$ even for small values of the cutoff velocity $v_{\rm cutoff}$ (see Section \ref{sec:HCM}). As this is also the only larger body located near the center of the family in the space of proper elements, we treat (20961) Arkesilaos as the largest remnant of the parent body, whose diameter we estimate to be $D_{\rm PB(SPH)} \simeq 87\,\rm km $. Given that the mass ratio of the largest remnant and the parent body, as derived from SPH simulations of Durda et al. (2007), is $M_{\rm LR}/M_{\rm PB} \simeq 0.02$ only, it seems this family inevitably originated from a catastrophic disruption. \subsection{Ennomos} In our previous work, we reported a discovery of a possible family associated with asteroid (4709) Ennomos. With new data, we can still confirm that there is a significant cluster near this body, but when we take into account our ``$N_{\rm{memb}}(v_{\rm{cutoff}})$'' criterion described above, it turns out that the family is rather associated with asteroid (17492) Hippasos. It is a relatively numerous group composed of almost 100~bodies, situated near the border of the stable librating zone $L_5$ at high inclinations, ranging from $I_{\rm p} \in\langle 26.86^\circ;30.97^\circ \rangle$, and $a_{\rm{p}}\in\langle 5.225; 5.338\rangle$~au. \subsection{2001\,UV$_{209}$} Using new data, we discovered a ``new'' family around asteroid (247341) $2001\,\rm UV_{209}$, which is the second and apparently the last observable family in our sample. Similar to the Ennomos family, it is located near the border of the $L_5$ zone on high inclinations $I_{\rm p} \in \langle 24.02^\circ;26.56^\circ \rangle$ and $a_{\rm{p}}\in\langle 5.218;5.320\rangle$~au. This family has an exceptionally steep slope of the SFD, with $\gamma=-8.6 \pm 0.9$, which may indicate a recent collisional origin or a disruption at the boundary of the libration zone, which may be indeed size-selective as explained in Chrenko et al. (2015). \section{Collisional models of the Trojan population}\label{sec:coll_model} In order to estimate the number of collisional families among $L_4$ Trojans, we performed a set of 100 simulations of the collisional evolution of Trojans with the Boulder code (Morbidelli et al., 2009) with the same initial conditions, but with different values of the random seed. \subsection{Initial conditions} We set our initial conditions of the simulations such that 4 Gyr of collisional evolution leads to the observed cumulative SFD of $L_4$~Trojans (red curve in Figure \ref{fig:Boulder_SFD}). We constructed the initial synthetic SFD as three power laws with the slopes $\gamma_{\rm a} = -6.60$ in the size range from $D_1=117\,\rm{km}$ to $D_{\rm max}=250\,\rm{km}$, $\gamma_{\rm b} = -3.05$ from $D_2=25\,\rm{km}$ to $D_1$ and $\gamma_{\rm c} = -3.70$ from $D_{\rm min}=0.05\,\rm{km}$ to $D_2$. The synthetic initial population was normalized to contain $N_{\rm norm}=11$ asteroids with diameters $D\geq D_1$. To calculate the target strength $Q_D^*$, we used a parametric formula of Benz and Asphaug (1999): \begin{equation}\label{Q*} Q_D^*=Q_0 R_{\rm PB}^a + B\rho_{\rm bulk}R_{\rm PB}^b, \end{equation} where $R_{\rm PB}$ is the parent body radius in centimetres, $\rho_{\rm bulk}$ its bulk density, which we set to be $\rho_{\rm bulk} = 1.3\,\,\rm{g}\,\rm{cm^{-3}}$ for synthetic Trojans (cf. Marchis et al., 2014). As of constants $a,\,b,\,B\,\rm{and}\,Q_0$ we used the values determined by Benz and Asphaug (1999) for ice at the impact velocity $v_{\rm imp}=3\,\rm km\,s^{-1}$, which are: $a=-0.39$, $b=1.26$, $B=1.2\,\,\rm erg\,cm^3\,g^{-2}$ and $Q_0=1.6\cdot10^7\,\rm erg\,g^{-1}$. In our model, we take into account only Trojan vs Trojan collisions, as the Trojan region is practically detached from the main belt. Anyway, main-belt asteroids with eccentricities large enough to reach the Trojan region are usually scattered by Jupiter on a time scale significantly shorter than the average time needed to collide with a relatively large Trojan asteroid. We thus assumed the values of collisional probability $P_{\rm i}=7.80\cdot 10^{-18}\,\,\rm km^{-2}\,yr^{-1}$ and the impact velocity $v_{\rm imp}=4.66\,\,\rm km\,s^{-1}$ (Dell'Oro et al., 1998). Unfortunately, Benz and Asphaug (1999) do not provide parameters for ice at the impact velocities $v_{\rm imp} > 3\,\rm km\,s^{-1}$. We also ran several simulations with appropriate values for basalt at impact velocity $v_{\rm imp}=5\,\rm km\,s^{-1}$ ($a=-0.36$, $b=1.36$, $B=0.5\,\,\rm erg\,cm^3\,g^{-2}$ and $Q_0=9\cdot10^7\,\rm erg\,g^{-1}$). Both models qualitatively exhibit the same evolution of SFD and they give approximately the same total numbers of disruptions and craterings occured, but for basalt the model gives three times fewer \textit{observable} families originated by cratering than for ice. The results for the ice match the observation better, so we will further discuss the results for ice only. \subsection{Long-term collisional evolution} The results of our simulations of the collisional evolution are shown in Figure \ref{fig:Boulder_SFD}. Our collisional model shows only little changes above $D>50\,\rm{km}$ over the last 3.85 Gyr (i.e. post-LHB phase only). Slopes of the initial synthetic population and the observed $L_4$ population differ by $\Delta \gamma < 0.1$ in the size range from 50~km to 100~km, while a relative decrease of the number of asteroids after 3.85 Gyr of collisional evolution is only about $12\,\%$ in the same size range. Hence, we can consider this part of the Trojan population as a representative sample of the source population, which is not much affected by collisional evolution. Therefore, these Trojans provide very useful information about the source population, from which they were captured (as modeled in Nesvorn\'y et al., 2013). \begin{figure} \centering \includegraphics[width=8.5cm]{figures/sfd_4000.eps} \caption{Simulations of the collisional evolution of $L_4$ Trojans with the Boulder code (Morbidelli et al., 2009). Shown here is the initial cumulative SFD of a synthetic population (black) and the SFD of the observed one (red). Green are the final SFDs of 100 synthetic populations with the same initial SFD but with different random seeds, after 4 Gyr of a collisional evolution. The evolution of bodies larger than $D> 50\,\rm{km}$ is very slow, hence we can consider this part of the SFD as captured population.} \label{fig:Boulder_SFD} \end{figure} \begin{figure*} \centering \renewcommand{\tabcolsep}{0pt} \begin{tabular}{cc} \includegraphics[width=9cm]{figures/Boulder_families_new_disrupt.eps} & \includegraphics[width=9cm]{figures/Boulder_families_new_cratering.eps} \\ \end{tabular} \caption{The dependence of the cumulative number (an average over 100 simulations) of catastrophic disruptions among Trojans (left panel) and cratering events (right panel) on the target diameter $D_{\rm PB}$ (black boxes), and a subset of those Trojan families, which should be detected in contemporary observational data, i.e. with the number of fragments $N(D>10\,\rm{km})>10$ (green boxes for disruptions and blue boxes for craterings). In other words, colour boxes represent simulated detections of families based on the expected effectiveness of our detection methods. This is the reason, why the cumulative number of the observable families does not strictly increase with the decreasing parent body size, but is rather constant under the limit of about 95~km in the case of catastrophic disruptions and 145~km in the case of craterings. There are also observed families marked for a comparison. Three of the four observed families in $L_4$ cloud originated by catastrophic disruption, while only one (Hektor) originated by cratering event (cf. Table \ref{fams_list}).} \label{fig:Boulder_families} \end{figure*} \subsection{An estimate of the number of observable families} From our set of simulations, we also obtained the number of collisions leading to collisional families among $L_4$ Trojans, namely catastrophic disruptions, where the mass ratio of the largest remnant and the parent body $M_{\rm{LR}}/M_{\rm{PB}} < 0.5$, and cratering events, where $M_{\rm{LR}}/M_{\rm{PB}} > 0.5$. As one can verify in Figure \ref{fig:Boulder_families}, these numbers are dependent on the diameter of the parent body $D_{\rm PB}$. However, not all of these collisions produce families which are in fact observable (detectable). There are generally two possible obstacles in the detection of a family in the space of proper elements: i) somewhat more concentrated background population, due to which our detection methods (both ``randombox'' and HCM, see Chapter \ref{sec:groups}) may fail, if the number of observed fragments is too low in comparison with the background, and ii) an observational incompleteness, which means that in the case of Trojans, a substantial part of fragments with sizes $D<10\,\rm km$ is still unknown, what again reduces a chance of a family detection. For these reasons, we constructed a criterion of observability that a synthetic family must fulfill in order to be detectable in the current conditions (i.e. we simulated a detection of synthetic families by the same methods we used to detect the real ones). The simplest criterion could be that a family must contain at least $N_{\rm min} = 10$ fragments with diameter $D\geq 10\,\rm{km}$. Within 100 simulations, there were 93 catastrophic disruptions of bodies with diameters $D_{\rm{PB}}>100\,\rm{km}$, but only 50 of them produced more than 10 fragments with $D\geq10\,\rm{km}$, see Figure \ref{fig:Boulder_families}. Hence, the probability that we would observe a collisional family originated by a catastrophic disruption of a parent body with $D_{\rm{PB}}>100\,\rm{km}$ is \textit{only} 0.50, which matches the observations (namely Eurybates family with $D_{\rm PB(SPH)}\simeq155\,\,\rm km$, see Table \ref{fams_prop}). This value is also roughly consistent with our previous estimate based on the stationary model (Bro\v{z} and Rozehnal, 2011), which gives the value 0.32 with new observational data. As one can also see in Figure \ref{fig:Boulder_families}, the number of cratering events is about one to two orders higher than the number of catastrophic disruptions, however, they do not produce enough fragments larger than $D\geq10\,\rm{km}$. For the parent body size $D_{\rm{PB}}>100\,\rm{km}$ there occurred almost 45,000 cratering events within 100 simulations which produced the largest fragment with $D_{\rm LF}\geq1\,\rm km$, but only 10 of them fulfill our criterion of observability. Hence, the probability that we can observe a family originated by a cratering of a parent body with $D_{\rm{PB}}>100\,\rm{km}$ is only 0.10, at least with contemporary data. From a statistical point of view, this can actually correspond to the Hektor family. As we have already demonstrated in Bro\v{z} and Rozehnal (2011), the number of families is not significantly affected by chaotic diffusion or by a ballistic transport outside the libration zone. \section{SPH simulations of Hektor family}\label{sec:SPH} As we have already mentioned in Section \ref{subsec:Hektor}, (624)~Hektor is very interesting Trojan asteroid with possibly bilobed shape and a small moon. Diameters of (624) Hektor stated in Marchis et al. (2014) are as follows: equivalent diameter $D_{\rm eq}=(250\pm26)\,\,\rm km$ for a convex model, the individual diameters of the lobes $D_{\rm A}=(220\pm22)\,\,\rm km$, $D_{\rm B}=(183\pm18)\,\,\rm km$ for a bilobed version. Estimated parameters of the moon are: the diameter $D_{\rm m}=(12\pm3)\,\,\rm km$, the semimajor axis $a_{\rm m}=(623\pm10)\,\,\rm km$, the eccentricity $e_{\rm m}=(0.31\pm0.03)$ and the inclination (with respect to the primary equator) $I_{\rm m}=(50\pm1)^\circ$. As we associate (624) Hektor with the collisional family, we would like to know, how the properties of the family are influenced by the shape of target body. We therefore performed a series of SPH simulations aiming to explain the origin of the Hektor family, for both cases of convex and bilobed shape of its parent body. \subsection{Methods and initial conditions} We simulated a collisional disruption using the smoothed-particle hydrodynamic code SPH5 (Benz and Asphaug, 1994). We performed two sets of simulations. In the first one, we simulated an impact on a single spherical asteroid. In the second, on a bilobed asteroid represented by two spheres positioned next to each other. The two touching spheres have a narrow interface, so that the SPH quantities do not easily propagate between them. In this setup, we are likely to see differences between sinlge/bilobed cases as clearly as possible. As for the main input parameters (target/impactor sizes, the impact velocity and the impact angle) we took the parameters of our best-fit SFDs, obtained by Durda et al. (2007) scaling method, see Section \ref{subsec:Durda_fit} and Figure \ref{Hektor_SPH_fit}. To simulate a collision between the parent body and the impactor we performed a limited set of simulations: i) a single spherical basalt target with diameter $D_{\rm PB} = 260\,\,\rm km$ vs a basalt impactor with diameter $D_{\rm imp} = 48\,\,\rm km$; ii) the single basalt target $D_{\rm PB} = 260\,\,\rm km$ vs an ice impactor (a mixture of ice and 30\,\% of silicates) with $D_{\rm imp} = 64\,\,\rm km$ (impactor diameter was scaled to get the same kinetic energy); iii) a bilobed basalt target approximated by two spheres with diameters $D_{\rm PB} = 200\,\,\rm km$ each (the total mass is approximately the same) vs a basalt impactor with $D_{\rm imp} = 48\,\,\rm km$; iv) a single spherical ice target $D_{\rm PB} = 260\,\,\rm km$ vs an ice impactor $D_{\rm imp} = 38\,\,\rm km$ (impactor diameter was scaled to get the same ratio of the specific kinetic energy $Q$ to the target strength $Q^*_{\rm D}$). The integration was controlled by the Courant number $C = 1.0$, a typical time step thus was $\Delta t \simeq 10^{-5}\,{\rm s}$, and the time span was $t_{\rm stop} = 100\,{\rm s}$. The Courant condition was the same in different materials, using always the maximum sound speed $c_{\rm s}$ among all SPH particles, as usually. We used $N_{\rm SPH, st}=10^5$~SPH particles for the single spherical target and $N_{\rm SPH, bt}=2\cdot 10^5$ for the bilobed one. For impactor $N_{\rm SPH, i}=10^3$ SPH particles. We assumed the Tillotson equation of state (Tillotson, 1962) and material properties, which are listed in Table \ref{tab:materials}. \begin{table}\vspace{-5mm}\centering \caption{Material constants used in our SPH simulations for basalt and silicated ice (30\,\% of silicates). Listed here are: the zero-pressure density $\rho_0$, bulk modulus $A$, non-linear compressive term $B$, sublimation energy $E_0$, Tillotson parameters $a$, $b$, $\alpha$ and $\beta$, specific energy of incipient vaporization $E_{\rm iv}$, complete vaporization $E_{\rm cv}$, shear modulus $\mu$, plastic yielding $Y$, melt energy $E_{\rm melt}$ and Weibull fracture parameters $k$ and $m$. Values we used for silicated ice are identical to those of pure ice, except density $\rho_0$, bulk modulus $A$ and Weibull parameters $k$ and $m$. All values were adopted from Benz and Asphaug (1999). } \small \renewcommand{\tabcolsep}{6pt} \begin{tabular}{c|c|c|c|} \hline\ quantity & basalt & silicated ice & unit\\ \hline\hline $\rho_0$ & 2.7 & 1.1 & ${\rm g}\,{\rm cm}^{-3}$ \\ $A$ & $2.67\cdot10^{11}$ & $8.44\cdot10^{10}$ & ${\rm erg}\,{\rm cm}^{-3}$ \\ $B$ & $2.67\cdot10^{11}$ & $1.33\cdot10^{11}$ & ${\rm erg}\,{\rm cm}^{-3}$ \\ $E_0$ & $4.87\cdot10^{12}$ & $1.00\cdot10^{11}$ & ${\rm erg}\,{\rm g}^{-1}$ \\ $a$ & 0.5 & 0.3 & -- \\ $b$ & 1.5 & 0.1 & -- \\ $\alpha$ & 5.0 & 10.0 & -- \\ $\beta$ & 5.0 & 5.0 & -- \\ $E_{\rm iv}$ & $4.72\cdot10^{10}$ & $7.73\cdot10^{9}$ & ${\rm erg}\,{\rm g}^{-1}$ \\ $E_{\rm cv}$ & $1.82\cdot10^{11}$ & $3.04\cdot10^{10}$ & ${\rm erg}\,{\rm g}^{-1}$ \\ $\mu$ & $2.27\cdot10^{11}$ & $2.80\cdot10^{10}$ & ${\rm erg}\,{\rm cm}^{-3}$ \\ $Y$ & $3.5\cdot10^{10}$ & $1.0\cdot10^{10}$ & ${\rm erg}\,{\rm g}^{-1}$ \\ $E_{\rm melt}$ & $3.4\cdot10^{10}$ & $7.0\cdot10^{9}$ & ${\rm erg}\,{\rm g}^{-1}$ \\ $k$ & $4.0\cdot10^{29}$ & $5.6\cdot10^{38}$ & ${\rm cm}^{-3}$ \\ $m$ & 9.0 & 9.4 & -- \\ \hline \vspace{2mm} \end{tabular} \label{tab:materials} \end{table} We terminated SPH simulations after 100 s from the impact. This time interval is needed to establish a velocity field of fragments and to complete the fragmentation. Then we handed the output of the SPH simulation as initial conditions to the N--body gravitational code Pkdgrav (Richardson et al., 2000), a parallel tree code used to simulate a gravitational reaccumulation of fragments. Unlike Durda et al. (2007), who calculated radii of fragments $R$ from the smoothing length $h$ as $R=h/3$, we calculated fragments radii from their masses $m$ and densities $\rho$ as $R = (m/(4\pi\rho))^{1/3}$. We ran Pkdgrav with the time step $\Delta t = 5.0\,\,\rm s$ and we terminated this simulation after $t_{\rm evol} = 3$~days of evolution. To ensure this is sufficiently long, we also ran several simulations with $t_{\rm evol} = 5$~days, but we had seen no significant differences between final results. We used the nominal value for the tree opening angle, ${\rm d}\theta = 0.5\,{\rm rad}$, even though for the evolution of eventual moons it would be worth to use even smaller value, e.g. ${\rm d}\theta = 0.2\,{\rm rad}$. \subsection{Resulting size-frequency distributions} \begin{figure} \centering \includegraphics[width=8.5cm]{figures/SPH_SFD_single_bilobe_imp_rmv.eps} \caption{A comparison of size-frequency distributions of the observed Hektor family (red dotted) and SFDs of synthetic families created by different SPH simulations, always assuming the impactor velocity $v_{\rm imp} = 4\,\,\rm km\,s^{-1}$ and the impact angle $\varphi_{\rm imp}~=~60^\circ$. For a single spherical target (green lines) we assumed the diameter $D_{\rm PB} = 260\,\,\rm km$, for a bilobe target (blue line) we approximated the lobes as spheres with diameters $D_{\rm PB} = 200\,\,\rm km$ each. The impactor size was assumed to be $D_{\rm imp} = 48\,\,\rm km$ in the case of basalt, $D_{\rm imp} = 64\,\,\rm km$ in the case of silicate ice impacting on basalt target (scaled to the same $E_{\rm imp}$) and $D_{\rm imp} = 38\,\,\rm km$ in the case of silicate ice impacting on ice target (scaled to the same $Q/Q^*_{\rm D}$). Fragments of the impactor were purposely removed from this plot, as they do not remain in the libration zone for our particular impact orbital geometry.} \label{fig:SPH_SFD} \end{figure} From the output of our simulations we constructed size-frequency distributions of synthetic families, which we compare to the observed one, as demonstrated in Figure \ref{fig:SPH_SFD}. As one can see, there are only minor differences between SFDs of families created by the impacts on the single and bilobed target, except the number of fragments with diameter $D<5\,\rm km$, but this is mostly due to different numbers of SPH particles. However, there are differences between ice and basalt targets. Basalt targets provide generally steeper SFDs with smaller largest remnants than the ice target. To make the comparison of these synthetic initial SFDs to each other more realistic, we removed the fragments of the impactor from our synthetic families. This is because fragments of the impactor often do not remain in the libration zone. Note that this procedure does not subtitute for a full simulation of further evolution; it serves just for a quick comparison of the SFDs. To match the observed SFD of the Hektor family more accurately, we should perform a much larger set of simulations with different sizes of projectiles and also different compositions (mixtures of ice and basalt). However, material parameters of these mixtures are generally not known. Regarding the material constants of pure ice, we have them for the impact velocity $v_{\rm imp} = 3\,\rm km\,s^{-1}$ only (Benz and Aspaugh, 1999). There are also some differences between SFDs of single and bilobe targets, so we should perform these simulations for each target geometry. However, we postpone these detailed simulations for future work; in this work we further analyse results of simulations with basalt targets and we focus on the evolution of the SFDs. It should be emphasized that the SFDs presented here correspond to very young synthetic families, hence they are not affected by any dynamical and collisional evolution yet. To reveal possible trends of the evolution by a ballistic transport and chaotic diffusion, we prepared initial conditions for the SWIFT integrator, similarly as described in Section \ref{sub:dyn_evol}, let the simulation run and monitored the corresponding evolution of the SFD. The results can be seen in Figure~\ref{fig:SPH_SFD_evol}. The biggest difference between $t=0$ and $t=1\,\,\rm Myr$ is caused by a ballistic transport outside the libration zone --- fragments (especially of the impactor) missing from the SFD at $t=1\,\,\rm Myr$ were perturbed too much to remain in the libration zone, at least for a given impact geometry. We actually tested two impact geometries: in the direction tangential and perpendicular to the orbit. \begin{figure} \centering \includegraphics[width=8.5cm]{figures/SPH_SFD_bilobe_evol.eps} \caption{A simulation of evolution of the SFD of a synthetic Hektor family due to a ballistic transport and chaotic diffusion. One can see here a rapid change of SFD within the first 1 Myr after the breakup as the fragments of the impactor leaved the libration zone in our impact geometry. This ballistic transport resulted in a reduction of the number of particularly larger bodies in our case. Further evolution due to the chaotic diffusion seems to cause the reduction of mostly smaller bodies. Note that the initial SFD (0 Myr) contains some fragments of the impactor, so the blue solid curve looks different than the curve in Figure \ref{fig:SPH_SFD}, where the fragments of the impactor were removed.} \label{fig:SPH_SFD_evol} \end{figure} This may be important for the method we used in Section \ref{subsec:Durda_fit} to derive a preliminary parent body size and other properties of the family. The SFDs obtained by Durda et al. (2007) were directly compared in their work to the main-belt families, however, there is a part of fragments among Trojans (in our case even the largest ones, see Figure \ref{fig:SPH_SFD_evol}), which cannot be seen in the space of resonant elements, because they do not belong to Trojans any more. Fortunately, values of pseudo-$\chi^2$ we computed in Section \ref{subsec:Durda_fit} depend rather weakly on the distribution of a few largest bodies. Even so, we plan to analyze SFDs of synthetic families more carefully in future works. \begin{figure} \centering \includegraphics[width=8.5cm]{figures/hist_velocity.eps} \caption{Velocity distributions of fragments originated in various SPH simulations (green, violet) in comparison with the model of Farinella et al. (1994) we used in our $N$-body simulations of isotropic disruption and dynamical evolution (see Section \ref{sub:dyn_evol}). Shown here is also the distribution of velocities after 1 Myr of evolution, i.e. of fragments that remained in libration zones.} \label{fig:SPH_v_evol} \end{figure} \subsection{Resulting velocity fields}\label{subsec:velocity_fields} In our $N$-body simulations, we used the model of isotropic disruption (Farinella et al., 1994). As we compared the synthetic family with the observed one (see Section \ref{sub:dyn_evol}), we simulated only the evolution of bodies with relatively low ejection velocities ($v<200\,\rm m\,s^{-1}$), because the observed family is confined by the cutoff velocity $v_{\rm cutoff} = 110\,\rm m\,s^{-1}$. Very small fragments with higher velocities may be still hidden in the background. Here, we compare Farinella's model to the velocity fields of fragments from SPH simulations, see Figure \ref{fig:SPH_v_evol}. We realized that Farinella's model is not offset substantially with respect to other velocity histograms, especially at lower velocities, $v < 200\,{\rm m}\,{\rm s}^{-1}$. On the other side, there remained some fragments of the impactor with velocities $v>2\,\rm km\,s^{-1}$ in our SPH simulations, which are not produced in the isotropic model. It does not affect a comparison of the synthetic and observed families in the space of proper elements, as these high-velocity fragments leaved the Trojan region in our case, but it does affect the SFD of the synthetic family. As a consequence, one should always analyse SFDs and velocity fields together. We also simulated a further evolution of the velocity field. After just 1 Myr of evolution, there remained no bodies with $v>1.5\,\rm km\,s^{-1}$ in our impact geometries, and as one can see in Figure \ref{fig:SPH_v_evol}, there was a rapid decrease in the number of fragments with initial $v>300\,\rm m\,s^{-1}$. The resulting histogram is again similar to that of the simple isotropic model. \subsection{Synthetic moons} \begin{table}\vspace{-5mm}\centering \caption{A comparison of the sizes and the orbital parameters (i.e. semimajor axis $a$, eccentricity $e$ and period $P$) of the observed moon of (624) Hektor as listed in Marchis et al. (2014), with the parameters of synthetic moons SPH I and SPH II captured in our SPH simulation of impact on the bilobed target.} \small \renewcommand{\tabcolsep}{3pt} \begin{tabular}{c|c|c|c|c|} \hline\ desig. & diam. [km] & $a$ [km] & $e$ & $P$ [days] \\ \hline\hline observed & $12\pm3$ & $623.5\pm10$ & $0.31\pm0.03$ & $2.9651\pm0.0003$ \\ \hline SPH I & 2.2 & 715 & 0.82 & 1.2 \\ SPH II & 2.7 & 370 & 0.64 & 0.4 \\ \hline \vspace{2mm} \end{tabular} \label{tab:moons} \end{table} In our simulation of the impact of basalt projectile on the bilobe-shape basalt target, we spotted two low-velocity fragments with original velocities $130\,\rm m\,s^{-1}$ and $125\,\rm m\,s^{-1}$, which were consequently captured as moons of the largest remnant. Their sizes and orbital parameters are listed in Table \ref{tab:moons}. These satellites were captured on orbits with high eccentricities ($e=0.82$ and 0.64 respectively), which are much higher than the eccentricity of the observed moon determined by Marchis et al. (2014) ($e=0.31\pm0.03$). However, this could be partly caused by the fact, that we handed the output of (gravity free) SPH simulations to the gravitational N-body code after first 100~s. Hence, fragments leaving the parent body could move freely without slowing down by gravity. More importantly, we do not account for any long-term dynamical evolution of the moons (e.g. by tides or binary YORP). When compared to the observed satellite, the diameters of the synthetic moons are several times smaller. This is not too surprising, given that the results for satellite formation are at the small end of what can be estimated with our techniques (median smoothing length $h = 2.3\,{\rm km}$; satellite radius $r \simeq 1.2\,{\rm km}$). The size of captured fragments could also be dependent on impact conditions as different impact angles, impactor velocities and sizes (as is the case for scenarios of Moon formation) which we will analyze in detail in the future and study with more focused simulations. \begin{figure*} \centering \begin{tabular}{cc} \\ \includegraphics[width=9cm]{figures/Hektor_single_basalt_080.eps} & \includegraphics[width=9cm]{figures/Hektor_bilobe_basalt_080.eps} \\ \end{tabular} \caption{A comparison of SPH simulations of a disruption of a single body (basalt) with diameter $D_{\rm target}=250\,\rm km$, by an impactor with the diameter $D_{\rm imp}=48\,\rm km$ (silicate ice) (\textit{left}) and a disruption of a bilobe basalt target, with $D_{\rm target}=198\,\rm km$ for each sphere, by an impactor with $D_{\rm imp}=46\,\rm km$ (silicate ice) (\textit{right}). Time elapsed is $t=80.1\,\rm{s}$ in both cases. There are notable physical differences between the two simulations, especially in the propagation of the shock wave, which is reflected from free surfaces, the number of secondary impacts, or the fragmentation (damage) of the target. Nevertheless, the amount of ejected material and the resulting size-frequency distributions do not differ that much (cf. Figure \ref{fig:SPH_SFD}).} \label{fig:SPH} \end{figure*} \section{Conclusions}\label{sec:conclusions} In this paper, we updated the list of Trojans and their proper elements, what allowed us to update parameters of Trojan families and to discover a new one (namely $2001\,\rm UV_{209}$ in $L_5$ population). We focused on the Hektor family, which seems the most interesting due to the bilobed shape of the largest remnant with a small moon and also its D-type taxonomical classification, which is unique among the collisional families observed so far. At the current stage of knowledge, it seems to us there are no major inconsistencies among the observed number of Trojan families and their dynamical and collisional evolution, at least in the current environment. As usual, we ``desperately'' need new observational data, namely in the size range from 5 to 10 km, which would enable us to constrain the ages of asteroid families on the basis of collisional modeling and to decide between two proposed ages of Hektor family, 1~to~4~Gyr or 0.1~to~2.5~Gyr. As expected, there are qualitative differences in impacts on single and bilobed targets. In our setup, the shockwave does not propagate easily into the secondary, so that only one half the mass is totally damaged as one can see in Figure \ref{fig:SPH}. On the other hand, the resulting SFDs are not that different, as we would expect. Even so, there is a large parameter space, which is still not investigated (i.e. the impact geometry with respect to the secondary, secondary impacts, the position in the orbit). SPH simulations of impacts on bilobed or binary targets thus seem very worthy for future research. Our work is also a strong motivation for research of disruptions of weak bodies (e.g. comets), better understanding the cometary disruption scaling law and also for experimental determination of material constants, which appear in the respective equation of state. As a curiosity, we can also think of searching for the remaining projectile, which could be still present among Trojans on a trajectory substantially different from that of family. A substantial part of projectile momentum is preserved in our simulations, so we may turn the logic and we may assume the projectile most likely came from the Trojan region and then it should remain in this region too. \section*{Acknowledgements} We thank Alessandro Morbidelli for his review which helped to improve the final version of the paper. The work of MB was supported by the grant no. P209/13/01308S and that of JR by P209/15/04816S of the Czech Science Foundation (GA CR). We acknowledge the usage of computers of the Stefanik Observatory, Prague.
d0dae4d92d1c1b80086c4a63956ab01445da7eee
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:intro} Discrete breathers are intrinsic localized modes that can emerge in networks of coupled nonlinear oscillators \cite{breather_reviews,Aubry}. They have been observed not only in Hamiltonian lattices but also in driven dissipative systems under certain conditions. Discrete breathers have been theoretically predicted or experimentally generated in a wide variety of physical systems such as Josephson junction arrays \cite{JJ}, coupled pendula chains \cite{pendula}, micro- and macro-mechanical cantilever arrays \cite{cantilever}, granular crystals \cite{granular}, nonlinear electrical lattices \cite{electric}, and double-strand DNA models \cite{DNA}, just to cite a few instances. Up to now, breathers have been mainly studied for the case of a harmonic external excitation, while various types of periodic excitations are in principle possible, depending upon the physical context under consideration. Since there are infinitely many different wave forms, a quite natural question is to ask how the generation and dynamics of breathers are affected by the presence of a generic periodic excitation. In this present work, we show that a relevant quantity properly characterizing the effectiveness of zero-mean periodic excitations $F(t)$ having equidistant zeros at controlling the generation and dynamics of discrete breathers is the \textit{impulse} transmitted by the external excitation over a half-period (hereafter referred to simply as the excitation's impulse \cite{1}, $I\equiv\int_{0}^{T/2}F(t)dt$, $T$ being the period)$-$ a quantity integrating the conjoint effects of the excitation's amplitude, period, and waveform. It is worth mentioning that the relevance of the excitation's impulse has been observed previously in quite different contexts, such as ratchet transport \cite{2}, adiabatically ac driven periodic (Hamiltonian) systems \cite{3}, driven two-level systems and periodically curved waveguide arrays \cite{4}, chaotic dynamics of a pump-modulation Nd:YVO$_{4}$ laser \cite{5}, topological amplification effects in scale-free networks of signaling devices \cite{6}, and controlling chaos in starlike networks of dissipative nonlinear oscillators \cite{7}. The rest of this paper is organized as follows. In Sec. II we introduce the model and further comment on some of its main features. Section III provides numerical evidence for the essential role of the excitation's impulse at generating breathers and controlling their stability for the prototypical cases of a hard $\phi^{4}$ potential and a sine-Gordon potential. A theoretical explanation of the effectiveness of the excitation's impulse in terms of energy and action is provided in Sec. IV. Finally, Sec. V is devoted to a discussion of the major findings and of some open problems. \section{Model system} \label{sec:model} The discrete nonlinear Klein--Gordon equation with linear coupling, linear damping, and external periodic excitation is one of the simplest equations where dissipative discrete breathers may arise:% \begin{equation} \ddot{u}_{n}+\alpha\dot{u}_{n}+V^{\prime}(u_{n})+C(2u_{n}-u_{n+1}% -u_{n-1})=F(t), \label{eq:dyn}% \end{equation} in which $V(u_{n})$ is an on--site (substrate) potential, $\alpha$ is the damping constant, $C$ is the coupling constant, while $F(t)$ is a zero-mean periodic excitation. \begin{equation} F(t)=(-1)^{hn}f_{0}f_{1,2}(t), \label{eq:driving}% \end{equation} { }in which $f_{0}$ is the driving amplitude, $h$ is a hardness parameter whose value is $0$ ($1$) when the on-site potential is soft (hard), while $f_{1}(t)$, $f_{2}(t)$ are two different periodic excitations that we will use as illustrative examples to show that the impulse is the relevant quantity controlling the effect of the external excitation on the generation and stability properties of breathers. These periodic excitations are given by \begin{equation} f_{1}(t)=N(m)\mathrm{sn}\left( \frac{2K(m)\omega_{\mathrm{b}}t}{\pi };m\right) \mathrm{dn}\left( \frac{2K(m)\omega_{\mathrm{b}}t}{\pi};m\right) , \label{eq:driving1}% \end{equation}% \begin{equation} f_{2}(t)=\mathrm{sn}\left( \frac{2K(m)\omega_{\mathrm{b}}t}{\pi};m\right) , \label{eq:driving2}% \end{equation} where $\mathrm{sn}(\cdot)\equiv\mathrm{sn}(\cdot;m)$ and $\mathrm{dn}% (\cdot)\equiv\mathrm{dn}(\cdot;m)$ are Jacobian elliptic functions of parameter $m$ [$K\equiv K(m)$ is the complete elliptic integral of the first kind], while $N(m)$ is a normalization function which is introduced for the elliptic excitation $f_{1}(t)$ to have the same amplitude $f_{0}$ and period $T\equiv2\pi/\omega_{\mathrm{b}}$ for any wave form (i.e., $\forall m\in\left[ 0,1\right) $ ). Specifically, the normalization function is given by \begin{equation} N(m)=\left[ \eta_{1}+\frac{\eta_{2}}{1-\exp[(m-\eta_{3})/\eta_{4}]}\right] ^{-1}, \end{equation} with $\eta_{1}=0.43932$, $\eta_{2}=0.69796$, $\eta_{3}=0.37270$, $\eta _{4}=0.26883$. In both excitations $f_{1,2}(t)$, when $m=0$, then $F(t)=(-1)^{hn}f_{0}\sin\left( \omega_{\mathrm{b}}t\right) $, that is, one recovers the standard case of a harmonic excitation \cite{Cubero}, whereas, for the limiting value $m=1$, the excitation $f_{1}(t)$ vanishes while the excitation $f_{2}(t)$ reduces to a square wave. It is worth noting that the excitations $f_{1,2}(t)$ have been chosen to exhibit the following properties. For the excitation $f_{1}(t)$, its impulse per unit of amplitude, $I_{1}(m)/I_{1}(0)$ with $I_{1}(m)=TN(m)/\left[2 K(m)\right] $, presents a single maximum at $m=m_{\max}\simeq0.717$. For the excitation $f_{2}(t)$, its impulse is written $I_{2}(m)=T\mathrm{arctanh}(\sqrt{m})/\left[2\sqrt{m}K(m)\right]$, and hence it corresponding normalized impulse $I_{2}(m)/I_{2}(0)$ grows monotonically from $1$ to $1.5$. Figure \ref{fig:driving} shows the time-dependence of both excitations over a period together with the dependence of their respective normalized impulses on the shape parameter $m$. The use of Jabobian elliptic functions as periodic excitations is mainly motivated by the fact that, after normalizing their (natural) arguments to keep the period as a fixed independent parameter, their waveforms can be changed by solely varying a {\em single} parameter: the elliptic parameter $m$, and hence the corresponding impulse will only depend on $m$ once the amplitude and period are fixed. \begin{figure}[tb] \begin{center} \begin{tabular}[c]{cc} \includegraphics[width=4cm]{fig1aN.pdf} \includegraphics[width=4cm]{fig1bN.pdf} \end{tabular} \end{center} \caption{External excitations (\ref{eq:driving1}) and (\ref{eq:driving2}) vs time over a period (panels (a) and (b), respectively). In the former case, dashed (blue), thin (purple), and thick (golden) lines correspond to $m=0$, $m=m_{\max}=0.717$, and $m=0.99$, respectively. In the latter case, the same lines correspond to $m=0$, $m=0.99,$ and $m=1-10^{-14}$, respectively. Notice that, when $m\rightarrow1$, the excitation $f_{2}$ becomes a square wave signal while $f_{1}$ vanishes. The insets show the respective normalized impulses of the excitations vs the shape parameter (see the text).}% \label{fig:driving}% \end{figure} The aim of this paper is to study the effectiveness of the excitation's impulse at controlling breathers arising in Eq. (\ref{eq:dyn}) by considering two prototypical on--site potentials. First, a hard $\phi^{4}$ potential, which was previously considered in \cite{Cubero} for the limiting case of a harmonic excitation $\left( m=0\right) $, and where it was shown that there exists a threshold value of $f_{0}$ such that breathers do not exist below it. Remarkably, such a threshold amplitude can be decreased in the presence of noise through a stochastic resonance mechanism. Second, a sine-Gordon potential \cite{Marin} so that Eq. (\ref{eq:dyn}) becomes the so-called Frenkel--Kontorova model (see, e.g., Refs. \cite{Kivshar,sG} for additional details), in which the emergence of discrete moving breathers \cite{moving} is indicated by the existence of a pitchfork bifurcation together with the appearance of an intermediate state. The existence of discrete breathers is characterized by using techniques based on the anti-continuous (AC) limit \cite{AClimit}. Thus, two periodic attractors must be found in such a limit (i.e., for the corresponding isolated nonlinear oscillator) such that the attractor with the largest amplitude is assigned to the central ($n=0$) site of the chain, while the other periodic attractor is assigned to the rest of the coupled oscillators. Such a solution is then continued from $C=0$ to the prescribed value of $C$. Since discrete breathers are periodic orbits in phase space, they can be calculated by means of a shooting method, i.e., they can be considered as fixed points of the map:% \begin{equation} \left[ \{u_{n}(0)\},\{\dot{u}_{n}(0)\}\right] \rightarrow\left[ \{u_{n}(T)\},\{\dot{u}_{n}(T)\}\right] . \end{equation} This analysis is accomplished by using a Powell hybrid algorithm complemented by an 5th-6th order Runge--Kutta--Verner integrator. To study the stability of discrete breathers, a small perturbation $\xi_{n}$ is introduced to a given $u_{n0}$ solution of Eq. (\ref{eq:dyn}) according to $u_{n}=u_{n0}+\xi_{n}$. Thus, one obtains the equation which is verified (to first order) by $\xi_{n}$: \begin{equation} \ddot{\xi}_{n}+\alpha\dot{\xi}_{n}+V^{\prime\prime}(u_{n0})\xi_{n}+C(2\xi _{n}-\xi_{n+1}-\xi_{n-1})=0. \end{equation} To determine the orbital stability of periodic orbits, a Floquet analysis can be performed so that the stability properties are deduced from the spectrum of the Floquet operator (whose matrix representation is the monodromy $\mathcal{M}$), given by \begin{equation} \left[ \{\xi_{n}(T)\},\{\dot{\xi}_{n}(0)\}\right] ^{T}=\mathcal{M}\left[ \{\xi_{n}(0)\},\{\dot{\xi}_{n}(0)\}\right] ^{T}, \end{equation} where $\lambda=\exp(i\theta)$ are the \textit{Floquet multipliers} while the values of $\theta$ are the \textit{Floquet exponents}. All eigenvalues must lie inside the unit circle if the breather is stable. \section{Numerical results} \label{sec:numerics} Our numerical study starts with the case of a hard $\phi^{4}$ potential, that is, $V(u)=u^{2}/2+u^{4}/4$. Notice that breathers in such a potential exhibit staggered tails due to its hardness. This means that the system must be driven following this pattern by taking $h=1$ in (\ref{eq:driving}) (see Figs. \ref{fig:hard}(a) and \ref{fig:hard}(b)). It has been shown for $m=0$ \cite{Cubero} that breathers exist if $f_{0}>f_{\mathrm{th}}$, i.e., the excitation amplitude must surpass a certain threshold. In general, this threshold is a function of the system parameters. Here we study the dependence of this threshold on the shape parameter $m$, while keeping fixed the remaining parameters. Figure \ref{fig:hard}(c) shows this dependence for the parameters $\alpha=0.1$, $C=1$, $\omega_{\mathrm{b}}=5$. For the excitation $f_{1}(t)${\textbf{ }}(\ref{eq:driving1}), it should be emphasized the existence of a minimum threshold at a critical value $m=m_{c}\approx0.664$ for such set of parameters; however, if any of such parameters were varied, $m_{c}$ would remain close to such a value. Although this critical value does not exactly match the value $m=$ $m_{\max}% \approx0.717$ at which the impulse $I_{1}\left( m\right) $ presents a single maximum, it is very close to the value $m=m_{1}=0.6416$ where the first harmonic of the Fourier expansion of the external driving presents a single maximum. Also, the waveforms corresponding to $m_{c}$ and $m_{\max}$ can be hardly distinguishable, as is shown in Fig. \ref{fig:hard}(d), which means that the values of their respective impulses are almost identical (the relative difference is only $\sim0.42\%$). The fact that $m_{c}$ does not change significantly when $\omega_{\mathrm{b}}$ and $C$ are varied implies that this property holds in the AC limit. Indeed, for the isolated oscillator we found that for the largest-amplitude attractor there exists a minimum value of its amplitude, $f_{\mathrm{th}}$, at $m=0.668\approx m_{c}$, while the smallest-amplitude attractor exists for any value of $f_{0}$ (i.e., $f_{\mathrm{th}}=0,\ \forall m$). Thus, the breather seems to inherit this key feature (impulse-induced threshold behaviour) of the largest-amplitude attractor of the isolated oscillator. It is worth mentioning that $\alpha$ must be sufficiently small in order that two periodic attractors can exist in the AC limit. For the excitation $f_{2}(t)$ (\ref{eq:driving2}), we found that the threshold amplitude $f_{\mathrm{th}}$ exhibits a monotonously decreasing behavior as a function of the shape parameter (see \ref{fig:hard}(c)), as expected from the monotonously increasing behavior of its impulse $I_{2}\left( m\right) $. Thus, the analysis of both periodic excitations $f_{1,2}(t)$ confirmed the same effect of the excitation's impulse on the amplitude threshold for the existence of breathers. \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig2a.pdf} & \includegraphics[width=4cm]{fig2b.pdf}\\ \includegraphics[width=4cm]{fig2c.pdf} & \includegraphics[width=4cm]{fig2d.pdf}\\ & \end{tabular} \end{center} \caption{Profiles of $u_{n},p_{n}\equiv\dot{u}_{n}$ (a) and phase space diagrams (b) of a breather in a $\phi^{4}$ chain for $m=0.6$ and $f_{0}=3$. (c) Threshold amplitude for the existence of breathers vs shape parameter $m$. Solid (blue) and dashed (red) lines correspond to excitations (\ref{eq:driving1}) and (\ref{eq:driving2}), respectively. Notice that in the former case, there exists a minimum at $m=m_{c}=0.664$ for $f_{\mathrm{th}% }=2.056$, whereas in the latter case, one sees a monotonously decreasing behavior. (d) External excitation (\ref{eq:driving1}) vs time over a period for two values of the shape parameter: $m=m_{c}$, for which $f_{\mathrm{th}}$ is minimum, and $m=m_{\max}=0.717$, for which the impulse is maximum. Fixed parameters: $\alpha=0.1$, $C=1$, $\omega_{\mathrm{b}}=5$.}% \label{fig:hard}% \end{figure} Next, we consider the Frenkel--Kontorova model, i.e., the case of a sine-Gordon potential: $V^{\prime}(u)=\sin(2\pi u)/(2\pi)$. This case, which was previously analyzed in detail for the limiting case of a harmonic excitation $\left( m=0\right) $ in \cite{Marin}, is much richer than the previous one due to the existence of exchange of stability bifurcations, Hopf / Neimark-Sacker bifurcations, chaos, moving breathers \cite{Meister}, and rotobreathers \cite{roto}. Figure \ref{fig:soft} shows the position, velocity, and phase space diagram of a typical breather. Since the sine-Gordon potential is soft, tails are unstaggered and hence one takes $h=0$ in (\ref{eq:driving}) in order that the periodic excitation fit this pattern. According to Ref. \cite{Marin}, for $\alpha=0.02$ and $f_{0}\gtrsim0.05$ the largest-amplitude attractor in the AC limit corresponds to a rotation, and hence it cannot be used for the analysis of breathers (the analysis of rotobreathers is beyond the scope of the present work). Thus, we are fixing $\alpha=f_{0}=0.02$ in our numerical simulations. Similarly to the case of a hard $\phi^{4}$ potential, we found a threshold for the existence of breathers inheriting the features of the largest-amplitude attractor of the AC limit \footnote{ The existence of such thresholds for the existence of breathers in driven-damped systems seems to be a general phenomenon, as demonstrated in \cite{Feckan} for localized periodic travelling waves in FPU lattices}. Additionally, we found an interesting behaviour arising from stability exchange bifurcations that leads to the onset of moving breathers \cite{moving}. In this kind of bifurcations, a site-centered breather (i.e., a breather with a single site excited at the AC limit) undergoes a supercritical pitchfork bifurcation becoming unstable past a critical value of the coupling, $C=C_{1},$ while a new kind of breather appears --- the so-called \textit{intermediate breather} --- where two adjacent sites are excited with different amplitudes. This intermediate breather disappears after undergoing a subcritical pitchfork bifurcation at $C=C_{1}^{\prime}$. At this coupling value, a site-centered breather (i.e., a breather with two adjacent sites excited with the same amplitude), which is unstable for $C<C_{1}^{\prime}$, changes its stability (see Fig. 22 in \cite{Aubry}, and Figs. 5 and 8 in \cite{Marin}). Figures \ref{fig:soft}(c) and \ref{fig:soft}(d) show the onset of this instability and the dynamics of a moving breather, respectively, for the excitation $f_{1}(t)$. The spatiotemporal patterns of moving breathers is illustrated by plotting their energy density:% \begin{equation} E_{n}=\frac{\dot{u}_{n}^{2}}{2}+V(u_{n})+\frac{C}{4}\left[ (u_{n}% -u_{n+1})^{2}+(u_{n}-u_{n-1})^{2}\right] . \end{equation} We found these results for frequencies over the range $1/2<\omega_{\mathrm{b}% }<2/3$. Note, however, that over the range $2/3<\omega_{\mathrm{b}}<1$ the discussed phenomenology can change due to the properties of Floquet exponents \cite{stab}. Indeed, for frequencies over the range $2/3<\omega_{\mathrm{b}% }<1$, the breather undergoes a Neimark--Sacker bifurcation as the coupling is increased, making it unstable past a critical coupling value $C=C_{2}$. This instability is characterized by the eventual destruction of the breather (i.e., the localization is lost and only a linear mode remains; see Fig. \ref{fig:soft2}). The critical value $C_{2}$ is much smaller than $C_{1}$ (in fact, $C_{2}$ is close to 0, i.e. to the AC limit). Therefore, it has no sense to study the emergence of moving breathers by stability exchange bifurcations. Note that this does not mean that moving breathers cannot exist for $\omega_{\mathrm{b}}>2/3$. The mechanism for the emergence of breathers when $\omega_{\mathrm{b}}>2/3$ is simply different: it is no more than the spontaneous motion described in Ref. \cite{Marin,Meister}. Notice that in Hamiltonian systems, moving breathers exist over this range of frequencies (cf. \cite{Kivshar}). Next, Fig. \ref{fig:soft3} shows the dependence of the critical values $C_{1},$ $C_{2}$ as functions of the shape parameter $m$. For the excitation $f_{1}(t)$, one sees that $C_{1}$ presents a minimum at $m\approx0.64$ when $\omega_{\mathrm{b}}=0.2\pi$, while $C_{2}$ presents a minimum at $m\approx0.67$ when $\omega_{\mathrm{b}}=0.8$. Notice that these values of the shape parameter are significantly close to $0.717\approx m_{\max}$, indicating once more again the effect of the excitation's impulse. The stability range increases as $m$ is increased from these values (see Fig. \ref{fig:soft3}). However, one expects that the pitchfork and Neimark-Sacker bifurcations disappear as $m\rightarrow1$ since in such a limit the excitation and the localization vanish. For the excitation $f_{2}(t)$, one sees that $C_{1}% ,C_{2}$ present a monotonously decreasing behavior, as expected from the monotonously increasing behavior of the impulse $I_{2}(m)$. Thus, the analysis of both periodic excitations $f_{1,2}(t)$ confirmed the same effect of the excitation's impulse on the critical coupling values $C_{1,2}$. \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig3a.pdf} & \includegraphics[width=4cm]{fig3b.pdf}\\ \includegraphics[width=4cm]{fig3c.pdf} & \includegraphics[width=4cm]{fig3d.pdf}\\ & \end{tabular} \end{center} \caption{Profiles of $u_{n},p_{n}\equiv\dot{u}_{n}$ (a) and phase space diagrams (b) of a breather in a chain with a sine-Gordon potential for the excitation (3) and $C=0.5$. (c) Modulus of the Floquet multiplier $\lambda$ vs coupling $C$. The dashed vertical line indicates de location of $C_{1}$. (d) Spatiotemporal pattern of the energy density of a moving breather emerging from the unstable solution for $C=0.52$. Fixed parameters: $\alpha=f_{0}% =0.02$, $m=0.66$, $\omega_{\mathrm{b}}=0.2\pi$.}% \label{fig:soft}% \end{figure} \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig4a.pdf} & \includegraphics[width=4cm]{fig4b.pdf}\\ & \end{tabular} \end{center} \caption{(a) Modulus of the Floquet multiplier $\lambda$ vs coupling $C$ for the excitation (3). The dashed vertical line indicates de location of $C_{2}$. (b) Spatiotemporal pattern of the energy density of an unstable breather for $C=0.1$. This instability finally leads to the destruction of the breather. Fixed parameters: $m=0.66$, $\alpha=f_{0}=0.02$, $\omega_{\mathrm{b}}=0.8$.}% \label{fig:soft2}% \end{figure} \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig5a.pdf} & \includegraphics[width=4cm]{fig5b.pdf}\\ & \end{tabular} \end{center} \caption{(a) Loci $C_{1}$ (see the text) of the exchange of stability bifurcation (i.e., onset of moving breathers) as a function of the shape parameter $m$ for $\alpha=f_{0}=0.02$, $\omega_{\mathrm{b}}=0.2\pi$. (b) Loci $C_{2}$ (see the text) of Neimark--Sacker bifurcation as a function of the shape parameter $m$ for $\alpha=f_{0}=0.02$, $\omega_{\mathrm{b}}=0.8$. Solid (blue) and dashed (red) lines correspond to excitations (\ref{eq:driving1}) and (\ref{eq:driving2}), respectively.}% \label{fig:soft3}% \end{figure} \section{Discussion} The numerical results discussed in Sec. \ref{sec:numerics} may be understood by considering the excitation's impulse through an energy-based analysis, including the properties of the action, of isolated oscillators. For the sake of clarity, we will consider, for example, the excitation $f_{1}(t)${\textbf{ }}(\ref{eq:driving1}) in the subsequent analysis. Indeed, every breather possesses a tail due to its localized character while the oscillators forming this tail effectively behave as linear oscillators presenting a small-amplitude attractor. Consequently, a breather can inherit some properties associated with the effective linear character of the oscillators forming its tail. We found indeed that breathers inherit the dependence on the shape parameter according to the impulse principle. Thus, we analyze the response of a linear (harmonic) oscillator subjected to a periodic anti-symmetric driving:% \begin{equation} \label{eq:linear_osc}\ddot{u}+\alpha\dot{u}+\omega_{0}^{2}u=f_{0}\sum _{k=0}^{\infty}G_{2k+1}\sin\left[ (2k+1)\omega_{\mathrm{b}}t\right] , \end{equation} where $G_{2k+1}$ are the Fourier coefficients of the non-harmonic excitation (\ref{eq:driving1}):% \begin{equation} G_{k}=\frac{\pi^{2}N(m)k}{2\sqrt{m}K^{2}(m)}\mathrm{sech}\left[ \frac{k\pi K(1-m)}{2K(m)}\right] . \end{equation} After some straightforward algebra, one obtains the solution% \begin{equation} u(t)=\sum_{k=0}^{\infty}\left[ A_{2k+1}\cos\left( (2k+1)\omega_{\mathrm{b}% }t\right) +B_{2k+1}\sin\left( (2k+1)\omega_{\mathrm{b}}t\right) \right] , \end{equation} where \begin{align} A_{k} & =f_{0}\frac{\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2}}{k^{2}% \omega^{2}\alpha^{2}+(\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2})^{2}}% G_{k},\nonumber\\ B_{k} & =f_{0}\frac{-k\omega_{\mathrm{b}}\alpha}{k^{2}\omega_{\mathrm{b}% }^{2}\alpha^{2}+(\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2})^{2}}G_{k}. \end{align} The action $J\equiv\frac{1}{2\pi}\oint\dot{u}\mathrm{d}u$ can be recast into the form% \begin{equation} J=\frac{1}{\pi}\int_{0}^{T/2}\left( \dot{u}(t)\right) ^{2}\mathrm{d}% t=\frac{\omega}{2}\sum_{k=0}^{\infty}k^{2}(A_{k}^{2}+B_{k}^{2}), \end{equation} with $T=2\pi/\omega_{\mathrm{b}}$ being the oscillator's period. Thus, the action of the linear oscillator can be finally expressed as% \begin{align} \label{eq:action_linear}J & =\frac{\omega_{\mathrm{b}}f_{0}}{2}\sum _{k=0}^{\infty}\mu_{2k+1}G_{2k+1}^{2},\nonumber\\ \mu_{k} & =k^{2}\frac{(\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2}% )^{2}+k^{2}\omega_{\mathrm{b}}^{2}\alpha^{2}}{[k^{2}\omega_{\mathrm{b}}% ^{2}\alpha^{2}+(\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2})^{2}]^{2}}. \end{align} Notice that $\mu_{k}$ does not depend on the particular waveform of the external periodic excitation $F(t)$, but depends on $\omega_\mathrm{b}$ and $\alpha$. Then, the dependence of the action on the shape parameter $m$ appears only in the $G_{k}^{2}$ terms. Now, after taking into account the fast decay of the Fourier coefficients with $k$, one numerically finds that the action presents a single maximum at $m=m_{\ell}$ which is very close to $m_{1}$, where $m_{1}$ is the shape parameter value at which $G_{1}$ presents a single maximum (recall from Sec. \ref{sec:numerics} that $m_{1}=0.6416$). Notice that $m_{\ell}$ depends on $\mu_{k}$ and, consequently, on $\omega_\mathrm{b}$ and $\alpha$. For instance, for the parameters sets taken in Figs. \ref{fig:hard} and \ref{fig:soft2}, i.e. ($\alpha=0.1$, $\omega_\mathrm{b}=5$) and ($\alpha=0.02$, $\omega_\mathrm{b}=0.8$), the value of $m_{\ell}$ is 0.646 and 0.644, respectively. Remarkably, the above mentioned properties also holds for the corresponding average energies $<E>$, which for the linear oscillator reads:% \begin{equation} <E>=\int_{0}^{T}\left( \frac{1}{2}\dot{u}^{2}(t)+\frac{\omega_{0}^{2}}% {2}u^{2}(t)\right) \end{equation} For the excitation (\ref{eq:driving1}), it can be recast into the simple form% \begin{align} \label{eq:energy_linear}<E> & =\frac{\pi f_{0}}{2}\sum_{k=0}^{\infty}% \mu^{\prime}_{2k+1}G_{2k+1}^{2},\nonumber\\ \mu^{\prime}_{k} & =\frac{(k^{2}\omega_{\mathrm{b}}^{2}+\omega_{0}% ^{2})[(\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2})^{2}+k^{2}\omega _{\mathrm{b}}^{2}\alpha^{2}]}{\omega_{\mathrm{b}}[k^{2}\omega_{\mathrm{b}}% ^{2}\alpha^{2}+(\omega_{0}^{2}-k^{2}\omega_{\mathrm{b}}^{2})^{2}]^{2}}. \end{align} Figure \ref{fig:action1} shows the dependence of the action and average energy of solutions of the linear oscillator [Eq. (\ref{eq:linear_osc})] on the shape parameter for the complete Fourier series and the main harmonic approximation ($G_{2k+1}=0,\ \forall k\geq1$) and two sets of the remaining parameters. \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig6a.pdf} & \includegraphics[width=4cm]{fig6b.pdf}\\ \includegraphics[width=4cm]{fig6c.pdf} & \includegraphics[width=4cm]{fig6d.pdf}\\ & \end{tabular} \end{center} \caption{Action $J$ [Eq. (\ref{eq:action_linear})] (top panels) and average energy $<E>$ [Eq. (\ref{eq:energy_linear})] (bottom panels) of solutions of the linear oscillator [Eq. (\ref{eq:linear_osc})] vs shape parameter $m$ for the complete Fourier series (solid lines) of excitation (3) and the main harmonic approximation ($G_{2k+1}=0\ \forall k\geq1$; dashed lines) and two sets of parameters: (left panels) $\omega_{\mathrm{b}}=0.8$, $f_{0}=0.02$, $\alpha=0.02$. (right panels) $\omega_{\mathrm{b}}=5$, $f_{0}=3$, $\alpha =0.1$.}% \label{fig:action1}% \end{figure} As already anticipated in Sec. \ref{sec:intro}, threshold phenomena associated with breathers' emergence and stability exhibit a high sensitivity to the excitation's impulse. To show this, we start with a general argument showing the relationship between energy changes and the quantities action and impulse for periodic solutions of isolated (nonlinear) oscillators. After integrating the corresponding energy equation over half a period (see, e.g., \cite{6,7}), one obtains% \begin{equation} \Delta E\equiv E(T/2)-E(0)=-\alpha\int_{0}^{T/2}\left( \dot{u}^{2}(t)\right) \mathrm{d}t+\int_{0}^{T/2}\dot{u}(t)F(t)\mathrm{d}t. \end{equation} Now, after applying the first mean value theorem for integrals \cite{GR} to the last integral on the right-hand side of Eq. (\ref{eq:energy_linear}) and recalling the definitions of action and impulse, one obtains% \begin{equation} \Delta E=-\alpha\pi J+T\dot{u}(t^{\ast})I, \end{equation} where $t^{\ast}\in\lbrack0,T/2]$ while $J$ and $I$ are the action and the impulse, respectively. Note that $t^{\ast}$ becomes independent of the excitation's waveform as $T\rightarrow0$ \cite{6,7}. It should be stressed that this limiting regime is unreachable for the present case of discrete breathers in nonlinear chains, specially in the soft potential case, due to breather frequencies are always below a maximum, and hence they cannot be increased arbitrarily. Therefore, we see that the dependence of $\Delta E$ on $m$ for any (linear or nonlinear) isolated oscillator relies on the dependence on $m$ of $J$ and $\dot{u}(t^{\ast})I$. For the case of a linear oscillator and the excitation $f_{1}(t)${\textbf{ }}(\ref{eq:driving1}), one readily obtains that $\dot{u}(t^{\ast})I$ can be expressed as $\sum_{k}\mu _{2k+1}^{(1)}G_{2k+1}^{2}$, where $\mu_{k}^{(1)}$ is independent on $m$, and hence the energy will present a maximum at $m=m_{\ell}^{\prime}\approx m_{\ell}$. We find that the dependence of the action on the shape parameter for nonlinear oscillators is quite similar to that of the discussed linear case. To connect this analysis of isolated oscillators with discrete breathers of nonlinear chains (\ref{eq:dyn}), one has to calculate the action of a breather, $J=\sum_{n}\oint\dot{u}_{n}\mathrm{d}u_{n}$, on the one hand, and to distinguish between the periodic attractors with large and the small amplitudes, on the other hand, since the orbits associated with the latter can substantially differ from those of a strictly linear oscillator despite of its relatively small oscillation amplitude. In any case, numerical simulations confirmed that the value $m=m_{\max}\approx0.717$ at which the impulse function $I_{1}(m)$ presents a single maximum is very close to $m_{\ell}$ in the sense that the waveforms corresponding to $m_{\max}$ and $m_{\ell}$ (and $m_{1}$) can be hardly distinguishable. Figure \ref{fig:action2} shows an illustrative example for the cases of a hard $\phi^{4}$ potential and a sine-Gordon potential, while Fig. \ref{fig:action_breather} shows, for the case of a hard $\phi^{4}$ potential, that the breather action presents a single maximum at $m=m_{b,\max}$ which is also very close to $m_{\ell}% \approx0.646$. \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig7a.pdf} & \includegraphics[width=4cm]{fig7b.pdf}\\ \includegraphics[width=4cm]{fig7c.pdf} & \includegraphics[width=4cm]{fig7d.pdf}\\ & \end{tabular} \end{center} \caption{Action $J$ of (top panels) large-amplitude attractors and (bottom panels) small-amplitude attractors corresponding to an isolated nonlinear oscillator with (left panels) a sine-Gordon potential for the excitation (3), $\omega_{\mathrm{b}}=0.8$, $f_{0}=0.02$, $\alpha=0.02$, and (right panels) a $\phi^{4}$ potential for $\omega_{\mathrm{b}}=5$, $f_{0}=3$, $\alpha=0.1$. The vertical dashed lines indicate the $m$ value where the action of the linear oscillator is maximum, i.e. $m=m_{\ell}$ (cf. Fig. \ref{fig:action1}).}% \label{fig:action2}% \end{figure} \begin{figure}[tb] \begin{center}% \begin{tabular} [c]{cc}% \includegraphics[width=4cm]{fig8a.jpg} & \includegraphics[width=4cm]{fig8b.pdf}\\ & \end{tabular} \end{center} \caption{(a) Action $J$ of a breather emerging in a $\phi^{4}$ chain vs $f_{0}$ and $m$ in the range $f_{0}\in(2.06,3)$ for the excitation (3). (b) Shape parameter value at which the action is maximum $m_{b,\max}$ vs amplitude $f_{0}$; horizontal line corresponds to $m=m_{\ell}=0.646$. Fixed parameters: $\omega_{\mathrm{b}}=5$, $\alpha=0.1,$ $C=1$.}% \label{fig:action_breather}% \end{figure} \section{Conclusions} We have shown through the example of a discrete nonlinear Klein-Gordon equation that varying the impulse transmitted by periodic external excitations is a universal procedure to reliably control the generation of stationary and moving discrete breathers in driven dissipative chains capable of presenting these intrinsic localized modes. We have analytically demonstrated that the enhancer effect of the excitation's impulse, in the sense of facilitating the generation of stationary and moving breathers, is due to a correlative increase of the breather's action, while numerical experiments corresponding to the cases of a hard $\phi^{4}$ potential and a sine-Gordon potential confirmed the effectiveness of the impulse\textbf{ }as the relevant quantity controlling the effect of the external excitation. The consideration of this relevant quantity opens up new avenues for studying external-excitation-induced phenomena involving intrinsic localized modes in discrete nonlinear systems, including, for instance, breather-to-soliton transitions and emergence of chaotic breathers. Our present work is aimed to explore these and related problems. \begin{acknowledgments} R. C. gratefully acknowledges financial support from the Junta de Extremadura (JEx, Spain) through Project No. GR15146. \end{acknowledgments}
a852461f06543c185aaa3a71c91facb103961664
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sec:intro} The CDM paradigm, within which the majority of matter in the Universe is weakly interacting, has enjoyed great success in explaining astrophysical and cosmological data. The inclusion of CDM in the now standard \ensuremath{\Lambda}CDM{} cosmological model makes clear predictions for, among other things, the mass function of gravitationally collapsed structures (e.g.\ \citealt{Jenkins0005260, Evrard0110246, Tinker0803.2706}), the ratio of baryonic mass to CDM in galaxy clusters (e.g.\ \citealt{Eke9708070, Kay0407058, Borgani0906.4370}), and the distribution of mass within the structures that form hierarchically in the Universe (e.g.\ \citealt{Bullock9908159, Navarro0311231, Navarro9611107, Gao0711.0746}). Observations testing the predicted mass function and cluster baryon fraction have largely validated CDM (e.g.\ \citealt{White1993Natur.366..429W, Bahcall9803277, Reiprich0111285, Ettori0211335, Ettori0904.2740, Allen0405340, Allen0706.0033, Allen1103.4829, Vikhlinin0812.2720, Mantz0909.3098, Mantz1407.4516, Rozo0902.3702, Sehgal1010.1025, Benson1112.5435, Hasselfield1301.0816, Planck1303.5080, Planck1502.01597}). Observational tests of the distribution of mass within bound structures is in some ways more challenging, requiring spatially resolved measurements of the gravitational potential of objects whose mass is predominately dark. Nevertheless, an extensive body of work now exists, using primarily X-ray and gravitational lensing observations of galaxy clusters \citep{Vikhlinin0507092, Voigt0602373, Zhang0603275, Schmidt0610038, Mandelbaum0805.2552, Host0907.1097, Newman1209.1391, Okabe1302.2728, Du1510.08193, Merten1404.1376, Shan1502.00313, van-Uitert1506.03817}. In this work, we revisit the subject using {\it Chandra}{} X-ray observations of a sample of massive, highly dynamically relaxed clusters. Although not representative of the cluster population at large, these systems are an ideal laboratory for deriving three-dimensional mass profiles based on observations of the intracluster medium (ICM) because departures from hydrostatic equilibrium and systematics due to projection are minimized. Both simulations and direct calibration using weak gravitational lensing indicate that the overall bias in {\it Chandra}{} X-ray mass estimates for this sample is small ($\ {\raise-.75ex\hbox{$\buildrel<\over\sim$}}\ 10$ per cent; \citealt{Nagai0609247, Applegate1509.02162}). The specific features of the cluster mass distribution that we consider here are (1) the concentration parameter of cluster mass profiles, and its dependence on mass and redshift; and (2) the shape of mass profiles, in particular departures from the baseline model defined by \citet*{Navarro9611107} (hereafter NFW). The selection of the sample of massive, relaxed clusters employed here is detailed in the first paper in this series (\citealt{Mantz1502.06020}; hereafter Paper~\citetalias{Mantz1502.06020}{}). Papers~\citetalias{Mantz1402.6212}{}, \citetalias{Mantz1509.01322}{} and \citetalias{Applegate1509.02162}{} \citep{Mantz1402.6212, Mantz1509.01322, Applegate1509.02162} employ {\it Chandra}{} data for this sample to respectively constrain cosmological parameters, through gas-mass fraction measurements; scaling relations and thermodynamics of the ICM; and the average bias of the X-ray mass determinations. Section~\ref{sec:data} reviews our procedure for constraining cluster mass profiles from X-ray data, for which complete details are available in Papers~\citetalias{Mantz1402.6212}{} and \citetalias{Mantz1509.01322}{}, and introduces the specific mass models employed in this work. In Section~\ref{sec:results}, we review the specific predictions of CDM for cluster mass profile concentrations and shapes, present our results, and compare them to others in the literature. Section~\ref{sec:conclusions} summarizes our conclusions. We assume a concordance flat \ensuremath{\Lambda}CDM{} cosmology with $h=0.7$ and $\mysub{\Omega}{m}=0.3$ throughout. Unless otherwise noted, quoted parameter uncertainties correspond to the maximum-likelihood (i.e.\ shortest) interval enclosing 68.3 per cent posterior probability, and best-fitting values are the posterior modes. \section{Data and Analysis} \label{sec:data} This work employs the sample of 40 massive, dynamically relaxed galaxy clusters identified in Papers~\citetalias{Mantz1502.06020}{} and \citetalias{Mantz1402.6212}{} (Table~\ref{tab:thetable}). For our purposes, dynamical relaxation was defined quantitatively in terms of X-ray image features, specifically the sharpness of the surface brightness peak and the alignment and symmetry of a series of standardized cluster isophotes; in addition all clusters in the sample must have an emission-weighted temperature of $>5$\,keV outside of their cores. The procedure for cleaning and fitting the {\it Chandra}{} data for these clusters is described in detail in Paper~\citetalias{Mantz1402.6212}{}; in Paper~\citetalias{Mantz1509.01322}{} the data reduction was updated to use a newer version of the {\it Chandra}{} calibration files (specifically {\sc caldb}\footnote{\url{http://cxc.harvard.edu/caldb/}} 4.6.2), and we use that version of the data reduction here. A complete list of the specific OBSIDs included in the analysis can also be found in that work. Note that, while the data for each of these 40 clusters can constrain the 2-parameter NFW mass model, the 3-parameter models discussed below can be constrained for only a subset of entire sample (see Sections~\ref{sec:gnfw} and \ref{sec:einasto}). \begin{table*} \centering \caption[]{ The galaxy cluster data set. Column [1] name, [2] redshift, [3--4] the radial range from which data is used to constrain mass profiles, [5--6] mass and concentration parameter constraints assuming the NFW profile, [7] central cooling luminosity, [8] inner slope constraints using the GNFW profile, and [9] Einasto profile shape parameter constraints. Quoted values are the medians of the marginalized posterior distributions for each parameter, and confidence intervals correspond to the 15.85th and 84.15th percentiles of the posterior (enclosing 68.3 per cent confidence). } \label{tab:thetable} \begin{tabular}{lcrrr@{ $\pm$ }lr@{ $\pm$ }lr@{ $\pm$ }lcc} \hline Name & Redshift & $R_\mathrm{min}$ & $R_\mathrm{max}$ & \multicolumn{2}{c}{$M_{200}$} & \multicolumn{2}{c}{$c$} & \multicolumn{2}{c}{$\mysub{L}{cool}$} & $\beta$ & $\alpha$ \\ & & (kpc) & (kpc) & \multicolumn{2}{c}{($10^{14}\ensuremath{\, M_{\odot}}$)} & \multicolumn{2}{c}{} & \multicolumn{2}{c}{$10^{45}\unit{erg}\unit{s}^{-1}$} & & \\ \hline Abell~2029 & 0.078 & 58 & 928 & 11.3 & 0.4 & 6.2 & 0.2 & 0.836 & 0.004 & $1.03_{-0.32}^{+0.25}$ & $0.42_{-0.17}^{+0.20}$ \\ Abell~478 & 0.088 & 71 & 1142 & 11.4 & 1.0 & 5.1 & 0.4 & 1.186 & 0.010 & $1.24_{-0.07}^{+0.06}$ & $0.31_{-0.08}^{+0.08}$ \\ PKS~0745$-$191 & 0.103 & 60 & 1190 & 14.6 & 0.8 & 5.0 & 0.2 & 1.944 & 0.006 & $0.90_{-0.02}^{+0.02}$ & $0.35_{-0.02}^{+0.02}$ \\ RX~J1524.2$-$3154 & 0.103 & 60 & 833 & 4.1 & 0.4 & 8.1 & 0.9 & 0.322 & 0.002 & $0.55_{-0.34}^{+0.38}$ & $0.56_{-0.13}^{+0.12}$ \\ Abell~2204 & 0.152 & 63 & 1001 & 13.0 & 1.2 & 7.7 & 0.7 & 1.976 & 0.014 & $0.51_{-0.23}^{+0.17}$ & -- \\ RX~J0439.0+0520 & 0.208 & 54 & 857 & 5.1 & 0.8 & 7.9 & 1.3 & 0.463 & 0.009 & $0.53_{-0.29}^{+0.36}$ & $0.58_{-0.14}^{+0.11}$ \\ Zwicky~2701 & 0.214 & 62 & 493 & 5.2 & 0.5 & 5.9 & 0.6 & 0.398 & 0.005 & -- & -- \\ RX~J1504.1$-$0248 & 0.215 & 48 & 1540 & 17.0 & 1.7 & 5.8 & 0.4 & 6.084 & 0.049 & -- & -- \\ Zwicky~2089 & 0.235 & 51 & 705 & 4.6 & 0.7 & 5.9 & 0.7 & 0.842 & 0.010 & -- & -- \\ RX~J2129.6+0005 & 0.235 & 103 & 1646 & 8.7 & 1.2 & 6.0 & 1.0 & 0.950 & 0.023 & $0.78_{-0.45}^{+0.34}$ & $0.41_{-0.19}^{+0.21}$ \\ RX~J1459.4$-$1811 & 0.236 & 147 & 1061 & 10.1 & 1.6 & 4.7 & 1.1 & 0.990 & 0.013 & -- & -- \\ Abell~1835 & 0.252 & 101 & 992 & 17.3 & 1.3 & 4.5 & 0.3 & 3.331 & 0.021 & $0.85_{-0.04}^{+0.04}$ & $0.36_{-0.02}^{+0.03}$ \\ Abell~3444 & 0.253 & 93 & 746 & 10.3 & 1.7 & 4.4 & 0.7 & 1.819 & 0.022 & -- & -- \\ MS~2137.3$-$2353 & 0.313 & 50 & 722 & 5.0 & 0.5 & 8.0 & 0.8 & 2.355 & 0.029 & -- & $0.30_{-0.12}^{+0.13}$ \\ MACS~J0242.5$-$2132 & 0.314 & 54 & 1447 & 7.1 & 2.0 & 7.8 & 2.3 & 2.812 & 0.082 & $1.43_{-0.32}^{+0.22}$ & -- \\ MACS~J1427.6$-$2521 & 0.318 & 46 & 876 & 3.8 & 0.6 & 10.4 & 2.4 & 0.468 & 0.016 & $0.83_{-0.53}^{+0.57}$ & $0.56_{-0.17}^{+0.13}$ \\ MACS~J2229.7$-$2755 & 0.324 & 55 & 1478 & 5.5 & 1.0 & 6.5 & 1.2 & 1.843 & 0.034 & $1.53_{-0.25}^{+0.17}$ & $0.06_{-0.03}^{+0.05}$ \\ MACS~J0947.2+7623 & 0.345 & 53 & 1233 & 12.9 & 1.7 & 5.8 & 0.6 & 4.734 & 0.060 & $1.32_{-0.17}^{+0.09}$ & -- \\ MACS~J1931.8$-$2634 & 0.352 & 68 & 1249 & 11.6 & 1.6 & 5.3 & 0.6 & 3.856 & 0.033 & -- & -- \\ MACS~J1115.8+0129 & 0.355 & 59 & 942 & 9.5 & 1.4 & 6.5 & 1.1 & 2.323 & 0.045 & $1.11_{-0.32}^{+0.27}$ & $0.28_{-0.09}^{+0.10}$ \\ MACS~J1532.8+3021 & 0.363 & 50 & 1592 & 11.1 & 1.1 & 4.9 & 0.3 & 4.669 & 0.035 & $1.20_{-0.32}^{+0.11}$ & $0.15_{-0.03}^{+0.08}$ \\ MACS~J0150.3$-$1005 & 0.363 & 60 & 637 & 4.0 & 0.8 & 7.8 & 1.9 & 0.892 & 0.031 & -- & -- \\ MACS~J0011.7$-$1523 & 0.378 & 51 & 817 & 7.8 & 1.5 & 7.0 & 1.6 & 0.747 & 0.025 & -- & $0.19_{-0.11}^{+0.25}$ \\ MACS~J1720.2+3536 & 0.391 & 52 & 1001 & 8.1 & 1.3 & 6.9 & 1.4 & 1.238 & 0.043 & -- & -- \\ MACS~J0429.6$-$0253 & 0.399 & 53 & 845 & 11.7 & 4.7 & 4.4 & 1.4 & 2.028 & 0.055 & -- & -- \\ MACS~J0159.8$-$0849 & 0.404 & 106 & 1362 & 14.8 & 2.5 & 5.6 & 1.3 & 2.192 & 0.057 & $0.77_{-0.46}^{+0.44}$ & $0.47_{-0.20}^{+0.20}$ \\ MACS~J2046.0$-$3430 & 0.423 & 44 & 1050 & 4.1 & 0.8 & 8.1 & 1.8 & 1.992 & 0.046 & $1.21_{-0.55}^{+0.28}$ & $0.16_{-0.11}^{+0.21}$ \\ IRAS~09104+4109 & 0.442 & 50 & 538 & 8.7 & 1.6 & 6.1 & 0.9 & 2.866 & 0.050 & -- & -- \\ MACS~J1359.1$-$1929 & 0.447 & 45 & 903 & 6.0 & 2.0 & 6.6 & 2.1 & 0.909 & 0.031 & -- & $0.08_{-0.03}^{+0.06}$ \\ RX~J1347.5$-$1145 & 0.451 & 51 & 1453 & 28.7 & 4.5 & 8.0 & 1.1 & 13.919 & 0.300 & $1.11_{-0.29}^{+0.23}$ & $0.22_{-0.08}^{+0.07}$ \\ 3C~295 & 0.460 & 52 & 551 & 6.0 & 1.8 & 5.9 & 1.5 & 0.974 & 0.026 & $0.91_{-0.55}^{+0.33}$ & -- \\ MACS~J1621.3+3810 & 0.461 & 57 & 919 & 7.3 & 1.0 & 7.2 & 1.3 & 1.183 & 0.027 & -- & -- \\ MACS~J1427.2+4407 & 0.487 & 47 & 663 & 6.7 & 1.1 & 6.9 & 1.3 & 1.993 & 0.063 & -- & -- \\ MACS~J1423.8+2404 & 0.539 & 44 & 799 & 7.2 & 0.9 & 7.0 & 0.8 & 3.551 & 0.053 & $1.33_{-0.20}^{+0.14}$ & $0.14_{-0.05}^{+0.08}$ \\ SPT-CL~J2331$-$5051 & 0.576 & 26 & 1652 & 5.6 & 1.2 & 6.2 & 0.9 & 1.329 & 0.088 & $0.48_{-0.26}^{+0.27}$ & $0.57_{-0.14}^{+0.11}$ \\ SPT-CL~J2344$-$4242 & 0.596 & 46 & 839 & 10.3 & 1.9 & 11.0 & 2.2 & 23.713 & 0.817 & -- & -- \\ SPT-CL~J0000$-$5748 & 0.702 & 28 & 1802 & 8.1 & 3.3 & 4.3 & 1.4 & 2.368 & 0.124 & -- & -- \\ SPT-CL~J2043$-$5035 & 0.723 & 43 & 912 & 6.1 & 1.1 & 4.6 & 0.9 & 3.115 & 0.089 & -- & $0.16_{-0.05}^{+0.08}$ \\ CL~J1415+3612 & 1.028 & 32 & 634 & 4.6 & 0.8 & 4.8 & 0.9 & 0.941 & 0.040 & -- & -- \\ 3C~186 & 1.063 & 28 & 511 & 5.5 & 1.3 & 4.8 & 1.1 & 2.934 & 0.119 & -- & $0.52_{-0.19}^{+0.14}$ \\ \hline \end{tabular} \end{table*} As in our previous papers, the spectral analysis of the X-ray data here involves a deprojection, assuming spherical symmetry and hydrostatic equilibrium, in order to infer the 3-dimensional properties of the hot gas and gravitating matter in the clusters. We consider three parametric models for the total mass profiles of clusters: the standard NFW profile, whose results were previously presented in Papers~\citetalias{Mantz1402.6212}{} and \citetalias{Mantz1509.01322}{}; a generalized NFW (GNFW) profile, in which the power-law slope of the density profile at small radii is a free parameter; and the Einasto profile \citep{Einasto1969Ap......5...67E}, which provides a more accurate description of simulated halos than the other two options, particularly at small radii \citep{Navarro0810.1522}. Up to normalizing factors, the density profiles associated with these models are \begin{eqnarray} \label{eq:profmodels} \mysub{\rho}{NFW}(r) &\propto& \left(\frac{r}{\mysub{r}{s}}\right)^{-1} \left(1 + \frac{r}{\mysub{r}{s}}\right)^{-2}, \\ \mysub{\rho}{GNFW}(r) &\propto& \left(\frac{r}{\mysub{r}{s}}\right)^{-\beta} \left(1 + \frac{r}{\mysub{r}{s}}\right)^{-(3-\beta)}, \nonumber\\ \mysub{\rho}{Einasto}(r) &\propto& \exp \left\{ -\frac{2}{\alpha}\left[\left(\frac{r}{r_{-2}}\right)^\alpha - 1\right] \right\}, \nonumber \end{eqnarray} where \mysub{r}{s}{} and $r_{-2}$ are scale radii. Figure~\ref{fig:demo} shows how the shapes of the GNFW and Einasto models differ from the NFW profile. Note that, while the NFW model is a special case of the GNFW profile with $\beta=1$, no value of $\alpha$ makes the Einasto profile completely equivalent to NFW, although values of $\alpha \approx 0.2$--$0.3$ can closely approximate it over the radial range probed by our data (roughly $0.1 \ {\raise-.75ex\hbox{$\buildrel<\over\sim$}}\ r/\mysub{r}{s} \ {\raise-.75ex\hbox{$\buildrel<\over\sim$}}\ 1.5$, in terms of the NFW scale radius). \begin{figure} \centering \includegraphics[scale=0.9]{fig_demo} \caption[]{ Comparison of the shapes of the GNFW and Einasto density profile models relative to the NFW model. Each profile is normalized at an ordinate of one. The data employed in this work are typically sensitive to radii of 0.1--1.5 in units of the NFW scale radius. } \label{fig:demo} \end{figure} The spectral analysis itself uses {\sc xspec}, within which we have modified the {\sc clmass} model \citep{Nulsen1008.2393} to include the GNFW and Einasto mass profiles in addition to NFW. This model treats the cluster ICM as a series of concentric, isothermal spherical shells of gas, in hydrostatic equilibrium with the gravitational potential given by one of the above parametric mass profiles. We simultaneously model absorption from Galactic hydrogen and, where necessary, the presence of foreground emission from the Galactic halo and/or the local bubble. Complete details of these aspects of the analysis can be found in Paper~\citetalias{Mantz1402.6212}{}. For every cluster, we identify and exclude from the analysis a central region, typically 50\,kpc in radius, where visible features in the ICM indicate that the assumptions of spherical symmetry and hydrostatic equilibrium are not met. This limits the radial range over which we can draw conclusions about the shape of the mass profiles, in particular excluding the central galaxies of each cluster, where the contribution of cold gas and stars to the total mass may be significant (e.g.\ \citealt{Newman1209.1391}). The outer radius used in our analysis is set by the need to measure temperature profiles without background modelling dominating the uncertainties (see Paper~\citetalias{Mantz1402.6212}{}). The specific radial range for each cluster over which the mass profiles are constrained is listed in Table~\ref{tab:thetable}. For each cluster and each mass model considered, our spectral analysis produces a Markov chain of the parameters of interest, for which the density of samples is proportional to the joint posterior density of the parameters. Rather than summarize these results by, e.g., a mean and standard deviation, we continue to use the full set of samples when constraining properties of the ensemble of clusters (e.g. the concentration-mass relation) in Section~\ref{sec:results}. In detail, the samples for each cluster are importance weighted and then directly numerically integrated over to provide a likelihood for the hierarchical model being fitted. This is especially important for constraining the distributions of the GNFW and Einasto shape parameters, as the posterior distributions of these parameters for individual clusters are frequently poorly constrained and highly asymmetric (Table~\ref{tab:thetable}). We additionally calculate central cooling luminosities ($\mysub{L}{cool}$) for each cluster in the sample. These we define as the bolometric luminosity emitted within the radius where the cooling time equals 5\,Gyr. The cooling time is calculated as \begin{equation} t_\mathrm{c} = \frac{3\, n_\mathrm{tot} \,kT}{2\, n_\mathrm{e} n_\mathrm{H} \,\Lambda(T,Z)}, \end{equation} where $T$ is the ICM temperature; $n_\mathrm{e}$, $n_\mathrm{H}$ and $n_\mathrm{tot}$ are the number densities of electrons, protons and all particles; $Z$ is the metallicity; and $\Lambda(T,Z)$ is the cooling function. These calculations are based on spectral analyses of each cluster that do not assume equilibrium or a particular underlying mass profile, using instead the {\sc projct} model in {\sc xspec}, as presented in Paper~\citetalias{Mantz1509.01322}{}. Like total luminosity, $\mysub{L}{cool}$ displays an overall trend with mass; fitting a power-law plus log-normal intrinsic scatter model, we find \begin{equation} \label{eq:Lcool-scaling} \frac{\mysub{L}{cool}}{10^{45}\unit{erg}\unit{s}^{-1}} = \left(2.9\pm0.3\right) \left(\frac{M_{200}}{10^{15}\ensuremath{\, M_{\odot}}}\right)^{1.3\pm0.3}, \end{equation} with a log-normal scatter of $0.70\pm0.09$. \section{Results and Discussion} \label{sec:results} This section presents our results and discusses them in the context of CDM predictions and other measurements in the literature. In Section~\ref{sec:cm}, we assume the NFW mass model and investigate the mass and redshift dependence of cluster concentrations. Sections~\ref{sec:gnfw} and \ref{sec:einasto} respectively consider the GNFW and Einasto generalizations to the NFW profile. \subsection{NFW Concentration--Mass Relation} \label{sec:cm} The concentration of mass within an NFW halo can be described by the parameter \begin{equation} c = \frac{r_{200}}{\mysub{r}{s}}, \end{equation} where $r_{200}$ is the radius within which the mean enclosed density is 200 times the critical density of the Universe at a cluster's redshift. Numerous dark-matter-only simulations have shown that the concentration of halos is expected to vary with both redshift and mass on average, and to have an intrinsic scatter of $\sim20$ per cent driven primarily by differences in accretion history \citep{Navarro9611107, Bullock9908159, Zhao0309375, Zhao0811.0828, Tasitsiomi0311062, Neto0706.2919, Gao0711.0746, Duffy0804.2486, Klypin1002.3660, Klypin1411.4001, Ludlow1206.1049, Ludlow1312.0945, Prada1104.5130, Dutton1402.7073}. Early work suggested a scaling relation similar to $c \propto (1+z)^{-1} \, M^{-0.13}$ \citep{Bullock9908159}. More recent work revealed that the dependence on both mass and redshift is reduced for the most massive (i.e., cluster-scale) halos (e.g.\ \citealt{Gao0711.0746, Klypin1002.3660, Prada1104.5130}), largely because these halos are preferentially recently formed (e.g.\ \citealt{Ludlow1312.0945}). \citet{Ludlow1206.1049} have shown that selecting the most dynamically relaxed halos from simulations recovers a power-law with mass out to the largest masses. Our current understanding, at least on the basis of N-body simulations, is that concentrations for individual clusters grow steadily as they accrete through minor mergers, but that sufficiently disruptive mergers can ``reset'' the concentration parameter to lower values ($c\sim3$), with the overall trends largely reflecting differences in the typical accretion histories of halos as a function of mass and redshift \citep{Ludlow1206.1049, Klypin1411.4001}. By construction, our sample contains only the most dynamically relaxed, massive clusters at any redshift. We therefore expect their accretion histories at the time of observation to be similar, and in particular that the time since the last major merger is significant in dynamical terms. As a consequence, we do not necessarily expect the redshift dependence found in simulations for all halos to apply to our sample, since it must partially reflect evolution in the rate of major mergers that will not be present in our sample. It therefore makes sense to consider models where both the dependence on mass and redshift are free parameters, \begin{equation} \label{eq:cm} c = c_0 \, \left(\frac{1+z}{1.35}\right)^{\kappa_\zeta} \left( \frac{M_{200}}{10^{15}\ensuremath{\, M_{\odot}}} \right)^{\kappa_m}. \end{equation} Our model also includes as a free parameter a log-normal intrinsic scatter in concentration, $\sigma_{\ln c}$, which is assumed to have the same value at all masses and redshifts. \begin{figure*} \centering \hspace{5mm} \includegraphics[scale=0.86]{fig_concentration} \hspace{2mm} \includegraphics[scale=0.86]{fig_concentration_residuals} \includegraphics[scale=0.86]{fig_concentration_cbar} \caption[]{ Left: Mass--concentration relation from the NFW analysis of the 40-cluster sample. The fitting procedure accounts for the covariance between mass and concentration, shown explicitly for RX\,J1347.5$-$1145 (the rightmost point), although for clarity uncertainties are shown as crosses otherwise. The solid line and dashed lines respectively show the best-fitting power law in mass and the corresponding 68 per cent predictive interval, including intrinsic scatter. Right: Residuals in concentration are plotted against residuals in central cooling luminosity. Residuals are normalized by the combined measurement and intrinsic scatter, and are with respect to the best-fitting scaling law with $M_{200}$ in both cases. The clusters with the highest values of $\mysub{L}{cool}$, RX\,J1347.5$-$1145 and SPT-CL\,J2344$-$4242, are shown as open circles in both panels. } \label{fig:cM} \end{figure*} The concentrations and masses of our clusters, derived from fitting the NFW profile, are shown in the left panel of Figure~\ref{fig:cM}. Note that the measurement correlation between $M_{200}$ and $c$ is typically strong, as illustrated by a joint constraint ellipse for the case of RX\,J1347.5$-$1145, and that this correlation is accounted for by our fitting method (Section~\ref{sec:data}). Fitting the model of Equation~\ref{eq:cm}, we find $\kappa_m = -0.16 \pm 0.07$ and $\kappa_\zeta = -0.17 \pm 0.26$. A dependence on mass, consistent with the expectation from simulations, is detected at $\sim2\sigma$ significance, while the redshift dependence is consistent with zero. We find an intrinsic scatter of $\sigma_{\ln c}=0.16\pm 0.03$, somewhat smaller that the $\sim0.2$ scatter found for all halos in simulations. The results are consistent with the notion that our sample consists of highly relaxed clusters, for which we expect a power-law with mass to hold even at the highest masses, with a smaller intrinsic scatter than the full population. This is reinforced by the fact that our concentrations are on average greater than those in simulations at the same masses. Selecting relatively relaxed halos from simulations based on the absence of substructure and the ratio of kinetic to potential energy, \citet{Ludlow1206.1049} find median concentrations of 4--5 in the mass range considered here, consistent with our observed clusters being typically more relaxed than the halos they selected.\footnote{\citet{Ludlow1206.1049} classify 31 per cent of halos at $z=0$ as relaxed, 3--4 times as many as we would expect to select from a mass-limited sample based on the criteria used to construct our cluster data set (Paper~\citetalias{Mantz1502.06020}{}).} These authors also find a clear correlation between halo concentration and the time since ``formation'', operationally defined as the time when the halo reaches half of its final mass (see also \citealt{Wechsler0108151}). Taking their $z=0$ and $z=1$ results as a guide, our measured concentrations in the range 5--8 place the clusters' formation times approximately between redshifts 1--2. Our selection of the most relaxed systems is also plausibly responsible for a reduced redshift dependence ($\kappa_\zeta$ consistent with zero) compared with all clusters in simulations. The right panel of Figure~\ref{fig:cM} shows the residuals from the best-fitting concentration--mass relation as a function of central cooling luminosity (as a proxy for the importance of non-gravitational physics), with the latter also expressed as a residual from the best-fitting $\mysub{L}{cool}$--mass relation (Equation~\ref{eq:Lcool-scaling}). While the two sets of residuals are not strongly correlated, it may be worth noting the two clusters with large positive residuals in both concentration and cooling luminosity, RX\,J1347.5$-$1145 and SPT-CL\,J2344$-$4242. If the relatively efficient cooling at work in the centers of these clusters has an effect on the concentration of the dark matter as well as the gas, boosting the value of $c$, this would not be represented in the dark-matter-only simulations used to predict the c--M relation. (The closest analogs of these systems in simulations may, of course, still have greater than average concentration.) Concentration--mass estimates based on X-ray data, mostly for low-redshift clusters, show broad agreement with the mass dependence predicted from N-body simulations \citep{Vikhlinin0507092, Voigt0602373, Zhang0603275, Schmidt0610038, Amodeo1604.02163}. More recently, measurements based on weak gravitational lensing have become possible. Weak lensing shear alone can constrain concentrations only for stacks or ensembles of clusters, but individual cluster constraints are possible in combination with strong lensing or magnification. Our own analysis of weak lensing data for a subset of the clusters used in this work finds an ensemble average concentration of $3.0^{+4.4}_{-1.8}$ (Paper~\citetalias{Applegate1509.02162}{}), consistent with our X-ray results here within the statistical uncertainties. Results using strong and weak lensing by X-ray selected clusters \citep{Okabe1302.2728} and from stacked analyses of optically selected clusters \citep{Mandelbaum0805.2552, Du1510.08193, Shan1502.00313, van-Uitert1506.03817} are in broadly good agreement with our results. However, an arguably better comparison is with the lensing analyses of the CLASH sample of clusters ($0.19<z<0.89$), which mostly consists of relaxed or nearly relaxed clusters (by our definition). Using CLASH strong+weak lensing, \citet{Merten1404.1376} found $\kappa_m=-0.32\pm0.18$ and $\kappa_\zeta=-0.14\pm0.52$, consistent with our results; incorporating magnification, \citet{Umetsu1507.04385} found $\kappa_m=-0.44\pm0.19$ while fixing $\kappa_\zeta=-0.668$. We conclude that our concentration--mass estimates for relaxed clusters are in good agreement with simulations and previous measurements, given the expectation that we have selected dynamically similar, and, in particular, the most dynamically relaxed, clusters at all redshifts. Direct tests of that selection on simulated cluster images would not be straightforward, given the inability of current hydrodynamic simulations to reproduce the cores of relaxed clusters. However, the sensitivity of our results to a few strongly cooling clusters motivates doing so in future work, as well as expanding the relaxed cluster sample if possible. \subsection{Generalized NFW Mass Profiles} \label{sec:gnfw} \begin{figure*} \centering \includegraphics[scale=0.9]{fig_gnfw_z} \hspace{1cm} \includegraphics[scale=0.9]{fig_gnfw} \caption[]{ Left: The GNFW inner slope, $\beta$, as a function of redshift, for the subset of our sample where this model can be constrained (see Equation~\ref{eq:profmodels}). When fitting individual clusters, $\beta$ is allowed to vary between 0.0 and 2.0. Right: Joint 68.3 and 95.4 per cent confidence constraints on the average value and Gaussian intrinsic scatter of the inner slope parameter. The NFW model corresponds to a value of $\beta=1$. } \label{fig:gnfw} \end{figure*} We next turn to constraints on the inner slope of the cluster mass profiles, as parametrized by the GNFW model. While a profile of the Einasto form is now known to provide a better description of simulated clusters over a large range in radii, the GNFW model still represents a simple modification of the baseline NFW model, and one that is arguably easier to interpret in light of the limited radial range of our data. N-body simulations generally place values of the inner slope of the GNFW profile, $\beta$ (Equation~\ref{eq:profmodels}), in the range 1.0--1.5, with $\beta=1$ corresponding to the NFW model \citep{Diemand0402267, Diemand0504215, Navarro0311231, Tasitsiomi0311062}. Unfortunately, due to its greater complexity compared to the NFW profile, the GNFW model cannot be fully constrained for all of the clusters in our sample using the current data. In practice, the scale radius parameter becomes unconstrained, taking on very large values, while the inner slope becomes relatively steep, taking values roughly appropriate for the average of the profile in the observed region (i.e., $\sim-2$). A straightforward way to eliminate these cases is to adopt the outer radius where the data can constrain the mass profile, $\mysub{r}{out}$, as an upper limit for the scale radius (when fitting the NFW profile, all scale radii are comfortably within the measured range). Those clusters for which the 95.4 per cent confidence interval for $\mysub{r}{s}$ includes values $>\mysub{r}{out}$ are removed from this portion of the analysis. We have checked that making this cut based on the 68.3 per cent confidence interval for $\mysub{r}{s}$ does not change our results apart from tightening the constraints slightly (because fewer clusters are excluded). Thus, while this procedure at some level introduces the prior that the mass profiles not be too far from NFW, our results are not very sensitive to it. The left panel of Figure~\ref{fig:gnfw} shows these individual constraints (see also Table~\ref{tab:thetable}). Using the 20 clusters that pass this check, we fit a model for the average value of $\beta$ and its cluster-to-cluster scatter, which we assume to be described by a Gaussian distribution with mean $\bar{\beta}$ and width $\sigma_\beta$. The joint constraints on these parameters are shown in the right panel of Figure~\ref{fig:gnfw}, and are consistent with the NFW value of $\beta=1$. The marginalized constraints are $\bar{\beta} = 1.02 \pm 0.08$ and $\sigma_\beta = 0.22 \pm 0.07$. We emphasize that the data used to fit this model do not extend all the way into the cluster centers, although they do include radii down to $\sim 0.1\,\mysub{r}{s}$ (see Section~\ref{sec:data}). Our analysis indicates that, on average, the NFW model provides a good description of the most relaxed clusters down to small radii. However, it also shows significant intrinsic scatter in $\beta$, at the $\sim20$ per cent level, even among the most relaxed clusters. Earlier work by \citet{Schmidt0610038} and \citet{Host0907.1097} using X-ray observations reached similar conclusions on the mean value of $\beta$, respectively finding $\bar{\beta}=0.88^{+0.26}_{-0.31}$ and $0.98<\bar{\beta}<1.19$ (both 95 per cent confidence intervals). Neither of these authors fit for intrinsic scatter explicitly, but their results also pointed to its presence, to the extent that the scatter in individual-cluster constraints on $\beta$ was not consistent with measurement uncertainties alone. More recently, \citet{Newman1209.1391} also found inner mass profile slopes of massive clusters to be consistent with CDM-only predictions, using a combination of strong and weak gravitational lensing constraints, at radii where stars do not dominate the mass budget. \subsection{Einasto Mass Profiles} \label{sec:einasto} \begin{figure*} \centering \includegraphics[scale=0.9]{fig_einasto_z} \hspace{1cm} \includegraphics[scale=0.9]{fig_einasto} \caption[]{ Left: The Einasto shape parameter, $\alpha$, as a function of redshift, for the subset of our sample where this model can be constrained (see Equation~\ref{eq:profmodels}). When fitting individual clusters, $\alpha$ is allowed to vary between 0.01 and 0.75. Right: Joint 68.3 and 95.4 per cent confidence constraints on the average value and Gaussian intrinsic scatter of the shape parameter. While no value of $\alpha$ perfectly corresponds to the NFW model, the values consistent with our data approximately reproduce it over the radial range probed (see Figure~\ref{fig:demo}). } \label{fig:ein} \end{figure*} In detail, the cluster mass profile shapes in simulations have been found to be better described by a continuously changing power-law, such as the Einasto form (Equation~\ref{eq:profmodels}), than by the NFW or GNFW models (e.g.\ \citealt{Navarro0810.1522}). Values of the shape parameter, $\alpha$, seen in simulations are mass and redshift dependent; for massive clusters at $z\ltsim1$, average values of $\sim0.2$--0.3 are typical \citep{Gao0711.0746, Dutton1402.7073, Klypin1411.4001}. While we do not expect our data to be capable of distinguishing the GNFW and Einasto models, due to the limited radial range covered, it is nonetheless interesting to see how values of $\alpha$ measured from the X-ray data compare to these expectations. Similarly to the GNFW case, degeneracies between model parameters prevent us from constraining the Einasto model for every cluster in our sample. We follow the same procedure as in Section~\ref{sec:gnfw} to remove from the analysis clusters where the scale radius of the Einasto model is not constrained. Unlike the GNFW case, the constraints on $\alpha$ for several of the remaining clusters are consistent with either the lower or upper bound of the uniform prior we adopt in the analysis, $0.01\leq\alpha\leq0.75$ (left panel of Figure~\ref{fig:ein}). While this makes our results dependent on the width of this prior, in practice the effect is small; for example, our results below are nearly identical to those we would obtain with a tighter prior of $0.01\leq\alpha\leq0.5$. As in the previous section, we fit a Gaussian model for the intrinsic distribution of Einasto shape parameters, parametrized by $\bar{\alpha}$ and $\sigma_\alpha$. The results are shown in the right panel of Figure~\ref{fig:ein}; the marginalized constraints are $\bar{\alpha} = 0.29 \pm 0.04$ and $\sigma_\alpha = 0.12 \pm 0.04$. Although the intrinsic scatter is relatively large, $\sim 40$ per cent, the mean value of the shape parameter is within the range expected for the most massive galaxy clusters at low redshifts \citep{Klypin1411.4001}. Note that, for this value of $\alpha$, the density profile is relatively similar to NFW over the range probed (Figure~\ref{fig:demo}); hence, this result is consistent with the value $\beta \approx 1$ arrived at when fitting the GNFW model. From their analysis of X-ray data for 11 clusters, \citet{Host0907.1097} found $0.14<\bar{\alpha}<0.26$ at 95 per cent confidence, with a best-fitting value of $\sim0.2$. This is somewhat smaller than our result, although we note that their cluster sample is also lower in mass on average ($M_{200}\sim2\E{14}\ensuremath{\, M_{\odot}}$), such that these results are not necessarily in conflict. \section{Conclusion} \label{sec:conclusions} We present constraints on mass profile models for massive, relaxed galaxy clusters based on an analysis of X-ray data from {\it Chandra}{}. This analysis assumes hydrostatic equilibrium between the ICM and the gravitational potential, and includes data only at cluster radii where we are confident that both departures from equilibrium and systematics due to background modeling are minimal. Assuming the NFW mass profile model, the measured concentration--mass relation has a power-law slope with mass of $\kappa_m=-0.16\pm0.07$, consistent with CDM simulations of cosmological structure formation. The measured relation is consistent with being constant as a function of redshift, a feature that is not seen in simulations; however, this is plausibly the result of our selection of the most dynamically relaxed clusters at a given redshift. Simulations including gas physics, on which the same selection procedure is replicated, would be required to perform a completely fair comparison. We detect an intrinsic scatter of $\sigma_{\ln c}=0.16\pm0.03$ in the concentration--mass relation. Two clear high-concentration outliers from the mean relation also have the largest central cooling luminosities in the sample, suggesting a role for baryonic physics in the scatter. However, the remaining clusters do not support a simple trend between concentration and cooling luminosity at fixed mass. When fitting a GNFW mass profile, where the slope at small radii is a free parameter, we find an average value that is consistent with the NFW model, $\bar{\beta}=1.02\pm0.08$. However, there is significant cluster-to-cluster scatter ($\sigma_\beta=0.22\pm0.07$), even within the relaxed sample of clusters analyzed here. For the Einasto profile, we similarly measure a mean value, $\bar{\alpha}=0.29\pm0.04$, which is consistent with CDM predictions for clusters of the mass studied here. In this case, we measure a larger fractional intrinsic scatter, $\sigma_\alpha=0.12\pm0.04$. Overall, our results confirm that the mass distribution within the most massive halos is in agreement with CDM predictions. There are clear opportunities to improve this analysis in the future, both by finding additional highly relaxed clusters and obtaining deeper X-ray data (especially to better constrain 3-parameter models like the GNFW and Einasto models). Although it is beyond the scope of this work, X-ray data for dynamically relaxed clusters at intermediate radii can potentially be combined with weak lensing (at larger radii, $\ {\raise-.75ex\hbox{$\buildrel>\over\sim$}}\ r_{500}$), and strong lensing or stellar velocity dispersions (at small radii, $\ {\raise-.75ex\hbox{$\buildrel<\over\sim$}}\ 0.5r_{2500}$), in the vein of e.g.\ \citet{Newman1209.1391}. While these additional data sets have their own challenges in the form of projection effects and velocity anisotropy, a careful combination for an appropriately chosen cluster sample potentially provides the most complete possible view of the mass distribution within clusters. \section*{Acknowledgements} We acknowledge support from the U.S. Department of Energy under contract number DE-AC02-76SF00515; from the National Aeronautics and Space Administration (NASA) through {\it Chandra}{} Award Numbers GO8-9118X and TM1-12010X, issued by the {\it Chandra}{} X-ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of NASA under contract NAS8-03060; and from NASA under Grant Number NNX15AE12G. \def \aap {A\&A} \def \aapr {A\&AR} \def \aaps {A\&AS} \def \statisci {Statis. Sci.} \def \physrep {Phys. Rep.} \def \pre {Phys.\ Rev.\ E} \def \sjos {Scand. J. Statis.} \def \jrssb {J. Roy. Statist. Soc. B} \def \pan {Phys. Atom. Nucl.} \def \epja {Eur. Phys. J. A} \def \epjc {Eur. Phys. J. C} \def \jcap {J. Cosmology Astropart. Phys.} \def \ijmpd {Int.\ J.\ Mod.\ Phys.\ D} \def \nar {New Astron. Rev.} \def ARA\&A {ARA\&A} \def AJ {AJ} \def A\&AR {A\&AR} \def ApJ {ApJ} \def ApJL {ApJL} \def ApJS {ApJS} \def \asl {Adv. Sci. Lett.} \def MNRAS {MNRAS} \def Nat {Nat} \def PASJ {PASJ} \def PASP {PASP} \def Sci {Sci} \def Comput.\ Phys.\ Commun. {Comput.\ Phys.\ Commun.} \def Geochim.\ Cosmochim.\ Acta {Geochim.\ Cosmochim.\ Acta} \def Nucl.\ Phys.\ A {Nucl.\ Phys.\ A} \def Phys.\ Lett.\ B {Phys.\ Lett.\ B} \def Phys.\ Rev.\ C {Phys.\ Rev.\ C} \def Phys.\ Rev.\ D {Phys.\ Rev.\ D} \def Phys.\ Rev.\ Lett. {Phys.\ Rev.\ Lett.} \def Space Sci.\ Rev. {Space Sci.\ Rev.}
0e6a1f6c3f944e84bed19aedc33dedb9ac6c3643
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \label{sect_intro} \begin{figure*}[ht] \vskip 0cm \begin{center} \includegraphics[width=0.5\textwidth]{fig1.pdf}\end{center} \caption{ Centrality dependence of $v_{2}$ scaled by number of quarks and participant eccentricity ($v_2/$($n_q$$\varepsilon_{\mathrm{part}}\{2\}$)) for ${K}^{0}_{S}$ and $\Lambda$ as a function of ($m_T - m$)/$n_q$ in $0-10$\%, $10-40$\% and $40-80$\% Au+Au collisions and $0-20$\% and $20-60$\% Cu+Cu collisions at \mbox{$\sqrt{\mathrm{\it s_{NN}}}$} = 200 GeV~\cite{cucu}. } \label{figure1} \end{figure*} \begin{figure*}[ht] \vskip 0cm \begin{center} \includegraphics[width=0.5\textwidth]{fig2.pdf}\end{center} \caption{The $v_{2}$ as function of $p_{T}$ for (a) $\pi$, proton and (b) $\phi$, $\Omega$ in Au+Au collisions at $\sqrt{s_{NN}}$ = 200 GeV for $0-80\%$ centrality~\cite{paper}.} \label{figure2} \end{figure*} \begin{figure}[t] \vskip 0cm \includegraphics[width=0.5\textwidth]{fig3.pdf} \caption{The ratio of $v_{2}(\phi)$ to $v_{2}(p)$ as function of $p_{T}$ in Au + Au collisions at $\sqrt{s_{NN}}$ = 200 GeV for two centralities, $0-30\%$ and $30-80\%$ The bands in panel (a) and (b) represent the hydro and transport model calculations for $v_{2}(\phi)/v_{2}(p)$, respectively~\cite{paper}. } \label{figure3} \end{figure} \begin{figure*}[ht] \vskip 0cm \begin{center} \includegraphics[width=0.7\textwidth]{fig4.pdf}\end{center} \caption{$v_{2}$ scaled by number of constituent quarks ($n_{q}$) as a function of $p_{T}/n_{q}$ and $(m_{T}-m_{0})/n_{q}$ for identified hadrons from Au + Au collisions at $\sqrt{s_{NN}}$ = 200 GeV for two centralities, $0-30\%$ and $30-80\%$. Ratios with respect to a polynomial fit to $K^{0}_{S}$ $v_{2}$ are shown in the corresponding lower panels~\cite{paper}. } \label{figure4} \end{figure*} \begin{figure*}[ht] \vskip 0cm \begin{center} \includegraphics[width=0.6\textwidth]{fig5.pdf}\end{center} \caption{$v_{2}$ scaled by number of constituent quarks ($n_{q}$) as a function of $p_{T}/n_{q}$ for identified hadrons from Pb + Pb collisions at $\sqrt{s_{NN}}$ = 2.76 TeV for various centrality intervals~\cite{ALICE_flow}.} \label{figure5} \end{figure*} \begin{figure}[t] \vskip 0cm \includegraphics[width=0.5\textwidth]{fig6.pdf} \caption{The difference in $v_2$ between particles ($X$) and their corresponding antiparticles ($\bar{X}$) as a function of beam energy for $0-80$\% central Au + Au collisions~\cite{BES2, BES3}.} \label{figure6} \end{figure} \begin{figure}[t] \vskip 0cm \includegraphics[width=0.5\textwidth]{fig7.pdf} \caption{The viscous hydrodynamical calculations without a hadronic cascade afterburner (VISH2+1) and with a hadronic cascade afterburner(VISHNU)~\cite{raimond, Song2011, Song2014, Song2015} } \label{figure7} \end{figure} \begin{figure}[t] \vskip 0cm \includegraphics[width=0.5\textwidth]{fig8.pdf} \caption{The comparison of ALICE measurements and VISHNU model calculations of $v_2$ as a function of $p_T$ at $\sqrt{s_{NN}}$ = 2.76 TeV for $10-20$\% central collisions~\cite{ALICE_flow, raimond, Song2011, Song2014, Song2015}.} \label{figure8} \end{figure} At the early stage of high energy relativistic heavy ion collisions, a hot and dense, strongly interacting medium named Quark Gluon Plasma (QGP) is created~\cite{rhicwp1, rhicwp2}. The subsequent system evolution is determined by the nature of the medium. Experimentally, the dynamics of the system evolution has been studied by measuring the azimuthal anisotropy of the particle production relative to the reaction plane~\cite{flow1, flow2, review}. The centrality of the collision, defined by the transverse distance between the centers of the colliding nuclei called the impact parameter, results in an `almond-shaped' overlap region that is spatially azimuthal anisotropic. It is generally assumed that the initial spatial anisotropy in the system is converted into momentum-space anisotropy through re-scatterings~\cite{rescatterings}. The elliptic flow, $v_{2}$, which is the second Fourier coefficient of the azimuthal distribution of produced particles with respect to the reaction plane, is defined as $v_{2}=\langle\cos 2(\varphi-\Psi)\rangle$, where $\varphi$ is the azimuthal angle of produced particle and $\Psi$ is the azimuthal angle of the reaction plane. The initial anisotropy in the coordinate space diminishes rapidly as the system expands. Thus, the driving force of $v_2$ quenches itself. Due to the self-quenching effect, the elliptic flow provides information about the dynamics at the early stage of the collisions~\cite{physics1, physics2, physics3}. Elliptic flow can provide information about the pressure gradients, the effective degrees of freedom, the degree of thermalization, and equation of state of the matter created at the early stage~\cite{review}. However, early dynamic information might be obscured by later hadronic rescatterings~\cite{hyrdo_cascade1, hyrdo_cascade2}. Strange hadrons, especially multi-strange hadrons and the $\phi$ meson are believed to be less sensitive to hadronic rescatterings in the late stage of collisions, as their freeze-out temperatures are close to the phase transition temperature and their hadronic interaction cross sections are expected to be small~\cite{multistrange1, multistrange2}. In this paper, I am going to review the elliptic flow results of strange and multi-strange hadron in relativistic heavy ion collisions from RHIC to LHC energies. \section{Discussions} \subsection{Centrality and system size dependence} The values of $v_2$ are usually divided by the initial spatial anisotropy, eccentricity, to remove the geometric effect in order to study the centrality and system size dependence of $v_2$. The participant eccentricity is the initial configuration space eccentricity of the participants which is defined by~\cite{ecc1,cucu} \begin{equation} \varepsilon_{\rm part} = \frac{\sqrt{(\sigma_{y}^{2} - \sigma_{x}^{2})+4(\sigma_{xy}^{2})}}{\sigma_{y}^{2} + \sigma_{x}^{2}}, \label{Equation:partiecc} \end{equation} In this formula, $\sigma_{x}^{2} = \langle x^{2} \rangle - \langle x\rangle^{2}$, $\sigma_{y}^{2} = \langle y^{2} \rangle - \langle y\rangle^{2}$ and $\sigma_{xy} = \langle xy \rangle - \langle x \rangle\langle y \rangle$, with $x$, $y$ being the position of the participating nucleons in the transverse plane. The root mean square of the participant eccentricity \begin{equation} \varepsilon_{\rm part}\{2\} = \sqrt{\langle\varepsilon_{\rm part}^{2}\rangle}, \label{Equation:epart2} \end{equation} is calculated from the Monte Carlo Glauber model~\cite{glauber} and Color Glass Condensate (CGC) model~\cite{CGC}. Figure~\ref{figure1} shows the centrality and system size dependence of ${K}^{0}_{S}$ and $\Lambda$ $v_2$ in $\sqrt{s_{NN}}$ = 200 GeV heavy ion collisions~\cite{cucu}. The eccentricity scaled $v_2$ has been further normalized by number of constituent quark ($n_q$) to make ${K}^{0}_{S}$ and $\Lambda$ results follow a same curve. The results from $0-20$\% and $20-60$\% central Cu + Cu collisions and from $0-10$\%, $10-40$\% and $40-80$\% central Au + Au collisions are presented. For a given collision system, stronger collectivity flow is apparent as higher scaled $v_{2}$ values in more central collisions. For both Au + Au and Cu + Cu collisions, larger collective flow is observed in larger system size which could be characterized by number of participants. Namely, the collisions with larger number of participants generate larger collective flow. \subsection{Multi-strange hadron and $\phi$ meson $v_2$} STAR experiment presented the first $v_2$ results of multi-strange hadrons based on $2 \times 10^{6}$ events collected in the year of 2001 - 2002~\cite{STAR_multi-strange}. Significant $v_2$ signals of $\Xi$ baryons which are similar to results for $\Lambda$ baryons are observed in Au + Au collisions at $\sqrt{s_{NN}}$ = 200 GeV. At low $p_T$ ($\mbox{$<~$} $ 2 GeV/$c$), the mass ordering is observed for $\Xi$ $v_2$ which is in agreement with the hydrodynamic model calculations. Due to limited statistics, the $v_2$ of $\Omega$ baryons have large statistical uncertainties, it is not clear whether $\Omega$ $v_2$ follows baryon or meson band at the intermediate $p_T$ range(2 - 5 GeV/$c$). But non-zero value of $v_2$ was clearly observed at that time. These results suggest that collective motion has been developed at parton phase in Au + Au collisions at $\sqrt{s_{NN}}$ = 200 GeV. Later, in the RHIC runs of the year 2010-2011, about 730 million minimum bias events were recorded by STAR. Sufficient statistics of multi-strange hadrons and $\phi$ mesons support the precise measurements on $v_2$. The multi-strange hadrons and the $\phi$ meson were reconstructed though the following decay channels: $\phi$ $\rightarrow$ $\it{K}^{+}$ + $\it{K}^{-}$, $\Xi^{-}$ $\rightarrow$ $\Lambda$ + $\pi^{-}$ ($\overline{\Xi}^{+}$ $\rightarrow$ $\overline{\Lambda}$ + $\pi^{+}$) and $\Omega^{-}$ $\rightarrow$ $\Lambda$ + $\it{K}^{-}$ ($\overline{\Omega}^{+}$$\rightarrow$ $\overline{\Lambda}$ + $\it{K}^{+}$). Figure~\ref{figure2} shows the $v_{2}$ as a function of $p_{T}$ for (a) $\pi$, protons and (b) $\phi$, $\Omega$ in Au + Au collisions at $\sqrt{s_{NN}}$ = 200 GeV for $0-80\%$ centrality~\cite{paper, paper1}. A comparison between $v_{2}$ of $\pi$ and protons, consisting of up ($\it u$) and down ($\it d$) light constituent quarks is shown in panel (a). Correspondingly, panel (b) shows a comparison of $v_{2}$ of $\phi$ and $\Omega$ containing $\it s$ constituent quarks. This is the first time that high precision measurement of $\Omega$ baryon $v_2$ up to 4.5 GeV/$c$ is available in experiments of heavy ion collisions. In the low $p_{T}$ region ( $p_{T}$ $\mbox{$<~$} $ 2.0 GeV/$c$), the $v_{2}$ of $\phi$ and $\Omega$ follows mass ordering. At intermediate $p_{T}$ ( 2.0 $\mbox{$<~$} $ $p_{T}$ $\mbox{$<~$} $ 5.0 GeV/$c$), a baryon-meson separation is observed. The $v_2$ results of $\phi$ mesons are consistent in two independent measurements at RHIC, PHENIX~\cite{PHENIX_phi} and STAR. It is evident that the $v_{2}(p_{T}) $ of hadrons consisting only of strange constituent quarks ($\phi$ and $\Omega$) is similar to that of light hadrons, $\pi$ and protons. However the $\phi$ and $\Omega$ do not participate strongly in the hadronic interactions, because of the smaller hadronic cross sections compared to $\pi$ and protons. It suggests the major part of the collectivity is developed during the partonic phase in high energy heavy ion collisions. ALICE experiment recently published multi-strange hadron and $\phi$ meson $v_2$ measurements in Pb + Pb collisions at $\sqrt{s_{NN}}$ = 2.76 TeV~\cite{ALICE_flow}. Also significant $v_2$ values for these particles are observed. Experimental measurements at RHIC and LHC indicates partonic collectivity has been built up in high energy heavy ion collisions. \subsection{Comparison of $\phi$ meson and proton $v_2$} The $\phi$ meson and proton show different sensitivity on the hadronic rescatterings. As discussed previously, the $\phi$ meson is less sensitive to the late hadron hadron interactions than light hadrons due to the smaller hadronic cross section. It means light hadrons (e.g. protons) would gain larger additional radial flow which modifies the $v_2$($p_T$) shape during final hadronic rescatterings. Hydrodynamical model calculations predict that $v_{2}$ as a function of $p_{T}$ for different particle species follows mass ordering, where the $v_{2}$ of heavier hadrons is lower than that of lighter hadrons~\cite{hydro}. The identified hadron $v_{2}$ measured in experiment indeed proves the mass ordering in the low $p_{T}$ region ($p_{T}$ $\mbox{$<~$} $ 2.0 GeV/$c$). Hirano {\it et al.} predict the mass ordering of $v_{2}$ could be broken between $\phi$ mesons and protons at low $p_{ T}$ ($p_{T}$ $\mbox{$<~$} $ 1.5 GeV/$c$) based on a model with ideal hydrodynamics plus hadron cascade process~\cite{hyrdo_cascade1, hyrdo_cascade2}. Here $\phi$ mesons and protons are chosen for the study, as their rest masses are quite close to each other. As the model calculations assign a smaller hadronic cross section for $\phi$ mesons compared to protons, the broken mass ordering is regarded as the different hadronic rescattering contributions on the $\phi$ meson and proton $v_2$. Figure~\ref{figure3} shows the ratios of $\phi$ $v_{2}$ to proton $v_{2}$ from model calculations and experimental data~\cite{paper, paper1}. This ratio is larger than unity at $p_{T}$ $\sim$ 0.5 GeV/$c$ for 0-30$\%$ centrality. It indicates breakdown of the expected mass ordering in that momentum range. This could be due to a large effect of hadronic rescatterings on the proton $v_{2}$. The data of 0-80\% centrality around 0.5 GeV/$c$ is quantitatively agrees with hydro + hadron cascade calculations indicated by the shaded red band in panel (a) of Fig.~\ref{figure3}, even though there is a deviation in higher $p_T$ bins. A centrality dependence of $v_{2}(\phi)$ to $v_{2}(p)$ ratio is observed in the experimental data. Namely, the breakdown of mass ordering of $v_{2}$ is more pronounced in 0-30$\%$ central collisions than in 30-80$\%$ peripheral collisions. In the central events, both hadronic and partonic interactions are stronger than in peripheral events. Therefore, the larger effect of late stage hadronic interactions relative to the partonic collectivity produces a greater breakdown of mass ordering in the 0-30$\%$ centrality data than in the 30-80$\%$. This observation indirectly supports the idea that the $\phi$ meson has a smaller hadronic interaction cross section. The ratio of $\phi$ $v_{2}$ to proton $v_{2}$ was also studied by using the transport models AMPT~\cite{ampt} and UrQMD~\cite{urqmd}. The panel (b) of Fig.~\ref{figure3} shows the $v_{2}(\phi)$ to $v_{2}(p)$ ratio for 0-30$\%$ centrality from AMPT and UrQMD models. The black shaded band is from AMPT with a hadronic cascade time of 0.6 fm/$c$ while the yellow band is for a hadronic cascade time of 30 fm/$c$. Larger hadronic cascade time is equivalent to stronger hadronic interactions. It is clear that the $v_{2}(\phi)/v_{2}(p)$ ratio increases with increasing hadronic cascade time. This is attributed to a decrease in the proton $v_{2}$ due to an increase in hadronic re-scattering while the $\phi$ meson $v_{2}$ is less affected. The ratios from the UrQMD model are much smaller than unity (shown as a brown shaded band in the panel (b) of Fig.~\ref{figure3}). The UrQMD model lacks partonic collectivity, thus the $\phi$ meson $v_{2}$ is not fully developed. Non of these models could describe the detailed shape of the $p_T$ dependence. In $\sqrt{s_{NN}}$ = 2.76 TeV Pb + Pb collisions at LHC, there is an indication that the $\phi$ meson $v_2$ is larger than the proton $v_2$ for the lowest $p_T$ bin~\cite{ALICE_flow, raimond}. Unfortunately, currently the uncertainties on the ALICE $\phi$ meson $v_2$ measurements are too large to conclude. \subsection{Number of constituent quark scaling} The Number of Constituent Quark (NCQ) scaling in $v_{2}$ in the intermediate $p_{T}$ range (2 $\mbox{$<~$} $ $p_T$ $\mbox{$<~$} $ 5 GeV/$c$) could be well reproduced by the quark coalescence~\cite{cola} or recombination~\cite{recom} mechanisms in particle production. The NCQ scaling indicates the collectivity in the parton level has been achieved in high energy heavy ion collisions at RHIC. Figure~\ref{figure4} shows number of constituent quarks ($n_{q}$) scaled $v_{2}$ as a function of transverse momentum scaled by $n_{q}$ ($p_{T}/n_{q}$) and transverse mass minus rest mass scaled by $n_{q}$ ($(m_{T}-m_{0})/n_{q}$) for identified hadrons from Au + Au collisions at $\sqrt{s_{NN}}$ = 200 GeV for two centralities, 0-30$\%$ and 30-80$\%$. To investigate the possible system size dependence deviation from NCQ scaling, the $K^{0}_{S}$ $v_{2}$ was fitted with a third-order polynomial function. Then the ratio to the $K^{0}_{S}$ fit was calculated. The lower panels of Fig.~\ref{figure4} show the results. Excluding pions, the scaling holds approximately within 10$\%$ for both 0-30$\%$ and 30-80$\%$ centralities. The pion is excluded as it is strongly affected by resonance decay process and non-flow correlations~\cite{dongx}. Figure~\ref{figure5} shows NCQ scaling at LHC energy. The maximum deviation from NCQ scaling is $\sim$20$\%$ at $\sqrt{s_{NN}}$ =2.76 TeV as observed by ALICE experiment~\cite{ALICE_flow}. Therefore, at top RHIC energy, NCQ scaling holds better than LHC energy. Recently, CMS collaboration presented the $v_2$ results of strange hadrons ($K_{S}^{0}$ and $\Lambda$) in $p$ + Pb collisions at $\sqrt{s_{NN}}$ = 5.02 TeV with event sample of large multiplicity~\cite{CMS1, CMS2}. A nice NCQ scaling (less than 10\% violation) is observed. It indicates the partonic level collectivity has been built up even in small $p$ + Pb colliding system. It would be interesting to compare the NCQ scaling using event samples with large and small multiplicity in the future. \subsection{Beam energy dependence} STAR experiment has covered the beam energies of $\sqrt{s_{NN}}$ = 7.7, 11.5, 14.5, 19.6, 27, 39, 62.4 and 200 GeV. During 2010 - 2014, a Beam Energy Scan program (phase I) was carried out at RHIC. The main motivation is to explore the nuclear matter phase structure in the higher net-baryon region. The most striking feature on the $v_2$ measurements is the observation of an energy dependent difference in $v_2$ between particles and their corresponding antiparticles~\cite{{BES1, BES2}}. Figure~\ref{figure6} shows the difference in $v_2$ between particles and their corresponding antiparticles as a function of beam energy. The difference between baryon and anti-baryon is much more pronounced than difference between mesons. Proton versus anti-proton and $\Lambda$ versus $\bar{\Lambda}$ show same magnitude of difference. This difference naturally breaks the number of constituent quark scaling (NCQ) in $v_2$ which is regarded as an evidence of partonic collectivity in the top energy heavy ion collisions at RHIC. It indicates the hadronic degrees of freedom play a more important role at lower collision energies. The data have also been compared to hydrodynamics + transport (UrQMD) hybrid model~\cite{hybrid} and Nambu-Jona-Lasino (NJL) model~\cite{NJL} which considers both partonic and hadronic potential. The hybrid model could reproduce the baryon (proton) data, but fails to explain the mesons; whereas the NJL model could qualitatively reproduce the hadron splitting. However, even if one tunes the $R_v$ parameter which is related to the partonic potential, NJL model fails to reproduce the magnitude for all hadron species simultaneously. Analytical hydrodynamic solution can reproduce the data within uncertainties~\cite{hydro_solution}. It predicts $\Delta v_{2}^{p} \mbox{$>~$} \Delta v_{2}^{\Lambda} \mbox{$>~$} \Delta v_{2}^{\Xi} \mbox{$>~$} \Delta v_{2}^{\Omega}$ for baryons. Future high precise data will clarify the validity of this description. \subsection{Comparison with hydrodynamic calculations} The $p_T$ differential $v_2$ could be modified by an increase on both collective and radial flow with increasing of colliding energy. It is qualitatively described by hydrodynamic calculations~\cite{Song2011}. The recent comparison between ALICE measurements and model calculations shows a nice agreement in $40-50$\% central collisions including strange baryon $\Lambda$ and multi-strange baryon $\Xi$. However, for more central collisions (e.g. $10-20$\%) a clear discrepancy is observed for protons, $\Lambda$ and $\Xi$~\cite{ALICE_flow}. Later, it was realized the hadronic rescatterings is important to be included in the hydrodynamic calculations for a fair comparison between data and models~\cite{Song2014}. In Fig.~\ref{figure7}, viscous hydrodynamical calculations with(VISHNU) and without (VISH2+1) a hadronic cascade afterburner are compared. The increase in mass splitting between identified particles for {\sc VISHNU} (solid curves) compared to {\sc VISH2+1} (dashed curves) illustrates the larger radial flow in the {\sc VISHNU} calculations due to the contribution of the hadronic cascade. The mass splitting between the pions and strange baryons ($\Lambda$) / multi-strange baryons ($\Xi$) does not change much, as small hadronic rescattering cross sections are assigned to these particles. The mass ordering observed in pure viscous hydrodynamical calculations is not preserved anymore between protons and strange baryons ($\Lambda$) / multi-strange baryons ($\Xi$) after including the hadronic interactions in VISHNU. Figure~\ref{figure8} shows the comparison between the $p_T$-differential $v_2$ measured by ALICE and the {\sc VISHNU} model. Even though VISHNU gives a very well description of kaons; clear discrepancy for protons, $\Lambda$ and $\Xi$ is observed. The VISHNU calculations under-predict the $v_2$ of protons and over-predict the $v_2$ of $\Lambda$ and $\Xi$. Obviously, the current theoretical framework of viscous hydrodynamics plus a hadron cascade afterburner does not describe the $v_2$ as a function of $p_T$ for identified particles in more central collisions better. One of the possible reasons is that hadronic interaction process for some particle species might not be well understood. \section{Summary} \label{sect_summary} In this paper, I review the elliptic flow results of strange and multi-strange hadrons in relativistic heavy ion collision from RHIC to LHC energies. The centrality and system size dependence of $v_2$ could be described by number of participants in both Au+Au and Cu+Cu collisions at $\sqrt{s_{NN}}$ = 200 GeV. The precise measurements of multi-strange hadron $v_2$, especially for the $\Omega$ baryons indicates the collectivity has been built-up in the early partonic stage of collisions. The comparison between the $v_2$ of $\phi$ mesons and protons shows a possible violation of hydrodynamics inspired mass ordering in $0-30$\% central collisions. It can be qualitatively explained by the different effects of late hadronic interactions on the $\phi$ meson and proton $v_2$. The NCQ scaling of identified particles in top energy heavy ion collisions at RHIC is better than LHC energy suggesting that coalescence might be the dominant hadronization mechanism at RHIC in the intermediate transverse momentum region (2 $\mbox{$<~$} $ $p_T$ $\mbox{$<~$} $ 5 GeV/$c$). Also, the NCQ scaling is observed in small colliding system, $p$ + Pb, at $\sqrt{s_{NN}}$ = 5.02 TeV. It indicates the partonic level of collectivity has also been reached in high energy $p$ + Pb collisions. At lower beam energy ($<$ $\sqrt{s_{NN}}$ = 39 GeV), a difference is observed between $v_2$ values of particles and anti-particles. Currently there is no theoretical framework can reproduce the data quantitatively. The recent comparison between viscous hydrodynamic calculations with a hadronic cascade afterburner and experimental data shows a discrepancy on the baryons which challenges the current knowledge on the hadronic interactions. \section{Acknowledgments} This work was supported in part by National Basic Research Program of China (973 program) under grand No. 2015CB8569, the National Natural Science Foundation of China under grant No. 11475070 and self-determined research funds of CCNU from the colleges' basic research and operation of MOE under grand No. CCNU15A02039.
c0565b9e1ef0d4e3fd581743098aebe91bd7ae99
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \subsection{} The notion of projective and affine geometry plays a crucial role in the emerging theory of $\mathbb{F}_1$, the field with one element. Combinatorially, following Tits \cite{anal}, such a projective space $\mathbb{P}$ of dimension $m$ is a complete graph on $m + 1$ vertices endowed with the full subgraph structure. Its automorphism group is isomorphic to the symmetric group $\texttt{S}_{m + 1}$, and the latter is the Weyl group of the projective general linear group of any projective space of dimension $m$ over any field. In \cite{KapranovUN}, Kapranov and Smirnov describe the same theory from the viewpoint of $\mathbb{F}_1$-vector spaces, and they end up with $\texttt{S}_{m + 1}$ as the group of $(m + 1)\times(m + 1)$-permutation matrices acting as linear automorphisms. In Deitmar's scheme theory \cite{Deitmarschemes2}, an affine space (of dimension $m$) over $\mathbb{F}_1$ is defined as $\texttt{Spec}(\mathbb{F}_1[Y_1,\ldots,Y_m])$, where $\mathbb{F}_1[Y_1,\ldots,Y_m]$ is the free abelian monoid generated by $Y_1,\ldots,Y_m$ (usually also containing an extra element $0$), and the $\texttt{Spec}$-construction generalizes naturally from commutative unital rings to commutative multiplicative monoids with an extra absorbing element $0$. As noticed by the author in (e.g.) \cite{KT-Chap-Mot} | see also \cite{LPL} | there is also a natural $\texttt{Proj}$-construction in this theory, and hence $\texttt{Proj}(\mathbb{F}_1[Y_1,\ldots,Y_{m + 1}])$ is the ``Algebro-Geom\-etric version'' of $\mathbb{P}$. Both notions fit perfectly together: the $m + 1$ closed points of $\texttt{Proj}(\mathbb{F}_1[Y_1,\ldots,Y_{m + 1}])$ correspond to the vertices of the graph, and linear subspaces of dimension $1$ indeed contain precisely two closed points. \subsection{} In most of the known scheme theories in characteristic $1$, a scheme essentially consists of a triple $(S,X,\gamma)$, where $S$ is a Deitmar scheme, $X$ a Grothendieck scheme, and $\gamma$ a map $S \mapsto X$ which comes from a form of base extension from $\mathbb{F}_1$ to $\mathbb{Z}$. (See for instance Deitmar \cite{Deitmarschemes2}, Connes and Consani \cite{ConCon}, and Thas \cite{KT-Chap-Mot}.) As such, Deitmar schemes play a very important role in Absolute Arithmetic. In general, many phenomena which are invisible at the $\mathbb{F}_1$-level, only become visible after having applied a base extension functor (say, to a commutative field) | Manin coined this principle with the term ``acquiring flesh.'' This principle is essential for the present note and we study it in the guise of the ``Hidden Points Principle.'' \subsection{} \label{issue} In \cite{KapranovUN}, the authors define the ``field extension'' $\mathbb{F}_{1^{\ell}} \Big/ \mathbb{F}_1$ (for any integer $\ell \geq 1$) as the cyclic group $\mu_{\ell}$ with an extra element $0$, no addition and usual multiplication. In coordinates, elements of the vector space $V(m + 1,\mathbb{F}_{1^{\ell}})$ are $(m + 1)$-tuples with at most one nonzero entry $a \in \mathbb{F}_{1^{\ell}}$; there is one vector (the zero vector), and there are $(m + 1)\ell$ vector lines (which are not necessarily different | see \S\S \ref{ghost}), and each contains only the zero vector. Everything passes in a straightforward way to affine spaces, and there arises a multiple occurrence of the aforementioned principle: \begin{itemize} \item[{\bf Aff.}] according to Deitmar's theory, affine $\mathbb{F}_1$-schemes only have one closed point, and the number of lines equals the dimension; \item[{\bf Proj.}] in the $\texttt{Proj}$-setting, the topology of $\texttt{Proj}(\mathbb{F}_1[Y_1,\ldots,Y_{m + 1}])$ and\\ $\texttt{Proj}(\mathbb{F}_{1^{\ell}}[Y_1,\ldots,Y_{m + 1}])$ is the same, so the latter also has $m + 1$ closed points. \end{itemize} \begin{remark}{\rm Note that \begin{itemize} \item[$\odot$] the number of vectors of $V(n,q^\ell)$ with $q \mapsto 1$ is $\lim_{q \mapsto 1}q^{n\ell} = 1$; \item[$\odot$] the number of vector lines of $V(n,q^\ell)$ with $q \mapsto 1$ is $\lim_{q \mapsto 1}\frac{q^{n\ell} - 1}{q^\ell - 1} = n$; \item[$\odot$] the number of lines of $\mathbf{AG}(n,q^\ell)$ with $q \mapsto 1$ is $\lim_{q \mapsto 1}\frac{q^{n\ell}(q^{n\ell} - 1)}{q^\ell(q^\ell - 1)} = n$. \end{itemize} } \end{remark} On the other hand, according to Kurokawa \cite{Kurozeta}, a $\mathbb{Z}$-scheme $\mathcal{X}$ is {\em of $\mathbb{F}_1$-type} if it comes with a counting polynomial $N_{\mathcal{X}}(T) \in \mathbb{Z}[T]$ (such that for any finite field $\mathbb{F}_q$, we have that $\vert \mathcal{X}_q \vert = N_{\mathcal{X}}(q)$). According to Connes and Consani \cite{ConCon} (see also Deitmar \cite{Deitmarschemes2} and Manin and Marcolli \cite{MM}), we should have that \begin{itemize} \item[{\bf Count.}] The equality $\vert \mathcal{X}_{\mathbb{F}_{1^{\ell}}} \vert = N(\ell + 1)$ holds. \end{itemize} So affine and projective spaces over $\mathbb{F}_{1^{\ell}}$ in Deitmar's setting have a number of hidden points which should be manifested through the counting polynomial, and which are invisible for the Zariski topology at the $\mathbb{F}_1$-level. (Put more bluntly: {\em the combinatorial side and the algebro-geometric side do not agree on field extensions}.) In this note, we present a way to handle this problem, and show that the algebro-geometric side and the combinatorial one actually {\em do} agree. \subsection{Acknowledgment} The author wants to express his gratitude to Nobushige Kurokawa for several very helpful communications about the paper \cite{Kurozeta}. \medskip \section{The functor of points} Let $k$ be a field, and consider an affine algebraic variety $V(\mathcal{A})$ with coordinate ring $A = k[t_1,\ldots,t_m]/\mathcal{A}$. Let $\gamma: k[t_1,\ldots,t_m] \mapsto A$ be the natural projection with kernel $\mathcal{A}$. Define $u_i := \gamma(t_i)$ for all $i$. If \begin{equation} \varphi:\ A\ \mapsto\ k \end{equation} is any morphism, then $(\varphi(u_1),\ldots,\varphi(u_m))$ is a $k$-point of $V(\mathcal{A})$. Vice versa, let $\overline{x} = (x_1,\ldots,x_m)$ be any $k$-point of $V(\mathcal{A})$. Then the following map defines a morphism $A \mapsto k$: \begin{equation} \varphi_{\overline{x}}:\ f\ \mapsto\ f(\overline{x}). \end{equation} \subsection*{Correspondence} For any morphism $\varphi: A \mapsto k$, we have that \begin{equation} \varphi \ \equiv \ \varphi_{\overline{x}}, \end{equation} where $\overline{x} = (\varphi(u_1),\ldots,\varphi(u_m))$. \\ This classical insight, which sees $k$-points as $k$-morphisms, also works for schemes in general. This is the motivation for the definition of {\em $\mathbb{F}_{1^\ell}$-rational point} of an $\mathbb{F}_1$-scheme $\mathcal{X}$, which is used in, e.g., \cite{ConCon}: \begin{equation} \mathcal{X}(\mathbb{F}_{1^\ell}) = \texttt{Hom}(\texttt{Spec}(\mathbb{F}_{1^\ell}),\mathcal{X}). \end{equation} When $\mathcal{X} = \texttt{Spec}(A)$ is an affine $\mathbb{F}_1$-scheme, we also have \begin{equation} \mathcal{X}(\mathbb{F}_{1^\ell}) \cong \texttt{Hom}(A,\mathbb{F}_{1^\ell}). \end{equation} \medskip \section{Spaces over $\mathbb{F}_{1^{\ell}}$} \label{spa-ext} It is our goal to find a common picture to the different models of affine and projective $\mathbb{F}_{1^{\ell}}$-spaces. Our starting point is the projection map \begin{equation} V(m + 1,\mathbb{F}_{1^{\ell}})\ \ \mapsto\ \ \texttt{Proj}(\mathbb{F}_{1^\ell}[Y_1,\ldots,Y_{m + 1}]). \end{equation} Following Kapranov and Smirnov \cite{KapranovUN} and \cite{KT-Chap-Comb}, $V = V(m + 1,\mathbb{F}_{1})$ has only one vector (corresponding to the zero vector), and $m + 1$ directions, which after base extension to a field, become vector lines. Each is given a coordinate of type $(-,1,-)$, and the linear maps act on these directions. Defining $\mathbf{PG}(m,\mathbb{F}_1)$ from $V$ in the usual way, each direction defines a projective point and the zero vector gets lost, so that automorphisms of $V$ carry over in a faithful manner to automorphisms of $\mathbf{PG}(m,\mathbb{F}_1)$. \subsection{Over $\mathbb{F}_{1^{\ell}}$ (ghost directions)} \label{ghost} Kapranov and Smirnov consider all directions of the form $(-,\alpha,-)$ with $\alpha \in \mu_{\ell}$ to define $V(m + 1,\mathbb{F}_{1^\ell})$. In fact, as any direction corresponds to a vector line, no new directions are introduced, but rather cyclotomic coordinates get into the picture. Making the transition to projective space as above, we get the same set of points (see also Proposition \ref{topell1}), as proportionality kills the cyclotomy (on the geometric level). If $X$ is a $k$-scheme, where $k$ is a field, then the closed points of $X$ represent orbits of the Galois group $\texttt{Gal}(\overline{k}/k)$, and all $\overline{k}$-rational points of $X \times_{k}\overline{k}$ are contained in the union of these orbits. So there is a natural map \begin{equation} \alpha: X(\overline{k}) \longrightarrow X, \end{equation} sending closed points of $X(\overline{k})$ to closed points of $X$, which is neither injective nor surjective in general. Going back to the spaces $\mathbf{PG}(n,\mathbb{F}_{1^\ell})$ of above, we see the closed points of $\mathbf{PG}(n,\mathbb{F}_{1^\ell})$ as orbits of $\texttt{Gal}(\overline{\mathbb{F}_1}/\mathbb{F}_{1^\ell})$. A stalk at an arbitrary closed point of $\mathbf{PG}(n,\mathbb{F}_{1^\ell})$ is isomorphic to \begin{equation} \mu_\ell \times \mathbb{F}_1[X_1,\ldots,X_n]. \end{equation} So on the algebraic level we can see the extension of the ground field | we consider the stalk as consisting, besides $0$, of $\ell$ distinct copies of $\mathbb{F}_1[X_1,\ldots,X_n] \setminus \{0\}$ equipped with a sharply transitive $\mu_\ell$-action, which is in accordance with the classical picture. Similar conclusions can be made for the affine spaces $\texttt{Spec}(\mathbb{F}_{1^\ell}[X_1,\ldots,\\ X_n])$ (where the cyclotomic directions are ghost directions defined on the associated vector space, but are not defined in the Deitmar scheme). For any finite field $\mathbb{F}_p$, where $p = 1$ is allowed, we have a natural identification \begin{equation} V(mn,\mathbb{F}_p)\ \overset{\sim}{\mapsto}\ V(m,\mathbb{F}_{p^n}), \end{equation} and so we have \begin{equation} V(m,\mathbb{F}_{1^\ell})\ \overset{\sim}{\mapsto}\ V(m\ell,\mathbb{F}_1)\ \overset{\otimes \mathbb{F}_p}{\mapsto}\ V(m\ell,\mathbb{F}_p)\ \overset{\sim}{\mapsto}\ V(m,\mathbb{F}_{p^\ell}), \end{equation} which justifies Kapranov and Smirnov's approach. Here, the base extension ``$\otimes_{\mathbb{F}_1}\mathbb{F}_p$'' can be defined using, e.g., the affine space: with $A = \mathbb{F}_1[X_1,\ldots,X_{m\ell}]$ we have \begin{equation} A \otimes_{\mathbb{F}_1}\mathbb{F}_p \ =\ (A \otimes_{\mathbb{F}_1} \mathbb{Z}) \otimes_{\mathbb{Z}} \mathbb{F}_p\ =\ \mathbb{Z}[A] \otimes_{\mathbb{Z}} \mathbb{F}_p, \end{equation} where $\mathbb{Z}[A]$ is the monoidal ring defined by $A$. \subsection{Hidden Points Principle} The following trivial observation reveals the Hidden Points Principle at the topological level. \begin{proposition}[\cite{KT-Chap-Mot}] \label{topell1} Topologically, the structure of $\mathbf{PG}(n,\mathbb{F}_{1^\ell})$ is independent of the choice of $\ell$. \end{proposition} {\em Proof}.\quad It suffices to observe that for any $\gamma \in \mu_\ell$ and any $i \in \{0,1,\ldots,n\}$, we have \begin{equation} (X_i) = (\gamma X_i), \end{equation} whence the topology is the same as (= homeomorphic to) that of $\mathbf{PG}(n,\mathbb{F}_1)$. \hspace*{\fill}$\blacksquare$ \\ Due to this fact, we need to introduce ``hidden points,'' which cannot be detected by the classical way in the Zariski topology. We can formulate the Hidden Points Principle as follows (for a Deitmar scheme $\mathcal{X}$ defined over $\mathbb{F}_{1^\ell}$): \begin{equation} \vert \texttt{Hom}(\texttt{Spec}(\mathbb{F}_{1^\ell}),\mathcal{X}) \vert\ \gg \ \vert \mathcal{X}_{\mathbb{F}_{1^\ell}} \vert, \end{equation} where on the right hand side, $\mathcal{X}_{\mathbb{F}_{1^\ell}}$ is the set of $\mathbb{F}_{1^\ell}$-rational points defined as closed points (i.e., coming from (next-to-) maximal ideals). In fact, the ratio \begin{equation} f(\mathcal{X},\ell) := \frac{\vert \texttt{Hom}(\texttt{Spec}(\mathbb{F}_{1^\ell}),\mathcal{X})\vert}{\vert \mathcal{X}_{\mathbb{F}_{1^\ell}}\vert} \end{equation} is strictly increasing for a fixed $\mathcal{X}$, as a function of $\ell \in \mathbb{N} \setminus \{0\}$. \\ In this note, we will look for a minimal enlargement of Deitmar schemes which reveals the hidden points in Deitmar schemes such as $\texttt{Spec}(\mathbb{F}_{1^\ell}[X_1,\ldots,X_m])$ and $\texttt{Proj}(\mathbb{F}_{1^\ell}[X_1,\ldots,X_m])$, and which resolves the issues of \S\S \ref{issue}. \medskip \subsection{The combinatorial picture | I} \subsubsection{} Let $\Gamma = (V,E)$ be a graph (undirected), with $V$ the vertex set and $E$ the edge set. A {\em cover} of $\Gamma$ is a graph $\widetilde{\Gamma} = (\widetilde{V},\widetilde{E})$ together with a surjective graph morphism \begin{equation} \gamma: \widetilde{\Gamma} \ \mapsto\ {\Gamma} \end{equation} such that locally $\gamma$ is a bijection (that is, for any vertex $v \in \widetilde{V}$, $\gamma$ induces a bijection between the edges incident with $v$ and the edges incident with $\gamma(v)$). If there is a positive integer $m$ such that any vertex of $\Gamma$ has a fiber of size $m$, then $(\widetilde{\Gamma},\gamma)$ is an {\em $m$-fold cover}. Now let $A = (V,E)$ and $B = (V',E')$ be two graphs. The {\em lexicographic product} $A[B]$ of $A$ and $B$, also denoted by $A \cdot B$, is the graph defined on the vertex set $V \times V'$, such that $(v,v') \sim (w,w')$ if either $v \sim w$, or ($v = w$ and $v' \sim w'$). We allow $B$ to be a directed. The lexicographic product $A \cdot B$ is in general not commutative, and, considered as the morphism defined by \begin{equation} \pi: A\ \mapsto\ B:\ (v,w) \ \mapsto\ v, \end{equation} not a cover of $A$. On the other hand, each vertex fiber has a constant size $\vert V' \vert$, and the morphism {\em is} surjective. Let $G$ be any group, and $S$ a (minimal) generating set of $G$. The {\em Cayley graph} $\Gamma(G,S)$ of $G$ with respect to $S$ is the directed graph with vertex set $G$, and directed edges of type $(v,sv)$, with $s \in S$. (With any fixed $s$ corresponds a color, but that is not important here.) \subsection*{Example} Let $G = \mu_n$ be any cyclic group, with $n \in \mathbb{N}^{\times} \cup \{ \vert \mathbb{N} \vert \}$, and let $S = \{ s\}$, with $s$ a generator. Then $\Gamma(\mu_n,S)$ is a directed cycle on $n$ vertices. \subsubsection{Combinatorial picture} One way of introducing a combinatorial model for $\mathbb{P}^m(\mathbb{F}_{1^\ell})$ could be to define it as $K_{m + 1}[\Gamma(\mu_{\ell}),\{s\}]$. This model essentially corresponds to the Kapranov-Smirnov model of vector spaces over extensions of $\mathbb{F}_1$. \includegraphics{Pic_arxiv} \subsubsection{Automorphism group} The automorphism group of $K_{m + 1}[\Gamma(\mu_{\ell}),\{s\}]$ is easily seen to be isomorphic to $\mu_{\ell} \wr \texttt{S}_{m + 1}$. The picture presented in this section is certainly richer than the topological model: it now possesses the right automorphism group, but the counting polynomial is not the predicted one. \medskip \section{Frames} \label{frame} Let $S$ be any monoid, written multiplicatively and with $0$, and $m \in \mathbb{N}^{\times}$. The {\em affine frame} $\mathcal{A}(m,S)$ is defined as the set $\{ (s_1,\ldots,s_m) \vert s_i \in S \}$. Let $(0,\ldots,0) =: \omega$. If $S \setminus \{0\}$ is also a group, the {\em projective frame} $\mathcal{P}(m - 1,S)$ is defined as $\mathcal{A}(m,S) \setminus \omega$ modulo proportionality (so an element of $\mathcal{P}(m - 1,S)$ is a class $(s_1:\cdots:s_m) := \{ \sigma(s_1,\ldots,s_m) \vert \sigma \in S, \sigma \ne 0 \}$, with $(s_1,\ldots,s_m) \ne \omega$). If $m = 0$, $\mathcal{P}(m - 1,S) = \emptyset$; if $m = 1$, $\mathcal{P}(m - 1,S)$ is a point, and \begin{equation} \mathcal{P}(1,S) = \{(1 : 0),(0 : 1)\} \cup \{(1 : s) \vert s \in S^{\times}\}. \end{equation} \subsection{Frames over $\mathbb{F}_{1^\ell}$} Let $\ell \in \mathbb{N}^{\times}$. The affine, respectively projective, frame {\em over} $\mathbb{F}_{1^\ell}$ of {\em dimension} $d$ is $\mathcal{A}(d,\mathbb{F}_{1^\ell})$, respectively $\mathcal{P}(d,\mathbb{F}_{1^\ell})$. (Here, $d \in \mathbb{N}$, respectively $d \in \mathbb{N} \cup \{-1\}$.) We will see $\mathcal{A}(d,\mathbb{F}_{1^\ell})$ and $\mathcal{P}(d,\mathbb{F}_{1^{\ell}})$ as models of respectively affine and projective spaces over $\mathbb{F}_{1^\ell}$. \subsection{Subframes} An (affine) {\em subframe} of $\mathcal{A}(d,\mathbb{F}_{1^{\ell}})$ is any subset of tuples $(x_1,\ldots,x_d)$ in which a fixed number $e$ of coordinates ($0 \leq e \leq d$), are fixed elements of $\mathbb{F}_{1^\ell}$, and the others take all possible values. Its {\em dimension} is $d - e$. Similarly as in the previous subsection, one defines projective subframes of dimension $d - e - 1$ of $\mathcal{P}(d - 1,\mathbb{F}_{1^\ell})$ from affine subframes of dimension $d - e$ of $\mathcal{A}(d,\mathbb{F}_{1^\ell})$ endowed with the proportionality relation. \subsection{The combinatorial picture | II. From spaces to frames} Let $\mathbb{A}^m_{1^\ell} =\\ \texttt{Spec}(\mathbb{F}_{1^\ell}[X_1,\ldots,X_m])$ be the Deitmar scheme of $m$-dimensional affine space over $\mathbb{F}_{1^\ell}$. According to the point of view of morphisms, a (rational) point should correspond to a morphism \begin{equation} \gamma: \mathbb{F}_{1^\ell}[X_1,\ldots,X_m]\ \mapsto\ \mathbb{F}_{1^\ell}. \end{equation} Obviously, any such $\gamma$ is completely determined by $(\gamma(X_1),\ldots,\gamma(X_m))$, and conversely, any such choice gives us a morphism. So $\texttt{Hom}(\texttt{Spec}(\mathbb{F}_{1^\ell}),\mathbb{A}_1^m)$ is in a natural bijective correspondence with the affine frame $\mathcal{A}(m,\mathbb{F}_{1^\ell})$, and \begin{equation} \vert \texttt{Hom}(\texttt{Spec}(\mathbb{F}_{1^\ell}),\mathbb{A}_1^m)\vert = (\ell + 1)^m. \end{equation} Similar remarks hold for projective space Deitmar schemes and projective frames. In the latter case, we have \begin{equation} \vert \texttt{Hom}(\texttt{Spec}(\mathbb{F}_{1^\ell}),\mathbb{P}_1^m)\vert = \sum_{i = 0}^m(\ell + 1)^i. \end{equation} \subsubsection{Example (case $\ell = 1$)} The number of rational points of $\mathcal{P}(d,\mathbb{F}_1)$ is $2^{d} + 2^{d - 1} + \cdots + 1$. Amongst these are the $d + 1$ closed points of the underlying Deitmar scheme. \subsection{Automorphism group} Obviously, since we do not have any addition in $\mathbb{F}_{1^\ell}$, we have that the automorphism group (defined in a natural way) of $\mathcal{A}(d,\mathbb{F}_{1^\ell})$, respectively $\mathcal{P}(d,\mathbb{F}_{1^\ell})$, is \begin{equation} \mu_\ell \wr \texttt{S}_d, \ \mathrm{respectively}\ \Big(\mu_\ell \wr \texttt{S}_{d + 1}\Big) \Big/ \mathbb{F}_{1^\ell}^\times. \end{equation} This picture is very interesting: in case of affine and projective spaces over real fields $k$, the respective automorphism groups act transitively on the $k$-linear subspaces of fixed dimension, and in particular on the $k$-points. This is not at all the case here; observe the following. (Denote the affine group by $\mathbf{AGL}_d(1^\ell)$ and the projective group by $\mathbf{PGL}_{d + 1}(1^\ell)$.) \begin{observation} Let $\mathcal{A}^r(d,\mathbb{F}_{1^\ell})$ be the set of elements in $\mathcal{A}(d,\mathbb{F}_{1^\ell})$ which have precisely $r$ nonzero entries ($r \in \{ 0,1,\ldots,d\}$). Similarly we define $\mathcal{P}^r(d,\mathbb{F}_{1^\ell})$ ($r \in \{1,\ldots,d + 1\}$). Then $\mathbf{AGL}_d(1^\ell)$ stabilizes each set $\mathcal{A}^r(d,\mathbb{F}_{1^\ell})$, and acts transitively on its elements. Also, $\mathbf{PGL}_{d + 1}(1^\ell)$ stabilizes each set $\mathcal{P}^r(d,\mathbb{F}_{1^\ell})$, and acts transitively on its elements. \hspace*{\fill}$\blacksquare$ \end{observation} So in this setting, the original closed Deitmar points form one orbit under the full automorphism group. \medskip \begin{remark}{\rm If $\mu_\ell \cong \mathbb{F}_q^\times$ for some prime power $q$, then note that the sets $\mathcal{A}(d,\mathbb{F}_{1^\ell})$ and $\mathcal{P}(d,\mathbb{F}_{1^\ell})$ coincide with the rational point sets of $\mathbb{A}^d_{\mathbb{F}_q}$ and $\mathbb{P}^d_{\mathbb{F}_q}$. } \end{remark} \medskip \section{Frames and enlarged Deitmar schemes} As we have seen, the data $(K_{m}[\Gamma(\mu_{\ell},S)],\mathcal{P}(m - 1,\mathbb{F}_{1^\ell}))$ reveals a number of hidden properties of (say) projective spaces $\texttt{Proj}(\mathbb{F}_{1^\ell}[X_1,\ldots,X_m])$. As the points of an $\mathbb{F}_{1^\ell}$-frame naturally correspond to $\mathbb{F}_{1^\ell}$-rational points of the corresponding space, we want to enlarge Deitmar schemes such that points of the frames associated to affine/projective/etc. spaces appear as closed points. Those can be seen if we (minimally) enlarge the category of Deitmar schemes in the following way, by using congruences. \subsection{Congruences and congruence ideals} Let $M$ be a multiplicative commutative monoid. A {\em congruence} $\mathcal{C} \subseteq M \times M$ on $M$ is an equivalence relation with the additional property that if $(u,v) \in \mathcal{C}$, then $(mu,mv) \in \mathcal{C}$ for each $m \in M$. If $(u,v) \in \mathcal{C}$, we also write $u \sim v$. If $\mathcal{C}$ is a congruence on $M$, one shows that $M/\mathcal{C}$ (or $M/\sim$) naturally inherits the structure of a monoid. Note that in general the union of two congruences $\mathcal{C}, \mathcal{C}'$ is not a congruence (as transitivity does not necessary hold anymore). (The congruence generated by $\mathcal{C}, \mathcal{C}'$ is essentially the transitive closure.) The intersection {\em is} a congruence, though. Any ideal $I$ in the monoid $M$ gives rise to a congruence, called ``Reese congruence" and denoted by $\mathcal{C}_I$; if $m \in M\setminus I$, the class $[m]$ is just $\{m\}$; if $a, b \in I$, then $[a] = [b]$. So $I$ defines one class, which we will also denote by $[I]$. If $M$ has an absorbing element $0$, then $0 \in I$, so in that case $a \sim 0 \sim b$. From now on we keep assuming that $M$ has a $0$. A congruence $\mathcal{C}$ is {\em prime} if from $(ab,ac) \in \mathcal{C}$ follows that either $(b,c) \in \mathcal{C}$ or $(a,0) \in \mathcal{C}$. (Alternatively, a congruence $\mathcal{C}$ on $M$ is {\em prime} if $M/\mathcal{C}$ is integral, i.e., if it satisfies the cancellation property.) Now let $\frak{p}$ be a prime ideal in the monoid $M$, so that $0 \in \frak{p}$. Let $a, b, c$ be elements in $M$, and suppose $ab = ac$ in $M/\mathcal{C}_{\frak{p}}$. If $ab \sim ac \in \frak{p}$, then either $a \in \frak{p}$, or $a \not\in \frak{p}$ and $b, c \in \frak{p}$. (If $ab = ac \not\in \frak{p}$, then $a, b, c \ne 0$.) If $M$ is supposed to be integral, then $\mathcal{C}_\frak{p}$ is a prime ideal of $M$. \subsection{Example: affine space} Consider $\texttt{Spec}(\mathbb{F}_{1^\ell}[X_1,\ldots,X_m])$, Deitmar version. Put $A := \mathbb{F}_{1^\ell}[X_1,\ldots,X_m]$. If $\frak{p}$ is a prime ideal, then $\mathcal{C}_{\frak{p}}$ is a prime congruence since $A$ is integral. \subsection{New Zariski topology $\texttt{Spec}^c(M)$} In Deitmar scheme theory, any set of elements $S$ in $\mathbb{F}_1[X_1,\ldots,X_m]$ defines an ideal $I(S)$, and the closed set corresponding to $I(S)$ classically contains the solutions of all the equations defined by its elements. Only equations of the form $P = 0$ arise, with $P$ an element in $\mathbb{F}_1[X_1,\ldots,X_m]$. We need to allow equations of the form $P = 1$, and more generally, over $\mathbb{F}_{1^\ell}$, of the form $P = \mu$. Let $\texttt{Spec}^c(M)$ be the set of prime congruences on $M$. Then the Zariski topology it comes with (and denoted in the same way) is defined by the closed sets \begin{equation} C(I) := \{ \frak{p}\ \vert\ I \subseteq \frak{p}, \frak{p}\ \text{prime congruence} \}, \end{equation} with $I$ any congruence on $M$. The key is that the closed set topology is generated by the sets \begin{equation} \mathcal{D}(a,b) := \{ \frak{p}\ \vert\ (a,b) \in \frak{p}, \frak{p}\ \text{prime congruence} \}. \end{equation} \subsection{Localization} Let $\frak{p}$ be any prime congruence. Define $S_{\frak{p}}$ as the monoid multiplicatively generated by the set \begin{equation} \{ b - a\ \vert\ (a,b) \in A \times A,\ a \not\sim_{\frak{p}} b \}. \end{equation} Then $A_{\frak{p}} := S_{\frak{p}}^{-1}A$ is defined as the monoid naturally generated by the fractions $\{ \frac{u}{v}\ \vert\ u \in A, v \in S_{\frak{p}} \}$. Here, we identify $u/v$ with $u'/v'$ if there is a $w \in S_{\frak{p}}$ such that $wuv' = wu'v$. \subsection{Sections and structure sheaf} Let $U$ be an open set. Then $\mathcal{O}_X(U)$ is defined as the set of maps \begin{equation} s : U \ \mapsto\ \coprod_{\frak{p} \in U}A_{\frak{p}} \end{equation} such that for each prime $\frak{p} \in U$, $s(\frak{p}) \in A_{\frak{p}}$, and for which there is a neighborhood $V$ of $\frak{p}$ in $U$ such that for all $\frak{q} \in V$, $s(\frak{q})$ is locally a fraction. This means that there are $a \in A$ and $f \in \bigcap_{\frak{q} \in V} S_{\frak{q}}$ so that $s = a/f$ over $V$. Note that $\mathcal{O}_X(U)$ carries the structure of a monoid. \subsection{Example: maximal ideals (closed points) for affine spaces} Consider again $A := \mathbb{F}_{1^\ell}[X_1,\ldots,X_m]$. For each coordinate monomial $P$ and each $\mu \in \mathbb{F}_{1^\ell}$, we consider the {\em base congrence} on $A$ generated by $P \sim \mu$. If, for example, $P = X_j$, the congruence is \begin{equation} \{ (QX_j^n,Q\mu^n) \vert Q \in A, n \in \mathbb{N}\} \cup \{ (Q\mu^n,QX_j^n) \vert Q \in A, n \in \mathbb{N} \}. \end{equation} It is easy to see that the maximal congruences are precisely those generated by base congruences of the following form \begin{equation} \{ (X_i,\mu_i)\ \vert\ i \in \{1,\ldots,n \}, \mu_i \in \mathbb{F}_{1^\ell} \}. \end{equation} In other words, the maximal congruences correspond to the rational points/frame points of $\mathcal{A}(m,\mathbb{F}_{1^\ell})$ as described in \S \ref{frame}. \subsection{General congruence schemes and $\texttt{Proj}^c$-schemes} For the theory of general congruence schemes and more details, we refer to Deitmar's paper \cite{congruence}. As for the $\texttt{Proj}^c$-construction, this is tersely described in \cite{MMKT2} in the case of projective spaces. One defines the {\em irreducible congruence} $\texttt{Irr}^c$ on $A = \mathbb{F}_{1^\ell}[X_1,\ldots,X_m]$ as \begin{equation} \texttt{Irr}^c := \Big\langle X_1 \sim 0, \ldots, X_m \sim 0 \Big\rangle, \end{equation} leaves it out, and applies a construction which is similar to that in Deitmar scheme theory. The next-to-maximal congruences in $A$ correspond to the closed points, and they correspond to the elements of the projective frame $\mathcal{P}(m - 1,\mathbb{F}_{1^\ell})$. \medskip \subsection{Maximal points vs. closed points} In the congruence setting of this section, one observes that for affine and projective spaces the set of rational points {\em coincides} with the set of closed points, a phenomenon which for proper fields only occurs for algebraically closed fields. \medskip \subsection{Absolute and geometric $\mathbb{F}_1$-Frobenius endomorphisms} Consider the algebraic closure $\overline{\mathbb{F}_1}$ of $\mathbb{F}_1$; by definition, it consists of all complex roots of unity, together with an absorbing element $0$, foreseen with multiplication. We define the {\em $\mathbb{F}_1$-Frobenius endomorphism}\index{absolute!Frobenius endomorphism} of degree $n \in \mathbb{N}$, denoted $\mathrm{Fr}^n_1$\index{$\mathrm{Fr}^n_1$}, to be the map \begin{equation} \mathrm{Fr}_1^n: \overline{\mathbb{F}_1} \longrightarrow \overline{\mathbb{F}_1}: x \longrightarrow x^n. \end{equation} Elements of $\mathbb{F}_{1^{d}} \cong \mu_d \cup \{0\} \leq \overline{\mathbb{F}_1}$ are characterized by the fact that they are the solutions of \begin{equation} {\mathrm{Fr}_1^{d + 1}}(x) = x, \end{equation} which is analogous to the fact that elements of finite fields $\mathbb{F}_{q^d} \leq \overline{\mathbb{F}_q}$ are singled out as fixed points of $\mathrm{Fr}^d$. The absolute $\mathbb{F}_1$-Frobenius map $\mathrm{Fr}_1^{d + 1}$ with $d \in \mathbb{N}^\times$ acts on an extended Deitmar scheme $\mathcal{Y}$ of finite type over $\mathbb{F}_{1^d}$ by acting trivially on the topology, and as $\mathrm{Fr}_1^{d + 1}$ on the structure scheaf. Passing to the scheme $\overline{\mathcal{Y}} := \mathcal{Y} \otimes_{\texttt{Spec}(\mathbb{F}_{1^d})}\texttt{Spec}(\overline{\mathbb{F}_1})$, we define the {\em geometric $\mathbb{F}_1$-Frobenius map} (still in the same degree) as the map $\mathrm{Fr}_{\mathrm{geom}}^{d + 1}: \overline{\mathcal{Y}} \mapsto \overline{\mathcal{Y}}$ with local comorphisms $y \otimes f \mapsto y^{d + 1} \otimes f$. It acts as \begin{equation} \mathbf{x} \mapsto \mathbf{x}^{d + 1} \end{equation} on frame points. \medskip \section{Zeta functions} In the inspiring note \cite{Kurozeta}, Kurokawa says that a $\mathbb{Z}$-scheme $\mathcal{X}$ is of {\em $\mathbb{F}_1$-type} if there exists a polynomial $N_\mathcal{X}$ in $\mathbb{Z}[X]$ such that, for each prime power $q$, we have \begin{equation} \vert \mathcal{X}_q \vert = N_{\mathcal{X}}(q). \end{equation} (Through one of Tate's conjectures, such schemes would come with a mixed Tate motive, a feature which one sees more clearly in the appropriate Grothendieck ring | see for instance \cite{KT-Chap-Mot} for a detailed discussion.) In fact, originally, in \cite{Kurozeta}, a $\mathbb{Z}$-scheme (of finite type) is of $\mathbb{F}_1$-type if its arithmetic zeta function $\zeta_{\mathcal{X}}(s)$ can be expressed in the form \begin{equation} \zeta_{\mathcal{X}}(s) = \prod_{k = 0}^n\zeta(s - k)^{a_k} \end{equation} with the $a_k$s in $\mathbb{Z}$. (Here, $s$ is a complex variable | cf. the remark right after the statement of Theorem \ref{Kuro}.) The definition through the counting polynomial then is derived from the following result. (In the next theorem, $\zeta(\cdot)$ is the classical Riemann zeta.) \begin{theorem}[Kurokawa \cite{Kurozeta}] \label{Kuro} Let $\mathcal{X}$ be a $\mathbb{Z}$-scheme of finite type. The following are equivalent. \begin{itemize} \item[{\rm (i)}] The arithmetic zeta function of $\mathcal{X}$ has the following form: \begin{equation} \zeta_{\mathcal{X}}(s) = \prod_{k = 0}^n\zeta(s - k)^{a_k} \end{equation} with the $a_k$s in $\mathbb{Z}$. \item[{\rm (ii)}] For all primes $p$ we have \begin{equation} \zeta_{\mathcal{X}\vert \mathbb{F}_p}(s) = \prod_{k = 0}^n(1 - p^{k - s})^{-a_k} \end{equation} with the $a_k$s in $\mathbb{Z}$. \item[{\rm (iii)}] There exists a polynomial $N_{\mathcal{X}}(Y) = \sum_{k = 0}^na_kY^k$ such that \begin{equation} \vert \mathcal{X}_{\mathbb{F}_{p^m}} \vert = N_{\mathcal{X}}(p^m) \end{equation} for all finite fields $\mathbb{F}_{p^m}$. \end{itemize} \end{theorem} For details about matters of convergence, we refer to Theorem \ref{eqext} below. Kurokawa defines the {\em $\mathbb{F}_1$-zeta function}\index{$\mathbb{F}_1$-zeta function} of a $\mathbb{Z}$-scheme $\mathcal{X}$ of $\mathbb{F}_1$-type as \begin{equation} \zeta_{\mathcal{X}\vert \mathbb{F}_1}(s) := \prod_{k = 0}^n(s - k)^{-a_k} \end{equation} with the $a_k$s as above. The {\em Euler characteristic}\index{Euler characteristic} is \begin{equation} \vert X_{\mathbb{F}_1} \vert := \sum_{k = 0}^na_k = N_{\mathcal{X}}(1). \end{equation} To fit this formula into {\bf Count}, put \begin{equation} \mathbb{F}_{1^0} = \mathbb{F}_{1^1}. \end{equation} (Although this seems rather artificial, it is also rather interesting that one makes {\bf Count} work by putting $0 = 1$!) \begin{theorem}[Kurokawa \cite{Kurozeta}] \label{Kurolimit} Let $\mathcal{X}$ be a $\mathbb{Z}$-scheme of $\mathbb{F}_1$-type. Then \begin{equation} \zeta_{\mathcal{X}\vert \mathbb{F}_1}(s)\index{$\zeta_{X\vert \mathbb{F}_1}(s)$} = \lim_{p \longrightarrow 1}\zeta_{\mathcal{X}\vert \mathbb{F}_p}(s)(p - 1)^{\vert \mathcal{X}_{\mathbb{F}_1} \vert}. \end{equation} Here, $p$ is seen as a complex variable (so that the left hand term is the leading coefficient of the Laurent expansion of $\zeta_{\mathcal{X} \vert \mathbb{F}_1}(s)$ around $p = 1$). \end{theorem} \medskip \subsection{Schemes of $\mathbb{F}_1$-type vs. schemes defined over $\mathbb{F}_1$} \label{VS} As Deitmar envisioned in \cite{Deitmarschemes2}, commutative monoids with a $1$ and $0 \ne 1$ should/could be seen as ``$\mathbb{F}_1$-algebras.'' For each such monoid $A$, we have an embedding $\mathbb{F}_1 \hookrightarrow A$, so a projection $\texttt{Spec}(A) \twoheadrightarrow \texttt{Spec}(\mathbb{F}_1)$. So $\texttt{Spec}(A)$ is ``defined over $\mathbb{F}_1$.'' (And one can make similar observations for general monoidal schemes.) Considering a $k$-algebra $\mathcal{A}$ for a field $k$, one can realize $\mathcal{A}$ as a quotient of a polynomial ring over $k$ with the kernel of an appropriate surjective morphism to $\mathcal{A}$. In particular, if \begin{equation} \iota: k\ \hookrightarrow\ R \end{equation} is an embedding of a field in a unital commutative ring, then $R$ comes with a $k$-algebra structure, so $R$ is a quotient of a polynomial ring over $k$. And any unital commutative ring is a $\mathbb{Z}$-algebra, of course. Over $\mathbb{F}_1$, that is to say, in the setting of monoidal schemes, this situation is less natural, as some unital commutative monoids with $0$ can only be realized as a quotient of a polynomial ring over $\mathbb{F}_1$ and a {\em congruence} (instead of a monoidal ideal). For instance, although we have an embedding $\mathbb{F}_1 \hookrightarrow \mathbb{Z}$, the latter, as a monoid, {\em needs} a congruence which is {\em not} associated to a monoidal ideal if one wants to see it as such a quotient. Whereas over $\mathbb{F}_{1^2}$, $\mathbb{Z}$ is indeed even {\em isomorphic} to a polynomial ring. So it seems more natural to say that $\texttt{Spec}(\mathbb{Z})$ is ``defined over $\mathbb{F}_{1^2}$,'' instead than over $\mathbb{F}_1$. But still, strictly speaking, if $\texttt{Spec}(A)$ is defined over $\mathbb{F}_{1^\ell}$, it should also be defined over $\mathbb{F}_1$ as well. With this discussion in mind, one can then decide when a $\mathbb{Z}$-scheme is defined over $\mathbb{F}_1$, or an extension. One should make a distinction in any case with Kurokawa's $\mathbb{Z}$-schemes ``of $\mathbb{F}_1$-type.'' If $\mathcal{X}$ is a $\mathbb{Z}$-scheme of $\mathbb{F}_1$-type \`{a} la Kurokawa, then any reasonable definition in the same vein of $\mathbb{Z}$-schemes over extensions of $\mathbb{F}_1$ will lead to the property that ``of $\mathbb{F}_{1^m}$-type'' implies ``of $\mathbb{F}_{1^n}$-type'' if $m$ divides $n$, but not necessarily the other way around. (This phenomenon will manifest itself in the next subsection.) So the situation is quite different (and more general) than the $\mathbb{F}_1$-algebra viewpoint. \medskip \subsection{Dirichlet series} A {\em Dirichlet series} is a series of the form \begin{equation} D(s) = \sum_{m = 1}^\infty\gamma(m)m^{-s} \end{equation} with $s$ a complex variable and $\gamma: \mathbb{N} \to \mathbb{C}$ a function. It can be shown that if $\sum_{m = 1}^\infty\vert \gamma(s)m^{-s}\vert$ does not diverge for all $s$, there exists a real number $\kappa$ such that $\sum_{m = 1}^\infty\vert \gamma(s)m^{-s}\vert$ converges for all $s$ with $\Re(s) > \kappa$ (that is, the Dirichlet series is absolutely convergent for these values). (The real part of a complex number $c$ is denoted by $\Re(c)$.) We will use the next well known result later on. \begin{theorem}[Uniqueness of coefficients] Suppose \begin{align} D(s) = \sum_{m = 1}^\infty\gamma(m)m^{-s} \ \text{and}\ E(s) = \sum_{m = 1}^\infty\beta(m)m^{-s} \end{align} are Dirichlet series, both absolutely convergent for $\Re(s) > \sigma$. If $D(s) = E(s)$ for each $s$ in some infinite sequence $\{s_k\}_k$ such that $\Re(s_k) \to \infty$ as $k \to \infty$, then $\gamma(n) = \beta(n)$ for each $n \in \mathbb{N}$. \end{theorem} \medskip \subsection{Schemes of $\mathbb{F}_{1^\ell}$-type} Let $V$ be nonsingular projective variety over $\mathbb{F}_q$, the finite field with $q$ elements. The {\em local zeta function} (or {\em congruence zeta function}) of $V$ is given by \begin{equation} \zeta_V(s) = Z(V,s)\ := \ \exp\Big(\sum_{m = 1}^\infty \frac{\vert V_{q^m} \vert}{m}q^{-sm}\Big). \end{equation} Sometimes $Z(V,s)$ is also denoted by $Z(V,q^{-s})$. The variable transformation $u = q^{-s}$ gives $Z(V,u)$ as a formal power series in $u$. To stress that the local zeta function is expressed ``with respect to $\mathbb{F}_q$,'' we also write $\zeta_{V \vert \mathbb{F}_q}(s)$ instead of $\zeta_{V}(s)$. Let $\ell$ be a positive nonzero integer. We will say that a $\mathbb{Z}$-scheme $\mathcal{X}$ is {\em of $\mathbb{F}_{1^\ell}$-type} if there exists a polynomial $N_{\mathcal{X}}(Y) = \sum_{k = 0}^na_kY^k$ such that for all primes $p$ and all extensions $\mathbb{F}_{p^{\ell m}}$ of $\mathbb{F}_{p^\ell}$, we have \begin{equation} \vert \mathcal{X}_{\mathbb{F}_{p^{\ell m}}} \vert = N_{\mathcal{X}}(p^{\ell m}). \end{equation} \begin{proposition} If the $\mathbb{Z}$-scheme $\mathcal{X}$ is of $\mathbb{F}_{1^m}$-type, $m \in \mathbb{N}_0$, then it is also of $\mathbb{F}_{1^r}$-type for any positive integer multiple $r$ of $m$. \hspace*{\fill}$\blacksquare$ \end{proposition} In the next theorem, for a complex number $c$, the modulus will be denoted by $\vert c\vert$. Also, $\log(\cdot)$ will denote the principal value complex logarithm. \begin{theorem} \label{eqext} Let $\mathcal{X}$ be a $\mathbb{Z}$-scheme of finite type, and $\ell$ a positive integer. The following are equivalent. \begin{itemize} \item[{\rm (i)}] There exist $k \in \mathbb{N}$ and integers $a_1,\ldots,a_k$, so that the following equality holds: \begin{equation} \prod_{\text{$p$ prime}}\zeta_{\mathcal{X}\vert\mathbb{F}_{p^\ell}}(s) = \prod_{k = 0}^n\zeta(\ell(s - k))^{a_k}, \end{equation} where $s \in \mathbb{C}$ and $\Re(s) > k + \frac{1}{\ell}$. \item[{\rm (ii)}] There exist $k \in \mathbb{N}$ and integers $a_1,\ldots,a_k$, so that the following equality holds for all primes $p$: \begin{equation} \zeta_{\mathcal{X}\vert \mathbb{F}_{p^\ell}}(s) = \prod_{k = 0}^n(1 - p^{\ell(k - s)})^{-a_k}, \end{equation} where $s \in \mathbb{C}$ and $\Re(s) > k + \frac{1}{\ell}$. \item[{\rm (iii)}] There exists a polynomial $N_{\mathcal{X}}(Y) = \sum_{k = 0}^na_kY^k \in \mathbb{Z}[Y]$ such that \begin{equation} \vert \mathcal{X}_{\mathbb{F}_{p^{\ell m}}} \vert = N_{\mathcal{X}}(p^{\ell m}) \end{equation} for all finite fields $\mathbb{F}_{p^{\ell m}}$. \end{itemize} \end{theorem} {\em Proof}.\quad We first show that (ii) and (iii) are equivalent. Observe that, with $q = p^\ell$, \begin{equation} \label{first} \log(\zeta_{\mathcal{X}\vert \mathbb{F}_{q}}(s)) = \ \sum_{m = 1}^\infty \frac{\vert \mathcal{X}_{q^m} \vert}{m}q^{-sm} \end{equation} while \begin{align} \label{second} \log\Big(\prod_{k = 0}^n(1 - p^{\ell(k - s)})^{-a_k}\Big) &= \sum_{k = 0}^n-a_k\log(1 - p^{\ell(k - s)}) \nonumber \\ &= \sum_{k = 0}^na_k\Big(\sum_{m = 1}^{\infty}\frac{p^{\ell(k - s)m}}{m} \Big) \nonumber \\ &= \sum_{m = 1}^\infty \frac{\Big(\sum_{k = 0}^na_k{p^{\ell mk}}\Big)}{m}q^{-sm}. \end{align} (Note that we can indeed expand $\log(1 - p^{\ell(k - s)})$ in its Mercator series since $\vert p^{\ell(k - s)} \vert < 1$.) Then assuming (ii) (equality of the expressions in (\ref{first}) and (\ref{second})), equality of the coefficients in both sides leads to (iii). Also, after assuming (iii), substituting $\vert \mathcal{X}_{\mathbb{F}_{p^{\ell m}}} \vert = N_{\mathcal{X}}(p^{\ell m})$ for all $p^{\ell m}$ in (\ref{first}) leads to (ii) (using (\ref{second})). Now assume (ii). Then \begin{align} \prod_{\text{$p$ prime}}\zeta_{\mathcal{X}\vert\mathbb{F}_{p^\ell}} &= \prod_{\text{$p$ prime}}\Big(\prod_{k = 0}^n(1 - p^{\ell(k - s)})^{-a_k}\Big) \nonumber \\ &= \prod_{k = 0}^n\Big(\prod_{\text{$p$ prime}}(1 - p^{\ell(k - s)})^{-a_k}\Big) \nonumber \\ &= \prod_{k = 0}^n\zeta(\ell(s - k))^{a_k}, \end{align} which yields (i).\\ Then assume (i). We obtain \begin{align} \prod_{\text{$p$ prime}}\zeta_{\mathcal{X}\vert \mathbb{F}_p^\ell} &= \prod_{k = 0}^n\zeta(\ell(s - k))^{a_k} \nonumber \\ &= \prod_{k = 0}^n\Big(\prod_{\text{$p$ prime}}(1 - p^{\ell(k - s)})^{-a_k}\Big) \nonumber \\ &= \prod_{\text{$p$ prime}}\Big(\prod_{k = 0}^n(1 - p^{\ell(k - s)})^{-a_k}\Big). \end{align} Taking complex logarithms of both sides yields: \begin{align} \sum_{\text{$p$ prime}}\Big(\sum_{m = 1}^\infty \frac{\vert \mathcal{X}_{q^m} \vert}{m}q^{-sm}\Big) &= \sum_{\text{$p$ prime}}\Big(\sum_{k = 0}^n-a_k\log(1 - p^{\ell(k - s)})\Big) \nonumber \\ &= \sum_{\text{$p$ prime}}\bigg(\sum_{k = 0}^na_k\Big(\sum_{m = 1}^{\infty}\frac{p^{\ell(k - s)m}}{m} \Big)\bigg). \end{align} It follows that \begin{align} \sum_{p^m\ \text{prime power}}\frac{\vert \mathcal{X}_{p^{\ell m}} \vert}{m}p^{-\ell ms} &= \sum_{p^m\ \text{prime power}}\Big(\sum_{k = 1}^n a_kp^{\ell mk} \Big)p^{-\ell ms}. \end{align} The uniqueness of the coefficients in the Dirichlet series lead to (iii). \hspace*{\fill}$\blacksquare$ \\ \begin{remark}{\rm The reader observes that the switchings above of summation symbols resp. product symbols can indeed be done as $\Re(s) > k + \frac{1}{\ell}$. (For the summation symbols, one can use Fubini's Theorem; switching of the product symbols reduces to summation through application of $\log(\cdot)$.) }\end{remark} \begin{remark}{\rm The author is not aware of a connection between the property expressed in (iii) of the previous theorem and (variations of) mixed Tate motives (as is the case for $\ell = 1$). } \end{remark} \begin{remark}{\rm Note that the equivalence between (ii) and (iii) remains true if stated for particular primes, and hence particular sets of primes. We will come back to this property further on. }\end{remark} Define the {\em $\mathbb{F}_{1^\ell}$-zeta function} of a $\mathbb{Z}$-scheme $\mathcal{X}$ of $\mathbb{F}_{1^\ell}$-type as \begin{equation} \zeta_{\mathcal{X}\vert \mathbb{F}_{1^\ell}}(s) := \prod_{k = 0}^n(\ell(s - k))^{-a_k} \end{equation} with the $a_k$s as in the previous theorem. \medskip Similarly as Theorem \ref{Kurolimit}, we have: \begin{theorem} Let $\mathcal{X}$ be a $\mathbb{Z}$-scheme of $\mathbb{F}_{1^\ell}$-type. Then \begin{equation} \zeta_{\mathcal{X}\vert \mathbb{F}_{1^\ell}}(s)\index{$\zeta_{X\vert \mathbb{F}_1}(s)$} = \lim_{p \longrightarrow 1}\zeta_{\mathcal{X}\vert \mathbb{F}_{p^\ell}}(s)(p - 1)^{\vert \mathcal{X}_{\mathbb{F}_1} \vert}. \end{equation} Here, $p$ is again seen as a complex variable. \end{theorem} {\em Proof}.\quad We have \begin{equation} \zeta_{X\vert \mathbb{F}_{1^\ell}}(s)(p - 1)^{\# X(\mathbb{F}_1)} = \prod_{k = 0}^n\big(\frac{1 - p^{\ell(k - s)}}{p - 1}\big)^{-a_k} \end{equation} so that for the limit we get \begin{eqnarray} \lim_{p \longrightarrow 1}\zeta_{X\vert \mathbb{F}_{1^\ell}}(s)(p - 1)^{\# X(\mathbb{F}_1)} &= &\prod_{k = 0}^n\big(\lim_{p \longrightarrow 1}\frac{1 - p^{\ell(k - s)}}{p - 1}\big)^{-a_k}\nonumber \\ &= &\prod_{k = 0}^n(\ell(s - k))^{-a_k}\nonumber \\ &= &\zeta_{X\vert \mathbb{F}_{1^\ell}}(s). \end{eqnarray} \hspace*{\fill}$\blacksquare$ \medskip \section{Schemes of $\mathbb{F}_1$-type versus schemes of $\mathbb{F}_{1^\ell}$-type} In general, $\mathbb{Z}$-schemes of $\mathbb{F}_{1^\ell}$-type are not necessarily of $\mathbb{F}_{1^m}$-type, with $m$ a natural divisor of $\ell$ (cf. Remark \ref{VS}). In particular, they need not be of $\mathbb{F}_1$-type, as the following example shows. \subsection{} Let $X^2 + uX + v \in \mathbb{Z}[X,X_2,\ldots,X_m]$ and $u$ odd, $\mathrm{gcd}(u,v) = 1$. It is easy to see that over $\mathbb{F}_{p^2}$, $p$ any prime, the number of points is $2(p^2 + 1)^m$. Taking $(u,v)$ such that $X^2 + uX + v$ is not reducible over every prime field $\mathbb{F}_p$, we obtain a scheme of $\mathbb{F}_{1^2}$-type which is not of $\mathbb{F}_1$-type. Its $\mathbb{F}_{1^2}$-counting polynomial is \begin{equation} 2(Y + 1)^m = 2\sum_{k = 0}^m\begin{pmatrix}m \\ k \end{pmatrix}Y^{m - k}, \end{equation} so its $\mathbb{F}_{1^2}$-zeta function is \begin{equation} \prod_{k = 0}^m(2(s - k))^{-2\begin{pmatrix}m \\ k \end{pmatrix}}. \end{equation} \subsection{} Let $\mathcal{X}$ be a $\mathbb{Z}$-scheme of finite type defined over $\mathbb{F}_1$. By definition, it means that $\mathcal{X}$ is a $\mathbb{Z}$-scheme of finite type which arises from Deitmar base extension to $\mathbb{Z}$ from a Deitmar scheme. Then Deitmar showed in \cite{Deitmarzeta} that there exists a natural number $e$ and a polynomial $N(Y) \in \mathbb{Z}[Y]$ such that for every prime power $q$ for which $\mathrm{gcd}(q - 1,e) = 1$, one has $\#\mathcal{X}(\mathbb{F}_q) = N(q)$. It might be interesting to consider $\mathbb{Z}$-schemes $\mathcal{X}$ which come with a counting polynomial $N_{\mathcal{X}}(Y)$ such that for some positive integer $\ell$, we have that \begin{equation} \#\mathcal{X}(\mathbb{F}_{p^\ell}) = N_{\mathcal{X}}(p^\ell), \end{equation} where $p$ varies over the primes, except possibly a {\em finite number of exceptions}. (When $\ell = 1$, such schemes are related to mixed Tate motives.) \\ Constructing examples is easy. Consider, for example, the affine conics with equation \begin{equation} \mathcal{C}(u,v): X^2 - uY^2 = v, \end{equation} with $u, v \in \mathbb{Z}$ (and $u, v$ fixed). Over a finite field $\mathbb{F}_{2^n}$ we have that $\mathcal{C}(u,v)$ has $2^n + 1$ points. If $p$ is an odd prime, we have that $\vert \mathcal{C}(u,v)\vert_{\mathbb{F}_{p^n}}$ is $p^n - 1$ if $u$ is a square, and $p^n + 1$ if $u$ is not. If $u$ would be a square in $\mathbb{F}_p$, it is also a square in $\mathbb{F}_{p^2}$, so $\vert \mathcal{C}(u,v) \vert_{\mathbb{F}_p} = p - 1$ and $\vert \mathcal{C}(u,v) \vert_{\mathbb{F}_{p^2}} = p^2 - 1$. If $u$ is not a square in $\mathbb{F}_p$, we have $\vert \mathcal{C}(u,v)\vert_{\mathbb{F}_p} = p + 1$, but in the quadratic extension $\mathbb{F}_{p^2}$ we know that $u$ {\em is} a square, so that $\vert \mathcal{C}(u,v) \vert_{\mathbb{F}_{p^2}} = p^2 - 1$. In other words, we have a fixed counting polynomial for all fields $\mathbb{F}_{p^2}$ ($p$ any prime) in odd characteristic, but not necessarily over $\mathbb{F}_p$. \subsection{} The {\em Grothendieck ring of schemes} of finite type over a field $k$, denoted as $K_0(\texttt{Sch}_{k})$, is generated by the isomorphism classes of schemes $\mathcal{X}$ of finite type over $\mathbb{F}_1$, $[\mathcal{X}]_{k}$, with the relation \begin{equation} [\mathcal{X}]_{k}= [\mathcal{X}\setminus \mathcal{Y}]_{k} + [\mathcal{Y}]_{k} \end{equation} for any closed subscheme $\mathcal{Y}$ of $\mathcal{X}$ and with the product structure given by \begin{equation} [\mathcal{X}]_{k}\cdot[\mathcal{Y}]_{k}= [\mathcal{X}\times_{k}\mathcal{Y}]_{k}. \end{equation} If it is clear what $k$ is, we will not write the index. Denote by ${\mathbb{L}} = [\mathbb{A}^1_{k}]$ the class of the affine line over $k$. By the first defining property, one obtains typical decompositions such as \begin{equation} [\mathbb{P}^n_k]_k = \mathbb{L}^n + \cdots + \mathbb{L} + 1. \end{equation} Let $\texttt{Sch}_{\mathbb{F}_{q}}$ be the category of schemes of finite type over the finite field $\mathbb{F}_{q}$. An {\em additive invariant} (or ``multiplicative Euler-Poincar\'{e} characteristic'') from $\texttt{Sch}_{\mathbb{F}_{q}}$ to $\mathbb{Z}$ is a map $\alpha$ which satisfies the following properties: \begin{itemize} \item[(AI1)] if $X, Y$ are in $\texttt{Sch}_{\mathbb{F}_{q}}$ and $X \cong Y$, then $\alpha(X) = \alpha(Y)$; \item[(AI2)] if $X$ is in $\texttt{Sch}_{\mathbb{F}_{q}}$ and $C$ is a closed subscheme, then $\alpha(X) = \alpha(C) + \alpha(X \setminus C)$; \item[(AI3)] if $X, Y$ are in $\texttt{Sch}_{\mathbb{F}_{q}}$, then $\alpha(X \times Y) = \alpha(X)\cdot\alpha(Y)$. \end{itemize} An additive invariant $\alpha$ factors uniquely through a ring morphism from $K_0(\texttt{Sch}_{\mathbb{F}_{q}})$ to $\mathbb{Z}$ by letting \begin{equation} \alpha(\sum_i[X_i]) = \sum_i\alpha[X_i]. \end{equation} An element $\gamma$ in $K_0(\texttt{Sch}_{\mathbb{F}_q})$, $\mathbb{F}_q$ a finite field, is said to be {\em polynomial-count} (according to \cite[Appendix]{Katz}) if there exists a polynomial $P(Y) = \sum_ia_iY^i \in \mathbb{C}[Y]$ such that for every finite extension $\mathbb{F}_{q^n}\Big/\mathbb{F}_q$, we have that $\vert \gamma(\mathbb{F}_q) \vert_{\mathbb{F}_{q^n}} = P(q^n)$. By \cite[Appendix]{Katz}, it follows that then $P(Y) \in \mathbb{Z}[Y]$. Note that $\vert \cdot \vert_{\mathbb{F}_{q^n}}$ is well defined on $K_0(\texttt{Sch}_{\mathbb{F}_{q^n}})$ by extension from the function $\vert \cdot \vert_{\mathbb{F}_{q^n}}$ on $\texttt{Sch}_{\mathbb{F}_{q^n}}$, using the fact that the latter function is an additive invariant from $\texttt{Sch}_{\mathbb{F}_{q^n}}$ to $\mathbb{Z}$. So in terms of \cite[Appendix]{Katz}, by Theorem \ref{eqext} a $\mathbb{Z}$-scheme $\mathcal{X}$ is of $\mathbb{F}_{1^\ell}$-type precisely if $[\mathcal{X}_{\mathbb{F}_{p^\ell}}] \in K_0(\texttt{Sch}_{\mathbb{F}_{p^\ell}})$ is polynomial-count for all primes $p$, where the corresponding counting polynomials {\em are the same for all primes $p$}. Also in terms of \cite[Appendix]{Katz}, this means that $[\mathcal{X}_{p^\ell}]$ is {\em zeta equivalent} to an element $P_{\mathcal{X}}$ in $\mathbb{Z}[\mathbb{L}]$ for all primes $p$, where $P_{\mathcal{X}}$ is independent of the prime $p$. \medskip \section{Examples} For affine and projective spaces, we obtain the following zeta functions (over $\mathbb{F}_p^\ell$ and $\mathbb{F}_{1^\ell}$, with $\ell \in \mathbb{N}^\times$): \begin{equation} \left\{\begin{array}{ccc} \zeta_{\mathbb{A}^n\vert \mathbb{F}_{p^\ell}}(s) &= &\Big({1 - p^{\ell(n - s)}}\Big)^{-1};\\ & & \\ \zeta_{\mathbb{A}^n\vert \mathbb{F}_{1^\ell}}(s) &= &\Big({\ell(s - n)}\Big)^{-1}, \end{array}\right. \end{equation} and \begin{equation}\left\{ \begin{array}{ccc} \zeta_{\mathbb{P}^n\vert \mathbb{F}_{p^\ell}}(s) &= &\Big({(1 - p^{-\ell s})(1 - p^{\ell(1 - s)})\cdots(1 - p^{\ell(n - s)})}\Big)^{-1}; \\ & & \\ \zeta_{\mathbb{P}^n\vert \mathbb{F}_{1^\ell}}(s) &= &\Big({\ell^{n + 1}s(s - 1)\cdots(s - n)}\Big)^{-1}. \end{array}\right. \end{equation} \bigskip We provide some more examples. In \cite{MMKT,MMKT2}, the authors have defined and studied a functor $\mathcal{F}_k$ for each field finite $k$, including $\mathbb{F}_1$, which maps any ``loose graph'' $\Gamma$ (which is a more general version of a graph, where edges with $0$ or $1$ vertices are allowed) to a $k$-scheme $\mathcal{F}(\Gamma) \otimes_{\mathbb{F}_1}k$. Here, $\mathbb{F}_1$-schemes are Deitmar schemes with the additional minimal congruences considered in this paper. It is shown in \cite{MMKT} that any such scheme comes with a counting polynomial, which is in fact independent of the field. So these schemes are of $\mathbb{F}_1$-type, so also of $\mathbb{F}_{1^\ell}$-type for any positive integer $\ell$. For loose trees, precise calculations are made in \cite{MMKT} to obtain the counting polynomials. The outcome is as follows. Let $\Gamma$ be a loose tree. We use the following notation: \begin{itemize} \item $D$ is the set of degrees $\{d_1, \ldots, d_m \}$ of $V(\Gamma)$ such that $1 < d_1 < d_2 < \ldots < d_m$; \item $n_i$ is the number of vertices of $\Gamma$ with degree $d_i$, $1\leq i \leq m$; \item $\displaystyle I= \sum_{i=1}^m n_i - 1$; \item $E$ is the number of vertices of $\Gamma$ with degree $1$. \end{itemize} Then, \begin{equation} \big[\Gamma\big]_{_{\mathbb{F}_1}} = \displaystyle\sum_{i = 1}^m n_i\underline{\mathbb{L}}^{d_i} - I\cdot\underline{\mathbb{L}} + I + E. \end{equation} This formula is expressed in terms of the Grothendieck ring of $\mathbb{F}_1$-schemes of finite type (see \cite{MMKT}). The symbol $\underline{\mathbb{L}}$ denotes the class of the affine line over $\mathbb{F}_1$. (The same polynomial arises after base extension to fields.) For {\em each} positive integer $\ell$, the $\mathbb{F}_{1^\ell}$-zeta function is thus given by \begin{equation} \zeta^{\mathbb{F}_{1^\ell}}_{\Gamma}(s)\ \ =\ \ \frac{(\ell(s - 1))^I}{(\ell s)^{E + I}}\cdot\displaystyle \prod_{k = 1}^m(\ell(s - k))^{-n_k}. \end{equation} If $D = \{d_1 = d \}$, $d > 1$ and $n_1 = 1$, then $I = 0$ and $E = 0$. This is the loose graph of affine $d$-space, and we obtain the $\mathbb{F}_{1^\ell}$-zeta function for $\mathbb{A}^d$ as above. (For the affine line, put $D = \emptyset$, $I = -1$ and $E = 1$.) \newpage
6dc37fa82d621423ba144f29334f49b5141f900d
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\subsection{Motivation} In a kinetic description of fluid motion the state of the gas is defined by a kinetic function $f(x,t,v),$ that determines the distribution of molecules at position $x$ and time $t$ according to the velocity $v,$ and a kinetic equation for $f,$ \begin{equation} \label{eq:Kinetic} \partial_t f{}+{}v\cdot\nabla{}_x f{}={}Q(f), \end{equation} where the right-hand side determines the changes in the kinetic density due to molecular interactions. In the kinetic models of gases, the collision operator $Q(f)$ verifies the following properties: \begin{enumerate} \item $Q$ has zero moments: \[ \int(1,v,|v|^2)Q(f)\,dv{}={}0; \] \item $Q(f){}={}0$ iff $f\in E_0,$ where $E_0$ is the set of minimizers of the problem \begin{equation*} \label{prob:min} \min\left\{ S(f)\,:\, f\geq0,\quad \int(1,v,|v|^2)f\,dv{}={}const.\right\} \end{equation*} with an entropy functional \[ S(f){}={}\int s(f)\,dv, \] were $s\,:\,\mathbb{R}\to\mathbb{R}$ is a convex, coercive function; \item interactions do not increase entropy: for any kinetic density $f,$ \[ \int Q(f)s'(f)\,dv\leq0. \] \end{enumerate} The above properties are due to the conservation of mass, momentum, and energy of molecular motion, and express the fact that molecular interactions have an effect on the kinetic density to ``relax" toward the set of equilibrium densities $E_0.$ Assuming that relaxation processes are instantaneous, the kinetic density takes values $f(x,t,\cdot)\in E_0,$ for all $(x,t),$ i.e, it is a function only of its $(1,v,|v|^2)$ moments, that we denote by $U{}={}(\rho,m,E)$ -- the macroscopic density, momentum and energy. The moments verify the system of Euler equations, \begin{equation} \label{eq:Euler} \int \left(\partial_t f{}+{}v\cdot\nabla{}_x f\right)\left(\begin{array}{c}1 \\ v\\|v|^2\end{array}\right)\,dv{}={}0, \end{equation} which should be supplemented with the initial/boundary conditions for a particular fluid flow in question. System of equations \eqref{eq:Euler} is a first order quasi-liner system of PDEs: \begin{equation} \label{eq:Hyperbolic} \partial_t U{}+{}\div_x F(U){}={}0, \end{equation} where the flux $F\,:\, \mathbb{R}^5\to\mathbb{R}^{5\times 3}.$ The conservation of entropy at the kinetic level ($Q(f){}={}0$) leads (for smooth kinetic functions) to the conservation of macroscopic entropy \[ \partial_t S{}+{}\div_x Q_e{}={}0, \] where \[ S(U){}={}\int s(f)\,dv,\quad Q_e(U){}={}\int vs'(f)\,dv. \] In flows away from vacuum, $\rho>0,$ the entropy is a strictly convex function of $U,$ and that makes system \eqref{eq:Hyperbolic} to be {\it symmetrizable, hyperbolic} system of conservation laws. The Cauchy problem for such systems is well-posed in classes of smooth functions, as was established in \cite{Garding, Kato}. Specifically, the following result holds, theorem 5.1.1 of \cite{Dafermos}. Consider a symmetrizable, hyperbolic system of $m$ conservation laws \begin{equation} \label{eq:Hyperbolic:2} \partial_t U{}+{}\div_x F(U){}={}G(U), \end{equation} with $F\in C^4(O)^{m\times d},$ $G\in C^3(O)^m,$ and entropy $S\in C^3(O),$ on an open subset $O\subset\mathbb{R}^d.$ Assume that $S$ is strictly convex on $O$ and the initial data $U_0(x)$ belong to a compact subset of $O,$ with \[ \nabla{} U_0\in H^l,\quad l>d/2. \] \begin{theorem*} There is a time interval $[0,T),$ on which system of equations \eqref{eq:Hyperbolic:2} with initial data $U_0$ has a unique classical solution $U(x,t).$ Solution belong to the class \[ \nabla{} U{}\in{}\cap_{k=0}^{l}C^k([0,T): H^{k-l}). \] \end{theorem*} The solution is constructed by a fixed point of a map, determined by a solution of the linearized equations \eqref{eq:Hyperbolic:2}. Alternatively, the solution can obtained in a zero viscosity limit, using the theory of parabolic systems. In the analysis of non-equilibrium flows, it might be desirable to approximate equation \eqref{eq:Kinetic} by a closed system of PDEs for a finite set of macroscopic parameters (moments). A moment closure is an example of such reduction of dimension, which is based on an ubiquitous idea of Galerkin approximation. A generic form of a moment closure was described in \cite{Gorban1, Gorban2}, and \cite{Levermore}. The moment closures of \cite{Grad:moments, Dreyer} are earlier, notable examples of such approximations. Let us consider moment closures is some detail, following \cite{Levermore} for the presentation. The closures are taken with respect to the moments \[ \int fl_i(v)\,dv,\quad i=1..k, \] where $\{l_i\}_{i=1}^k$ is a set ``elements". Given the tendency of the kinetic density to an equilibrium in $E_0,$ it is reasonable to include polynomials $\{1,v,|v|^2\}$ among moments. Further restrictions on the set of moments can be imposed by requirement that the corresponding system of PDEs \eqref{eq:Hyperbolic:2} verifies Galilean and rotational symmetries, see \cite{Levermore}. Denote \[ E^*{}={}\span\{l_i,\,i=1..k\}. \] Consider a minimization problem \begin{equation} \label{minimization} \min\left\{ S(f)\,:\, \int fl_i(v)\,dv{}={}U_i=const.\right\} \end{equation} where the entropy functional is as above. For a given vector $U{}={}(U_0,..,U_k),$ the problem (typically) has a unique minimizer $f_0,$, determined by the conditions \begin{eqnarray*} \exists g_0\in E^*,\quad g_0\in\partial S(f_0),\\ \int f_0 l_i(v)\,dv{}={}U_i,\quad i=1..k. \end{eqnarray*} The first condition defined $g_0{}={}s'(f_0)\in E^*,$ or $f_0{}={}(s^*)'(g_0),$ where $s^*$ is the Legendre transform of $s.$\footnote{ We always assume the duality pairing between functions to be $ \langle f,g \rangle{}={}\int fg\,dv. $ A functional space $V$ for kinetic density $f$ can be define \[ V{}={}\left\{f\,:\, \int (1+|v|^{m_0})|f|\,dv<+\infty\right\}, \] where $m_0$ is the highest degree of polynomials in the set $\{l_i\}.$ In this section we proceed informally, identifying, for example, a subdifferential $\partial S(f)$ with a function $s'(f).$ } Define $E$ to be a set of minimizers $f_0$ for all choices of moments vector $U.$ A moment closure of \cite{Levermore} is defined as a system of equations \begin{equation} \label{eq:Moment_closure:1} \begin{cases} \int \left( \partial_t f{}+{}v\cdot\nabla{}_x f- Q(f)\right)l_i(v)\,dv{}={}0,\quad i=1..k,\\ f(x,t,\cdot)\in E, \end{cases} \end{equation} which is equivalent to a first order quasi-linear system of PDEs of type \eqref{eq:Hyperbolic:2} in unknowns $U_i(x,t){}={}\int f(x,t,v)l_i(v)\,dv.$ The corresponding entropy equation reads: \[ \partial_t\int s(f)\,dv{}+{}\div_x\int vs'(f)\,dv{}={}\int Q(f)s'(f)\,dv. \] Following \cite{Levermore}, the convexity of $S=S(U)$ can be conveniently expressed using dual, hydrodynamic variables $\alpha_i=\alpha_i(x,t),$ related to $f$ by the condition \[ s'(f(x,t,v)){}={}\sum_i \alpha_i(x,t)l_i(v)\in E^*. \] Indeed, the definition of $U$ is stated as \begin{equation} \label{eq:Uj} U_j{}={}\int fl_j(v)\,dv{}={}\int (s')^{-1}(\sum_i\alpha_il_i(v))\,dv. \end{equation} Since $S(U){}={}\int s(f)\,dv$ we obtain that \[ \partial_{U_j}S{}={}\int s'(f)\partial_{U_j}f\,dv{}={}\int (\sum_i\alpha_il_i)\partial_{U_j}f\,dv{}={}\alpha_j. \] Thus, $\partial^2_{U_jU_k}S{}={}\partial_{U_k}\alpha_j.$ Also, from \eqref{eq:Uj} we also get \[ \partial_{\alpha_i}U_j{}={}\int\frac{l_j(v)l_i(v)}{s''(\sum\alpha_il_i)}\,dv, \] which makes $\partial_\alpha U$ a positive definite. Its inverse is positive definite as well, and so is $\nabla{}^2_{U}S.$ \subsubsection{Orthogonality in primal variables} An alternative way to derive system \eqref{eq:Moment_closure:1} is to use the differential structure of sets appearing in the optimization problem \eqref{minimization}, following the approach of \cite{Gorban1, Gorban2}. Consider a kinetic density $f$ and a set of density with the same $l_i$ moments: \[ M(f){}={}\left\{ \tilde{f}\,:\, \int (\tilde{f}-f)l_i(v)\,dv{}={}0,\,i=1..k\right\}. \] Let $f_0\in M_f\cap E_0$ be the minimizer of problem \eqref{minimization}. Define the tangent plane to $M_f$ at $f_0$ as \[ T_{M(f)}=\left\{\hat{f}\,:\, \int \hat{f}l_i(v){}={}0,\,i=1..k\right\}. \] Since it is independent of $f$ we simply write $T_M.$ The set $E,$ defined above, can be defined as \[ f_0\in E\iff \int s'(f_0)\hat{f}\,dv{}={}0,\quad \forall \hat{f}\in T_{M}. \] With this definition we can define the tangent space to $E$ at $f_0,$ denoted by $T_E(f_0),$ as the set of vectors $\bar{f}$ such that \[ \forall \hat{f}\in T_{M},\quad \lim_{t\to0}t^{-1}\int s'(f_0+t\bar{f})\hat{f}\,dv{}={}0. \] The condition for being an tangent vector can be equivalently stated as \begin{equation} \label{eq:ortho:1} \int s''(f_0)\hat{f}\bar{f}\,dv{}={}0,\quad \forall \hat{f}\in T_M. \end{equation} One can interpret this condition as orthogonality between the tangent spaces to $M$ and $E,$ in a weighted $L^2$ space with scalar product \[ (\hat{f},\bar{f}){}={}\int \hat{f}\bar{f}\,s''(f_0)dv. \] In this notation, $T_M{}={}\left(T_E(f_0)\right)^{\perp}.$ Now, the moment closure system \eqref{eq:Moment_closure:1} can be equivalently expressed as a differential inclusion \begin{equation} \label{eq:Levermore} \begin{cases} \partial_tf +v\cdot\nabla{}_x f-Q(f)\in \left(T_E(f_0)\right)^{\perp},\\ f(x,t,\cdot)\in E, \end{cases} \end{equation} or, by noticing that $\partial_t f\in T_E(f),$ as an equation \begin{equation} \label{eq:Gorbin} \begin{cases} \partial_tf{}={} \mbox{\rm proj}_{T_E(f)}(-v\cdot\nabla{}_x f+Q(f)),\\ f(x,t,\cdot)\in E. \end{cases} \end{equation} The later equation is in the form used in \cite{Gorban1, Gorban2}. \subsubsection{Orthogonality in dual variables} Finally, lets consider yet another way to pose a moment closure, expressing \eqref{eq:Levermore} is dual kinetic variables \[ l{}={}s'(f),\quad (l\in\partial S(f)). \] Condition \eqref{eq:ortho:1} carries over to $l$--variables and becomes, \begin{equation} \label{eq:ortho:1.5} \int (s^*)''(l_0)\hat{l}\bar{l}\,dv{}={}0,\quad \forall \hat{l}\in E^*,\,\hat{l}\in T_{M^*(l_0)}. \end{equation} It expresses the orthogonality of linear space $E^*$ and tangent space to $M^*(l_0){}={}\{l\,:\, \int ((s^*)'(l)-(s^*)'(l_0))l_i(v)\,dv{}={}const.\}$ Orthogonality condition \eqref{eq:ortho:1.5} has another interpretation. Recall from the Convex Analysis, \cite{Ekland}, proposition 2.4, that the values of the primal problem \[ \min\left\{ S(f)\,:\, f\geq0,\,\int(f-\check{f})l_i(v)\,dv{}={}0,\,\,i=1..k\right\} \] (for a fixed $\check{f}\geq0,$) and its dual \begin{equation} \label{prob:dual} \sup\left\{ \int l\check{f}\,dv{}-{}S^*(l)\,:\,l\in E^*\right\} \end{equation} where $S^*(l){}={}\int s^*(l)\,dv$ is the Legendre transform of $S(f),$ are equal. The minimizer $f_0$ and the maximizer $l_0$ are determined by the conditions \begin{eqnarray} l_0\in\partial S(f_0),\quad (l_0{}={}s'(f_0),\,f_0{}={}(s^*)'(l_0)), \notag \\ \label{cond:moments} \int (f_0-\tilde{f})l_i(v)\,dv{}={}0,\quad i=1..k. \end{eqnarray} Choose $h$ that verifies condition \eqref{eq:ortho:1.5}: \begin{equation*} \int (s^*)''(l_0)hl_i\,dv{}={}0,\quad i=1..k, \end{equation*} and let $l_0\in E^*.$ Consider the above maximization problem with $\tilde{f}{}={} (s^*)'(\tilde{l}),$ $\tilde{l}{}={}l_0+th.$ Let $l^{t}_0$ be the corresponding maximizer. Re-writing condition \eqref{cond:moments} as \[ \int ( (s^*)'(l_0+th)-(s^*)'(l^{t}_0))l_i(v)\,dv{}={}0. \] we see that due to assumptions on $h,$ \[ l^{t}_0-l_0{}={}o(t),\quad l_0{}={}\mbox{\rm proj}_{E^*}(\tilde{l}), \] where the projection with respect ot weighted $L^2$ norm. In other words, the solution of the optimization problem \eqref{prob:dual} with $\tilde{f}{}={} (s^*)'(\tilde{l}),$ coincides (to the first order of distance from $E^*$) with the projection of $\tilde{l}$ onto $E^*.$ This considerations allow us to re-write the moment closure equations in \eqref{eq:Levermore} as a differential inclusion in dual variable $l:$ \begin{equation} \label{eq:Misha} \begin{cases} \partial_t l{}+{}v\cdot\nabla{}_x l{}-{}\tilde{Q}(l)\in (E^*(l))^{\perp},\\ l(x,t,\cdot)\in E^*, \end{cases} \end{equation} where the collision operator equals \[ \tilde{Q}(l){}={}\frac{Q((s^*)'(l))}{(s^*)''(l)}. \] We certainly could have arrived at \eqref{eq:Misha} directly from \eqref{eq:Levermore}, but the above arguments show that there is also an underlying variational principle. Let us remark, that the dual kinetic variables and weighted $L^2$ spaces, discussed above, has been in use in the theory of Boltzmann equations since the work of Hilbert\cite{Hilbert}, where they appear in a context of linearization of \eqref{eq:Kinetic}. In a typical linearization analysis, kinetic density is represented in terms of a dual variable $h,$ as $f{}={}f_0(1+h),$ where $f_0$ is a Maxwellian ($f_0\in E_0$). \subsection{Results} In this work we establish the existence of classical solutions to a class of systems of PDEs \eqref{eq:Hyperbolic:2} corresponding to \eqref{eq:Misha}, by solving the later problem in a space of kinetic functions. We assume that initial data $l^0(x,\cdot)$ take values in $E^*,$ ranging in a neighborhood of a constant state $\bar{l}\in E^*,$ and, is in Sobolev's $H^3$ space, as a function of $x.$ Two types of collision operators are considered. In the first model, the collision operator is absent, $\tilde{Q}=0,$ and we are dealing with projection of a transport equation onto $E^*.$ In the second model, we consider a non-linear BGK--type operator, in dual variable $l:$ \[ \tilde{Q}(l){}={}\Pi^0_{l}-l,\quad \Pi^0_l{}={}\mbox{\rm proj}_{E_0^*(l)}(l). \] The corresponding operator in the primal variable, $Q(f){}={}(s^*)''(l)\tilde{Q}(l),$ $l{}={}(s)'(f),$ verifies properties (1), (2), and (3) of the collision operators, stated at the beginning of the Introduction. It is a first order (in the distance from $E_0$) approximation of the classical BGK operator. We show that classical solutions of \eqref{eq:Misha} can be constructed in zero limit of step $h,$ of a time-discretization of \eqref{eq:Misha}, in which transport, collision and projection are computed in succession, over time intervals $(nh, (n+1)h],$ as defined in \eqref{Projection}, \eqref{approx:continuous:1}, \eqref{Projection:2}. The analysis is based on entropy estimates for the kinetic density and its $x$--derivatives, that are similar to the estimates for linearized Boltzmann equation obtained in \cite{Grad:asymptote}. The choice of the approximating scheme is not accidental. In fact, it is the convergence of that particular scheme that we're interested in, rather than finding a new way to prove the existence of classical solutions for a class of PDEs \eqref{eq:Hyperbolic:2}. The reason for this, is an observation that the time discretization of \eqref{eq:Misha}, with $E^*{}={}E_0^*,$ and $\tilde{Q}{}={}0,$ is linked to a hydrodynamic limit of a gas-kinetic equation \eqref{eq:Kinetic} with the right-hand side containing a large factor $h^{-1}.$ The projection of the kinetic density to an equilibrium $E_0$ can be loosely related to result of collisions, since the later amounts to relaxing the density toward equilibrium, while conserving the moments. In this respect, our convergence result can be compared with the works on the fluid dynamic limit of Boltzmann and related equations, with some representative results given in \cite{Nishida_Boltzmann, Caflisch, CaflischPapa, BB, BV}. In the present setting, the convergence takes place for all times $t$ inside an interval $[0,T]$ determined by the initial data. No initial layer is present since the dynamics is smooth and starts from the target manifold $E^*.$ Finally, let us mention that discrete transport--projection approximations appear in many areas of PDEs. Some examples of the method, in the context of Boltzmann equation and scalar conservation laws can be found in \cite{Gorban3, Brenier1, Brenier2}. \end{section} \begin{section}{Transport equation} \subsection{Notation and auxiliary lemmas} Let $\{l_i(v)\}_{i=1}^{k}$ be a set of $k+1$ linearly independent on an open set of $\mathbb{R}^d$ polynomials. We assume that the highest degree polynomial is $l_k(v){}={}|v|^{m_0},$ for some $m_0>0,$ and the lowest degree polynomial $l_1(v){}={}1.$ Denote \[ E^*{}={}\span\{l_i,\:\,i=1..k\}. \] We choose entropy density to be $s(f){}={}f^p,$ with $p\in(1,6/5).$ The Legendre transform of $s$ equals \[ s^*(l){}={}c_p(l_+)^{p/{p-1}}, \] with $c_p{}={}\frac{p-1}{p^{p/(p-1)}},$ and $l_+$ is a positive part of $l.$ For a notational convenience we define the weight function with respect to variable $-l,$ rather than $l,$ \[ w(l){}={}\bar{c}_p(l_{-})^{\frac{2-p}{p-1}}, \] where $\bar{c}_p>0,$ and $l_{-}\geq0$ -- the negative part of $l.$ For the range of $p$ defined above, $w\in C^4(\mathbb{R}),$ convex function, supported on $l\geq0.$ With the above choice of an entropy $s$, kinetic densities $f\in E,$ are smooth and compactly supported functions. The analysis critically depends on last two properties. We consider a Cauchy problem \begin{equation} \label{model:1} \begin{cases} \partial_t l {}+{}v\cdot\nabla{}_x l{}\in{} (E^*(l))^\perp, & (x,t,v)\in \mathbb{R}^3\times(0,T)\times\mathbb{R}^3,\\ l(x,t,\cdot)\in E^*, & (x,t)\in\mathbb{R}^4_+, \\ l(x,0,v){}={}l^0(x,v), & (x,v)\in\mathbb{R}^6, \end{cases} \end{equation} where notation $E^*(l)$ denotes space $E^*$ with weighted $L^2$ norm \[ \|\tilde{l}\|^2_{w(l)}{}={}\int \tilde{l}^2\,w(l)\,dv. \] Following \cite{Dafermos}, we make the following definition. \begin{definition} A classical solution $l,$ of \eqref{model:1} as a Lipschitz continuous in $(x,t,v)$ function $l(x,t,\cdot),$ such that for a.e. $(x,t),$ $t\geq0,$ the set of $k+1$ equations holds, \[ \int \left( \partial_t l{}+{}v\cdot\nabla{}_x l\right)l_i(v)\,dv{}={}0,\quad i=1..k; \] for all $(x,t),t\geq0,$ $l(x,t,\cdot)\in E^*;$ for $t=0,$ and all $(x,v),$ $l(x,0,v){}={}l^0(x,v).$ \end{definition} A care should be taken to avoid degenerate situation when the weight $w(l)$ is zero. \begin{definition} Let $(R,\delta_1,r,\delta_2)$ be positive numbers. We say that $l\in E^*$ has property $P(R,\delta_1,r,\delta_2)$ if \begin{enumerate} \item $\forall v,$ with $|v|>R,$ $ l(v)\geq0; $ \item $\forall v,$ with $|v|<R,$ $ l(v)\geq -\delta_1; $ \item there is a ball $B_r$ of radius $r$ such that $ l(v)\leq -\delta_2, $ for all $v\in B_r.$ \end{enumerate} \end{definition} Let $\bar{l}{}={}\sum_i\bar{\gamma}_il_i(v)\in E^*$ be a constant reference state with \[ \bar{\gamma}_0<0,\quad \bar{\gamma}_k>0. \] The solutions we construct are in a neighborhood of $\bar{l}.$ The following lemma is easily verified. \begin{lemma} \label{lemma:nondegenerate} There are positive numbers $(R,\delta_1,r,\delta_2)$ such that \[ \bar{l}\in P\left(R/2,\delta_1/2,2r,2\delta_2\right), \] and $\forall \varepsilon>0$ there is $\Delta>0,$ such that, if $l\in E^*$ and \[ \int_{B_R}|l-\bar{l}|^2\,dv\leq \Delta, \] then \[ l\in P\left(\frac{1+\varepsilon}{2}R,\frac{1+\varepsilon}{2}\delta_1, \frac{2}{1+\varepsilon}r,\frac{2}{1+\varepsilon}\delta_2\right). \] \end{lemma} Let $\Delta_2$ be a number corresponding to $\varepsilon=1,$ and $\Delta_1<\Delta_2$ be the number corresponding to $\varepsilon{}={}1/2,$ in the above lemma. The numbers $(R,\delta_1,r,\delta_2)$ and the corresponding balls $B_R,$ $B_r$ from the definition of the property $P,$ will be fixed in the following analysis. Let the initial date $l^0(x,v)$ be such that \begin{equation} \label{cond:initial} \begin{cases} \forall x\in\mathbb{R}^3, \quad l^0(x,\cdot)\in E^*,\\ l^0-\bar{l}\in L^2(B_R; H^3(\mathbb{R}^3)), \\ \sup_x\int_{B_R}|l^0(x,v)-\bar{l}(v)|^2\,dv\leq \Delta_1. \end{cases} \end{equation} We use a weighted ``norm" for $l,$ defined as: \begin{equation} \label{X} \|l\|_{X(l)}{}={}\left(\iint (\sum_{\alpha,|\alpha|\leq 3}|D^\alpha_x l|^2)w(l)\,dvdx\right)^{1/2}. \end{equation} \subsection{Statement of the result} The discrete-time algorithm approximating differential inclusion \eqref{model:1} is defined in the following way. Let $h>0$ be the time step and define $N{}={}\lceil T/h\rceil.$ Given the values of $l^{n-1},$ we define \begin{equation} \label{Projection} l^n(x,\cdot){}={}\mbox{\rm proj}_{E^*(l^{n-1}(x,\cdot))}l^{n-1}(x-h\cdot,\cdot), \end{equation} where the projection is in weighted $L^2$ space with weight $w(l^{n-1}).$ If the weight is not zero identically, the projection is uniquely defined by conditions \begin{equation} \label{Projection_cond} \forall x\in\mathbb{R}^3,\,i=1..k,\quad \int l^n(x,v)l_i(v)w(l^{n-1}(x,v))\,dv{}={} \int l^{n-1}(x-hv,v)l_i(v)w(l^{n-1}(x,v))\,dv. \end{equation} In what follows we use the shorthand notation $w^n(x,v){}={}w(l^n(x,v)).$ \begin{theorem} \label{theorem:1} Let $\bar{l},l^0$ be as described above and $h\in(0,1].$ There is time $T>0,$ independent of $h,$ such that all functions $l^n,$ $n=0..\lceil T/h\rceil,$ in \eqref{Projection} are well defined, and the family $\{l^h\}$ of interpolations of $l^n$'s, defined in \eqref{approx:continuous:1}, converges as $h\to0$, uniformly on compact set in $\mathbb{R}^3\times[0,T]\times\mathbb{R}^3$ to a unique classical solution of \eqref{model:1}. \end{theorem} \begin{remark} The uniqueness of classical solutions (in fact strong-weak uniqueness) follows from the uniqueness of classical solutions of corresponding PDEs \eqref{eq:Hyperbolic:2}, with strictly convex entropy, see, for example, \cite{Dafermos}, theorem 5.3.1. \end{remark} The theorem is based on the fact that functions $l^n$ are bounded in strong norms, which can be heuristically explained as follows. Suppose that the approximation $\{l^n\}$ is well-defined. Denote by $\hat{l}^{n-1}(x,v){}={}l^{n-1}(x-hv,v),$ and for any $x\in\mathbb{R}^3,$ the distance \[ |l(x,\cdot)-\bar{l}(\cdot)|^2_{w^{n}}{}={}\int |l(x,v)-\bar{l}(v)|^2w(l^n(x,v))\,dv. \] From \eqref{Projection} we obtain \[ |l^n-\bar{l}|^2_{w^{n-1}}\leq |\hat{l}^{n-1}-\bar{l}|^2_{w^{n-1}}. \] We will show that \begin{equation} \label{est:zero} |l^n-\bar{l}|^2_{w^n}{}\leq{}|\hat{l}^{n-1}-\bar{l}|^2_{w(\hat{l}^{n-1})}{}+{}R_n, \end{equation} where $R_n$ accounts for changes in the weights from $w^n$ to $w^{n-1},$ and from $w(l^{n-1})$ to $w(\hat{l}^{n-1}).$ The remainder is such that \begin{equation} \label{R1} \int R_n\,dx{}={}O(h), \end{equation} provided that all $l^n$ are smooth, as measured by \eqref{X}. Integrating \eqref{est:zero} we obtain \begin{equation} \label{est:one} \int|l^n-\bar{l}|^2_{w^n}\,dx{}\leq{}\int|l^{n-1}-\bar{l}|^2_{w^{n-1}}\,dx{}+{}O(h). \end{equation} To estimate the spacial derivatives we use an orthogonal decomposition (in topology of $L^2_{w^{n-1}}(\mathbb{R}^3)$) \[ (l^n-\bar{l}){}+{}(\hat{l}^{n-1}-l^n){}={}\hat{l}^{n-1}-\bar{l}, \] to obtain \[ D^\alpha_x(l^n-\bar{l}){}+{}D^\alpha_x(\hat{l}^{n-1}-l^n){}={}D^\alpha_x(\hat{l}^{n-1}-\bar{l}), \] which implies that \[ |D^\alpha_x(l^n-\bar{l})|^2_{w^{n-1}}{}\leq{}|D^\alpha_x(\hat{l}^{n-1}-\bar{l})|^2_{w^{n-1}}{}+{}R^\alpha_n, \] where $R^\alpha_n$ accounts for spacial derivatives of the weight function, and has property \eqref{R1}. From this, by changing the weights, \[ |D^\alpha_x(l^n-\bar{l})|^2_{w^{n}}{}\leq{}|D^\alpha_x(\hat{l}^{n-1}-\bar{l})|^2_{w(\hat{l}^{n-1})}{}+{}\tilde{R}^\alpha_n, \] which leads to \begin{equation} \label{est:two} \int|D^\alpha_x(l^n-\bar{l})|^2_{w^n}\,dx{}\leq{}\int|D^\alpha_x(l^{n-1}-\bar{l})|^2_{w^{n-1}}\,dx{}+{}O(h). \end{equation} With derivatives of order 3, \eqref{est:one}, \eqref{est:two} lead to a priori estimates on $l^n$ as measured by $X(l^n).$ The above arguments are formalized in lemmas \ref{lemma:1}--\ref{lemma:Final} below, after which we show that properly interpolated in time sequence $l^n$ converges to a classical solution of \eqref{model:1}. \subsection{Proof of theorem \ref{theorem:1}} Consider the sequence $\{l^n\}$ determined from the initial data and \eqref{Projection}. We will assume in this section the following hypothesis. \begin{hypothesis} For all $n=1..N,$ and all $x\in\mathbb{R}^3,$ \[ l^n(x,\cdot)\in P(R,\delta_1,r,\delta_2). \] \end{hypothesis} Under this hypotheses, functions $l^n=\sum_i\gamma_i^n(x)l_i(v)$ from \eqref{Projection} are well-defined and we proceed to derive energy estimates. \begin{lemma} \label{lemma:1} There is $C>0,$ independent of $(n,h),$ such that \begin{equation} \label{est:1} \sup_x \int_{B_R}|l^{n-1}(x-hv,v)-l^{n-1}(x,v)|^2\,dv\leq Ch^2\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}. \end{equation} \end{lemma} \begin{proof} The estimate follows directly by applying lemma \ref{Sobolev:1} and lemma \ref{Sobolev:2} from the Appendix to function $l^{n-1}.$ \end{proof} \begin{lemma} There is $C>0,$ independent of $(n,h),$ such that \begin{equation} \label{est:2} \sup_x\int_{B_R}|l^{n}(x,v)-l^{n-1}(x,v)|^2\,dv\leq Ch^2\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}, \end{equation} and \begin{equation} \label{eneqry:00} \int \int_{B_R} |l^n-l^{n-1}|^2\,dvdx\leq Ch^2\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}. \end{equation} \end{lemma} \begin{proof} Indeed, using \eqref{Projection} we obtain \[ \sup_x\int|l^{n}(x,v)-l^{n-1}(x,v)|^2w^{n-1}\,dv\leq C\sup_x\int_{B_R}|l^{n-1}(x-hv,v)-l^{n-1}(x,v)|^2\,dv, \] and the first statement of the lemma follows from \eqref{est:1}, and the facts that $w^{n-1}$ is strictly positive on the ball $B_r,$ and norms with respect to balls $B_r$ and $B_R$ are equivalent. Similarly, from conditions \eqref{Projection} and Hypothesis 1 we get \[ \iint |l^n-l^{n-1}|^2w^{n-1}\,dvdx\leq C\int_{B_R}\int |l^{n-1}(x-hv,v)-l^{n-1}(x,v)|^2\,dxdv \] which is less than $ Ch^2\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}, $ by estimate \ref{est:1}. \end{proof} \begin{lemma} \label{lemma:3} There is $C>0,$ independent of $(n,h),$ such that \[ \int\int_{B_R} |D^\beta l^n-D^\beta l^{n-1}|^2\,dxdv{}\leq{}Ch^2(\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}{}+{} \|l^{n-1}-\bar{l}\|^4_{X(l^{n-1})}), \] for any multi-index $\beta,$ with $|\beta|=1;$ \[ \int\int_{B_R} |D^\beta l^n-D^\beta l^{n-1}|^2\,dxdv{}\leq{}Ch^2(\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}{}+{} \|l^{n-1}-\bar{l}\|^4_{X(l^{n-1})}{}+{}\|l^{n-1}-\bar{l}\|^6_{X(l^{n-1})}), \] for any multi-index $\beta,$ with $|\beta|=2.$ \end{lemma} \begin{proof} Let $D$ be generic notation for the first derivative in $x.$ By applying it to \eqref{Projection} we find that \begin{eqnarray} \label{diff:1} \int(Dl^n - Dl^{n-1})l_iw^{n-1}\,dv &=& \int (Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v))l_iw^{n-1}\,dv\\ &&+\int (l^{n-1}(x-hv,v)-l^n(x,v))l_iw'(l^{n-1})Dl^{n-1}\,dv. \notag \end{eqnarray} It follows that for any $x,$ \begin{eqnarray*} \int_{B_R}|Dl^n- Dl^{n-1}|^2\,dv{}&\leq&{} C\int_{B_R}|Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v)|^2\,dv\\ &&{}+{} C\int_{B_R}|l^{n-1}(x-hv,v)-l^{n-1}(x,v)|^2|Dl^{n-1}|^2\,dv\\ &&{}+{}\int_{B_R}|l^n-l^{n-1}|^2|Dl^{n-1}|^2\,dv\\ &\leq& C\int_{B_R}|Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v)|^2\,dv\\ &&{}+{}\int_{B_R}|Dl^{n-1}|^2\,dv \int_{B_R}|l^{n-1}(x-hv,v)-l^{n-1}(x,v)|^2\,dv\\ &&{}+{}\int_{B_R}|Dl^{n-1}|^2\,dv \int_{B_R}|l^{n}-l^{n-1}|^2\,dv. \end{eqnarray*} Using \eqref{est:1}, \eqref{est:2} and lemma \ref{Sobolev:2} we get the first inequality in the lemma. \[ \int\int_{B_R}|Dl^n- Dl^{n-1}|^2\,dv{}\leq{} Ch^2( \|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})} + \|l^{n-1}-\bar{l}\|^4_{X(l^{n-1})}). \] The second inequality is obtained by differentiating \eqref{diff:1} and repeating the arguments above. \end{proof} \begin{lemma}[Zero order entropy estimate] There is $C>0,$ independent of $(n,h),$ such that \begin{equation} \label{energy:0} \iint |l^n-\bar{l}|^2w^{n}\,dvdx{}\leq {} \iint |l^{n-1}-\bar{l}|^2w^{n-1}\,dvdx+Ch\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}. \end{equation} \end{lemma} \begin{proof} To get the estimate we use \eqref{Projection} to write \begin{multline*} \iint |l^n-\bar{l}|^2w^{n-1}\,dvdx{}\leq{}\iint|l^{n-1}(x-hv,v)-\bar{l}|^2w^{n-1}dvdx\\ {}\leq{} \iint|l^{n-1}(x,v)-\bar{l}|^2w^{n-1}\,dvdx{}+{}\iint|l^{n-1}(x,v)-\bar{l}|^2|w^{n-1}-w(l^{n-1}(x+hv,v))|\,dvdx \end{multline*} The last term can be estimated as \begin{multline*} \iint|l^{n-1}(x,v)-\bar{l}|^2|w^{n-1}-w(l^{n-1}(x+hv,v))|\,dvdx\\ {}\leq{} C\int\int_{B_R}|l^{n-1}(x,v)-\bar{l}|^2|l^{n-1}(x,v)-l^{n-1}(x-hv,v)|\,dvdx\\ {}\leq{}C\int \sup_{v\in B_R} |l^{n-1}(x,v)-l^{n-1}(x-hv,v)|\int_{B_R}|l^{n-1}(x,v)-\bar{l}|^2\,dv\,dx\\ {}\leq{}C \int \int_{B_R}|l^{n-1}(x,v)-l^{n-1}(x-hv,v)|\,dv\int_{B_R}|l^{n-1}(x,v)-\bar{l}|^2\,dv\,dx\\ {}\leq{}C\left(\sup_x\int_{B_R}|l^{n-1}(x,v)-l^{n-1}(x-hv,v)|\,dv\right)\left(\int\int_{B_R}|l^{n-1}(x,v)-\bar{l}|^2\,dvdx\right)\\ {}\leq{} Ch\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}, \end{multline*} where we used equivalence of norms in $v,$ and \eqref{est:1}. Using this in the previous inequality results in the statement of the lemma. \end{proof} Higher order energy estimates are obtained by differentiating conditions \eqref{Projection} and following the arguments of the previous lemma. \begin{lemma}[Third order entropy estimates] \label{lemma:energy:3} There is $C>0,$ independent of $(n,h),$ such that for any multi-index $\alpha,$ with $|\alpha|=3,$ \begin{multline} \label{energy:3} \iint |D^\alpha l^n|^2w^n\,dxdv{}\leq{}\iint |D^\alpha l^{n-1}|^2w^{n-1}\,dxdv{}+{}Ch\left( \|l^n-\bar{l}\|^3_{X(l^n)}{}+{} \|l^n-\bar{l}\|^2_{X(l^n)}\right.\\ \left.{}+{}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}{}+{}\|l^{n-1}-\bar{l}\|^3_{X(l^{n-1})}{}+{}\|l^{n-1}-\bar{l}\|^5_{X(l^{n-1})}\right). \end{multline} \end{lemma} \begin{proof} After differentiating \eqref{Projection} by $D^\alpha{}={}\partial_{x_1}^{\alpha_1}\partial_{x_2}^{\alpha_2}\partial_{x_3}^{\alpha_3}$ we obtain \begin{eqnarray*} \int D^\alpha l^nl_iw^{n-1}\,dv&=&\int D^\alpha l^{n-1}(x-hv,v)l_i w^{n-1}\,dv\\ &&{}-{} \int (l^n-l^{n-1})l_iD^\alpha w^{n-1}\,dv{}-{}\int(l^{n-1}(x,v)-l^{n-1}(x-hv,v))l_iD^\alpha w^{n-1}\\ &&{}-{}\sum_{\beta,\gamma}\int D^\beta(l^n-l^{n-1})l_iD^\gamma w^{n-1}\,dv\\ &&{}-{}\sum_{\beta,\gamma}\int D^\beta(l^{n-1}(x,v)-l^{n-1}(x-hv,v))l_iD^\gamma w^{n-1}\,dv,\quad i=1..k, \end{eqnarray*} where multi-indices's $\beta,\gamma$ are such that $\beta+\gamma=\alpha,$ $|\beta|,|\gamma|>0.$ From this we obtain \begin{eqnarray*} \int |D^\alpha l^n|^2w^{n-1}\,dv &\leq& \int|D^\alpha l^{n-1}(x-hv,v)|^2w^{n-1}\\ &&+ 2\int|l^n-l^{n-1}||D^\alpha l^n||D^\alpha w^{n-1}|\,dv\\ &&+ 2\int|l^{n-1}(x,v)-l^{n-1}(x-hv,v)||D^\alpha l^n||D^\alpha w^{n-1}|\,dv\\ &&+2\sum_{\beta,\gamma}\int |D^\beta(l^n-l^{n-1})||D^\alpha l^n||D^\gamma w^{n-1}|\,dv\\ &&+2\sum_{\beta,\gamma}\int |D^\beta(l^{n-1}(x,v)-l^{n-1}(x-hv,v))||D^\alpha l^n||D^\gamma w^{n-1}|\,dv. \end{eqnarray*} Labeling the last four terms as $I_1,..,I_4,$ we write \begin{eqnarray} \label{l:1} \iint |D^\alpha l^n|^2w^{n}\,dvdx &\leq& \iint|D^\alpha l^{n-1}(x-hv,v)|^2w(l^{n-1}(x-hv,v))\,dvdx\\ && +\iint|D^\alpha l^{n-1}(x-hv,v)|^2|w(l^{n-1}(x-hv,v)-w(l^{n-1}(x,v)|\,dvdx \notag \\ &&+ \iint|D^\alpha l^n|^2|w^n-w^{n-1}|\,dvdx{}+{}\sum_j \int I_j\,dx. \notag \end{eqnarray} In this way we obtained inequality \begin{eqnarray*} \iint |D^\alpha l^n|^2w^{n}\,dvdx &\leq& \iint|D^\alpha l^{n-1}|^2w^{n-1}\,dvdx\\ &&{}+{}\int J_1\,dx+\int J_2\,dx{}+{}\sum_j \int I_j\,dx, \end{eqnarray*} where by $J_1,J_2$ we denote the second and the third terms on the right in \eqref{l:1}. It remains to show that integrals of $J_i's$ and $I_j's$ are of the order $h.$ \begin{eqnarray*} \int J_1\,dx &\leq&C\int_{B_R}\left(\int |D^\alpha l^{n-1}(x-hv,v)|^2\,dx\right) \sup_x|l^{n-1}(x-hv,v)-l^{n-1}(x,v)|\,dv\\ &\leq& Ch\int_{B_R}\left(\int |D^\alpha l^{n-1}(x-hv,v)|^2\,dx\right) \|l^{n-1}-\bar{l}\|_{H^3(\mathbb{R}^3)}\,dv\\ &\leq& Ch\sup_{v\in B_R}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}{}\leq{}Ch \|l^{n-1}-\bar{l}\|^3_{X(l^{n-1})}. \end{eqnarray*} \begin{eqnarray*} \int J_2\,dx &\leq& C\int_{B_R}\int |D^\alpha l^n|^2|l^n-l^{n-1}|\,dvdx\\ &\leq& C\int \left(\int_{B_R}|l^n-\bar{l}|^2\,dv\right)^{1/2}\int_{B_R}|D^\alpha l^n|^2\,dv\,dx\\ &\leq& C\sup_x \left(\int_{B_R}|l^n-\bar{l}|^2\,dv\right)^{1/2}\|l^n-\bar{l}\|^2_{X(l^n)}{}\leq{}Ch\|l^n-\bar{l}\|^3_{X(l^n)}, \end{eqnarray*} where in the last inequality we used \eqref{est:2}. Consider now \begin{eqnarray*} \int I_1\,dx &\leq&C\int\int_{B_R}|l^n-l^{n-1}||D^\alpha l^n||Dl^{n-1}|^3\,dvdx {}+{}C\int\int_{B_R}|l^n-l^{n-1}||D^\alpha l^n||Dl^{n-1}||D^2l^{n-1}|\,dvdx\\ &&+ C\int\int_{B_R}|l^n-l^{n-1}||D^\alpha l^n||D^3l^{n-1}|\,dvdx{}={}K_1+K_2+K_3, \end{eqnarray*} where $Dl,D^2l,D^3l$ denote all derivatives in $x$ of orders 1,2, and 3. Then, \begin{eqnarray*} K_1&\leq& C\int \sup_{v\in B_R}\left(|l^n-l^{n-1}||Dl^{n-1}|^2\right)\int_{B_R}|D^\alpha l^n|^2+|Dl^{n-1}|^2\,dv\,dx\\ &\leq& C\int \sup_{v\in B_R}|l^n-l^{n-1}|\sup_{v\in B_R}|Dl^{n-1}|^2\int_{B_R}|D^\alpha l^n|^2+|Dl^{n-1}|^2\,dv\,dx\\ &\leq& C\int \left(\int_{B_R}|l^n-l^{n-1}|^2\,dv\right)^{1/2}\int_{B_R}|Dl^{n-1}|^2\,dv\int_{B_R}|D^\alpha l^n|^2+|Dl^{n-1}|^2\,dv\,dx\\ &\leq& C\sup_x \left(\int_{B_R}|l^n-l^{n-1}|^2\,dv\right)^{1/2}\sup_x\int_{B_R}|Dl^{n-1}|^2\,dv\left(\|l^n-\bar{l}\|^2_{X(l^n)}{}+{}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}\right). \end{eqnarray*} Using estimates of lemma \ref{Sobolev:1} and \eqref{est:2}, we conclude that \begin{equation*} K_1\leq Ch\|l^{n-1}-\bar{l}\|^3_{X(l^{n-1})}\left(\|l^n-\bar{l}\|^2_{X(l^n)}{}+{}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}\right). \end{equation*} Estimates on $K_2,K_3$ are similar. They lead to: \begin{eqnarray} \label{est:I1} \int I_1\,dx&\leq&Ch\left(\|l^{n-1}-\bar{l}\|_{X(l^{n-1})}+\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}+\|l^{n-1}-\bar{l}\|^3_{X(l^{n-1})}\right)\\ &&{}\times{}\left(\|l^n-\bar{l}\|^2_{X(l^n)}{}+{}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}\right). \notag \end{eqnarray} The estimate on $I_2$ is analogous to that of $I_1,$ where we use \eqref{est:1} instead of \eqref{est:1}. It lead to the estimate \eqref{est:I1}. Consider $I_4:$ \begin{eqnarray*} \int I_4\,dx&\leq& C\int\int_{B_R}|Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v)||D^\alpha l^n||D^2l^{n-1}|\,dvdx\\ &&{}+{}C\int\int_{B_R}|Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v)||D^\alpha l^n||Dl^{n-1}|^2\,dvdx\\ &&{}+{}C\int\int_{B_R}|D^2l^{n-1}(x-hv,v)-D^2l^{n-1}(x,v)||D^\alpha l^n||Dl^{n-1}|\,dvdx{}={}M_1+M_2+M_3. \end{eqnarray*} \begin{eqnarray*} M_3 &\leq &C\sup_{v\in B_R,x\in\mathbb{R}^2}|Dl^{n-1}|\left(\int\int_{B_R}|D^2l^{n-1}(x-hv,v)-D^2l^{n-1}(x,v)|^2\,dvdx\right)^{1/2}\\ &&{}\times{}\left(\int\int_{B_R}|D^\alpha l^n|^2\,dvdx\right)^{1/2}\\ &\leq&Ch\|l^n-\bar{l}\|_{X(l^n)}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}, \end{eqnarray*} where in the last inequality we used lemmas \ref{Sobolev:1} and \ref{Sobolev:2}. By exactly the same argument, \[ M_2\leq Ch\|l^n-\bar{l}\|_{X(l^n)}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}, \] as well. Consider now \begin{eqnarray*} M_3&\leq& C\|D^\alpha l^n\|_{L^2(B_R\times\mathbb{R}^3)}\|D^2 l^{n-1}\|_{L^6(B_R\times\mathbb{R}^3)}\|Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v)\|_{L^3(B_R\times\mathbb{R}^3)}. \end{eqnarray*} Using Sobolev's inequalities we find that \begin{eqnarray*} M_3&\leq& C\|D^\alpha l^n\|_{L^2(B_R\times\mathbb{R}^3)}\|D^3 l^{n-1}\|_{L^2(B_R\times\mathbb{R}^3)}\|Dl^{n-1}(x-hv,v)-Dl^{n-1}(x,v)\|^{1/2}_{L^2(B_R\times\mathbb{R}^3)}\\ &&{}\times{}\|D^2l^{n-1}(x-hv,v)-D^2l^{n-1}(x,v)\|^{1/2}_{L^2(B_R\times\mathbb{R}^3)}. \end{eqnarray*} Using lemma \ref{Sobolev:2} this is less than \[ Ch\|l^n-\bar{l}\|_{X(l^n)}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}. \] Thus we showed that \[ \int I_4\,dx{}\leq{}Ch\|l^n-\bar{l}\|_{X(l^n)}\|l^{n-1}-\bar{l}\|^2_{X(l^{n-1})}. \] It remains to estimate $I_3.$ Notice, that it has structure similar to that of $I_4.$ Once we establish the estimates in lemma \ref{lemma:3}, $\int I_3\,dx$ is estimated in a similar way, leading to \[ \int I_3\,dx{}\leq{} Ch\|l^n-\bar{l}\|_{X(l^n)}\left( \|l^{n-1}-\bar{l}\|_{X(l^n-1)}{}+{} \|l^{n-1}-\bar{l}\|^2_{X(l^n-1)}{}+{} \|l^{n-1}-\bar{l}\|^3_{X(l^n-1)}\right), \] which completes the proof of lemma \ref{lemma:energy:3}. \end{proof} Collecting all energy estimate we conclude the next lemma. \begin{lemma} \label{lemma:Final} Assume (without loss of generality) that for all $n,$ $\|l^n-\bar{l}\|_{X(l^n)}\leq 1.$ There is $C>0$ independent of $(n,h)$ such that for all $n=0..N,$ \begin{equation} \label{est:Final} \|l^n-\bar{l}\|^2_{X(l^n)}{}\leq{}\|l^0-\bar{l}\|^2_{X(l^0)}{}+{}CT. \end{equation} \end{lemma} Now we impose a smallness condition on $T:$ \[ CT< \Delta_2-\Delta_1, \] (see lemma \ref{lemma:nondegenerate}), which implies that for all $n, $ and $x,$ $l^n(x,\cdot)\in P(R,\delta_1,r,\delta_2),$ verifying Hypothesis 1. \subsubsection{Convergence of the scheme} In this section we show that properly interpolated on time axis solution of the discrete scheme converges to a classical solution of \eqref{model:1}. Let $\{l^n(x,v)\}$ be the sequence verifying estimate \eqref{est:Final}. Define a continuous function $l^h(x,t,v)$ by \begin{equation} \label{approx:continuous:1} l^h(x,t,v){}={}\left\{ \begin{array}{ll} l^n(x-(t-nh)2v,v), & t\in[nh,nh+h/2],\\\\ \left(1{}-{}2\dfrac{(n+1)h-t}{h}\right)l^{n+1}(x,v){}+{} 2\dfrac{(n+1)h-t}{h}l^n(x-hv,v),& t\in[nh+h/2,(n+1)h]. \end{array} \right. \end{equation} Notice that $\partial_t l^h$ is piecewise constant in $t.$ By construction, for all $(x,v)$ and all $t\not=nh,$ $l^h$ is a solution of the equation \begin{equation} \label{eq:approx} \partial_t l^h{}={}-2v\cdot\nabla{}_xl^h\mathbbm{1}_{A_h}(t){}+{} \frac{2}{h}(l^{\lfloor t/h\rfloor+1}-l^{\lfloor t/h\rfloor})\mathbbm{1}_{[0,T]\setminus A_h}(t), \end{equation} where \[ A_h{}={}\cup_{n=0}^{N-1}[nh,nh+h/2]. \] The following bounds are easily verified given \eqref{est:Final}. \begin{lemma}[Bounds] \label{lemma:bounds} The following statements hold. \[ l^h-\bar{l}\quad\mbox{\rm bounded in } L^\infty((0,T)\times B_R; H^3(\mathbb{R}^3)); \] \[ D_{x,t,v}l^h\quad\mbox{\rm bounded in } L^\infty(\mathbb{R}^3\times(0,T)\times B_R); \] \[ \forall (x,t),\quad \mbox{\rm supp}\; _vl^h(x,t,\cdot)\subset B_R. \] \end{lemma} \begin{lemma}[Compactness] \label{lemma:compactness} There a sequence (still labeled) $h\to0,$ and a continuous function $l(x,t,v),$ such \[ l-\bar{l}\in L^\infty((0,T)\times B_R;H^3(\mathbb{R}^3)); \] \[ D_{x,t,v}l\in L^\infty(\mathbb{R}^3\times(0,T)\times B_R); \] \[ \mbox{ $l^h$ converges to $l$ uniformly on compact sets of $\mathbb{R}^3\times[0,T]\times\mathbb{R}^3;$} \] for any $(x,t),$ \[ \mbox{\rm supp}\; _vl(x,t,\cdot)\subset B_R, \] and \[ l(x,t,\cdot)\in E^*. \] \end{lemma} \begin{proof} Given the bounds of lemma \ref{lemma:bounds}, it remains to prove the inclusion $l(x,t,\cdot)\in E^*.$ Fix $t\in(0,T].$ Let integer $n_h$ be such that \[ n_hh\leq t<n_hh+h. \] Using the definition of function $l^h$ we find that for $t\in[n_hh,n_hh+h/2],$ \[ |l^{n_h}(x,v)-l^{h}(x,t,v)|{}\leq{}|l^{n_h}(x,v)-l^{n_h}(x-(t-n_hh)2v,v)|, \] and for $t\in[n_hh+h/2,n_hh+h),$ \begin{multline*} |l^{n_h}(x,v)-l^{h}(x,t,v)|{}\leq{}\left(1-2\frac{(n_h+1)h-t}{h}\right)|l^{n_h+1}(x,v)-l^{n_h}(x,v)|\\{}+{} 2\frac{(n_h+1)h-t}{h}|l^{n_h}(x-hv,v)-l^{n_h}(x,v)|. \end{multline*} Using \eqref{est:1}, \eqref{est:2}, and \eqref{est:Final}, we find that \[ \sup_x\int_{B_R}|l^{n_h}(x,v)-l^h(x,t,v)|^2\,dv\to0. \] It follows that for fixed $(x,t),$ $l(x,t,\cdot){}={}\lim l^h(x,t,\cdot)$ is also a limiting point of a sequence $\{l^{n_h}(x,\cdot)\}\subsetE^*.$ Since $E^*$ is closed, $ l(x,t,\cdot)\in E^*. $ \end{proof} We conclude the analysis by taking the limit in the equation \eqref{eq:approx}. By $W(l)$ we denote the anti-derivative of $w,$ normalized by $W(0){}={}0.$ \begin{lemma}[Limiting equations] For any $\psi\in C^\infty_0(\mathbb{R}^4),$ -- test function, and for any $i=1..k$ it holds: \begin{equation} \label{eq:weak} \iiint W(l)l_i(v)\{\partial_t\psi{}+{}v\cdot\nabla{}_x\psi\}\,dvdxdt{}={}0; \end{equation} for a.e. $(x,t):$ \begin{equation} \label{eq:strong} \int\left\{\partial_tW(l){}+{}v\cdot\nabla{}_x W(l)\right\}l_i(v)\,dv{}={}0; \end{equation} for a.e. $(x,t):$ \begin{equation} \label{eq:strong1} \int\left\{\partial_t l{}+{}v\cdot\nabla{}_x l\right\}l_i(v)\,w(l)dv{}={}0; \end{equation} vector \[ U_i(x,t){}={}\int l_i(v)W(l)\,dv,\quad i=1..k, \] is a classical solution of the system of PDEs \eqref{eq:Hyperbolic:2} corresponding to the closure \eqref{model:1}. \end{lemma} \begin{proof} Only the first statement needs a proof. Others follow from that, since $l$ is Lipschitz continuous in $(x,t,v).$ We multiply equation \eqref{eq:approx} by $w(l^h)\psi(x,t)l_i(v),$ and consider the integral \[ 2\iiint \mathbbm{1}_{A_h}(t)l_i(v)W(l^h)v\cdot\nabla{}_x\psi\,dvdxdt. \] Due to the uniform convergence of $l^h,$ the integral converges to \[ \iiint l_i(v)W(l)v\cdot\nabla{}_x\psi\,dvdxdt. \] Consider the integral \begin{multline*} \frac{2}{h}\int_{[0,T]\setminus A_h}\iint l_i(v)w(l^h)(l^{\lfloor t/h\rfloor+1}(x,v)-l^{\lfloor t/h\rfloor}(x-hv,v))\psi\,dvdxdt \\ {}={}\frac{2}{h}\int_{[0,T]\setminus A_h}\iint l_i(v)w(l^{\lfloor t/h\rfloor})(l^{\lfloor t/h\rfloor+1}(x,v)-l^{\lfloor t/h\rfloor}(x,v))\psi\,dvdxdt\\ {}+{} \frac{2}{h}\int_{[0,T]\setminus A_h}\iint l_i(v)(w(l^h)-w(l^{\lfloor t/h\rfloor}))(l^{\lfloor t/h\rfloor+1}(x,v)-l^{\lfloor t/h\rfloor}(x,v))\psi\,dvdxdt\\ + \frac{2}{h}\int_{[0,T]\setminus A_h}\iint l_i(v)(w(l^h)-w(l^{\lfloor t/h\rfloor}))(l^{\lfloor t/h\rfloor}(x-hv,v)-l^{\lfloor t/h\rfloor}(x,v))\psi\,dvdxdt. \end{multline*} The first integral on the right equals zero by the definition of $l^{\lfloor t/h\rfloor+1}.$ The other two are of the order $h,$ due to estimates \eqref{est:1}, \eqref{est:2}, and \eqref{est:Final}. \end{proof} Finally, since the classical solutions of the problem in question are unique, we conclude that the original sequence $l^h$ converges to $l$ as $h\to0.$ \end{section} \begin{section}{BGK-type equation} In section we consider a problem of projecting a BGK-type model \eqref{eq:Misha} onto a finite dimensional space $E^*.$ Let $E^*_0=\span\{1,v,|v|^2\},$ and $E^*{}={}\span\{ l_i(v),\,i=1..k\},$ with $E^*_0\subsetE^*.$ Let $\Pi^0_l$ denote the projection of $l$ onto $E^*_0$ in $L^2$ space with the weight $w=w(l).$ We consider the problem \begin{equation} \label{model:2} \begin{cases} \partial_t l {}+{}v\cdot\nabla{}_x l{}-{}(\Pi^0_l-l){}\in{} (E^*(l))^\perp, & (x,t,v)\in \mathbb{R}^3\times(0,T)\times\mathbb{R}^3,\\ l(x,t,\cdot)\in E^*, & (x,t)\in\mathbb{R}^4_+, \\ l(x,0,v){}={}l^0(x,v), & (x,v)\in\mathbb{R}^6. \end{cases} \end{equation} Let $\bar{l}\in E^*_0,$ and $l^0=l^0(x,\cdot)\inE^*,$ be the functions verifying assumptions \eqref{cond:initial} of the previous section. Given $h\in(0,1],$ and the values of approximation $l^{n-1}(x,t,v),$ define approximation at the next level, $l^n$ as \begin{eqnarray} \hat{l}^{n-1}(x,v)&=&l^{n-1}(x-hv,v),\notag \\ \label{Projection:2} \check{l}^{n-1}(x,v)&=&\hat{l}^{n-1}(x,v){}+{}h(\Pi^0_{\hat{l}^{n-1}}{}-{}\hat{l}^{n-1})\\ &&=(1-h)\hat{l}^{n-1}(x,v){}+{}h\Pi^0_{\hat{l}^{n-1}},\quad \Pi^0_{\hat{l}^{n-1}}{}={}\mbox{\rm proj}_{E_0(l^{n-1})}\hat{l}^{n-1},\notag \\ l^n(x,v) &=&\mbox{\rm proj}_{E(l^{n-1})}\check{l}^{n-1}. \notag \end{eqnarray} We prove \begin{theorem} \label{theorem:2} Let $\bar{l},l^0$ be as described above and $h\in(0,1].$ There is time $T>0,$ independent of $h,$ such that all functions $l^n,$ $n=0..\lceil T/h\rceil,$ in \eqref{Projection:2} are well defined, and the family $\{l^h\}$ of interpolations of $l^n$'s, defined in \eqref{inter:1}--\eqref{inter:3}, converges as $h\to0$, uniformly on compact set in $\mathbb{R}^3\times[0,T]\times\mathbb{R}^3$ to a unique classical solution of \eqref{model:2}. \end{theorem} \begin{proof} We start with the heuristic arguments of the proof. It is based on the orthogonality of $\check{l}^{n-1}-l^n$ to $E^*,$ and orthogonality of $\hat{l}^{n-1}-\Pi^0_{\hat{l}^{n-1}}$ to $E^*_0.$ Consider the following orthogonal decompositions (w.r.t. weight $w^{n-1}{}={}w(l^{n-1})$): \begin{eqnarray} \label{orth:1}\check{l}^{n-1}-\bar{l} &=& (l^n-\bar{l}){}+{}(\check{l}^{n-1}-l^n),\\ \label{orth:2}\hat{l}^{n-1}-\bar{l} &=& (\Pi^0_{\hat{l}^{n-1}}-\bar{l}){}+{}(\hat{l}^{n-1}{}-{}\Pi^0_{\hat{l}^{n-1}}). \end{eqnarray} From the first equation we obtain \[ |l^n{}-{}\bar{l}|^2_{w^{n-1}}{}\leq{}|\check{l}^{n-1}-\bar{l}|^2_{w^{n-1}}{}\leq{}(1-h)|\hat{l}^{n-1}-\bar{l}|^2_{w^{n-1}}{}+{}h|\Pi^0_{\hat{l}^{n-1}}-\bar{l}|^2_{w^{n-1}}. \] Here and below $h\in(0,1).$ From the second equation we get \[ |\Pi^0_{\hat{l}^{n-1}}-\bar{l}|^2_{w^{n-1}}\leq |\hat{l}^{n-1}-\bar{l}|^2_{w^{n-1}}, \] which leads to \[ |l^n{}-{}\bar{l}|^2_{w^{n-1}}{}\leq{}|\hat{l}^{n-1}-\bar{l}|^2_{w^{n-1}}. \] It implies, as explained in theorem \ref{theorem:1}, \[ \int |l^n{}-{}\bar{l}|^2_{w^{n}}\,dx{}\leq{}\int |l^{n-1}-\bar{l}|^2_{w^{n-1}}\,dx{}+{}O(h), \] where $O(h)$ measures the change in weights. Differentiating equations \eqref{orth:1} and \eqref{orth:2} in $x,$ we get: \begin{eqnarray*} \label{orth:3}D^\alpha_x(\check{l}^{n-1}-\bar{l}) &=& D^\alpha_x(l^n-\bar{l}){}+{}D^\alpha_x(\check{l}^{n-1}-l^n),\\ \label{orth:4}D^\alpha_x(\hat{l}^{n-1}-\bar{l}) &=& D^\alpha_x(\Pi^0_{\hat{l}^{n-1}}-\bar{l}){}+{}D^\alpha_x(\hat{l}^{n-1}{}-{}\Pi^0_{\hat{l}^{n-1}}). \end{eqnarray*} It follows that \[ |D^\alpha_x(l^n-\bar{l})|^2_{w^{n-1}}{}\leq{}|D^\alpha_x(\check{l}^{n-1}-\bar{l}|^2_{w^{n-1}}{}+{}R_n, \] and \[ |D^\alpha_x(\Pi^0_{\hat{l}^{n-1}}-\bar{l})|^2_{w^{n-1}}{}\leq{}|D^\alpha_x(\hat{l}^{n-1}-\bar{l})|^2_{w^{n-1}}{}+{}\tilde{R}_n, \] where $\int R_n\,dx,\,\int \tilde{R}_n\,dx{}={}O(h).$ By varying the weights we arrive at \[ \int|D^\alpha_x(l^n-\bar{l})|^2_{w^n}\,dx{}={}\int|D^\alpha_x(l^{n-1}-\bar{l}|^2_{w^{n-1}}\,dx{}+{}O(h). \] By appropriately changing the arguments of lemmas \ref{lemma:1}--\ref{lemma:Final} we obtain \begin{lemma} There are $T,C>0$ such that for all $n=0..\lceil T/h\rceil,$ \[ \|l^n-\bar{l}\|_{X(l^n)}{}\leq C. \] For all $x\in\mathbb{R}^3,$ $l^n(x,\cdot)\in P(R,\delta_1,r,\delta_2).$ \end{lemma} We omit the proof of this lemma. A continuous process $l^h(x,t,v),$ $t\in[0,T]$ is constructed by a linear interpolation between states $l^n, \hat{l}^n, \check{l}^n,$ and $l^{n+1}$ on interval $[nh,(n+1)h]:$ \\ for $t\in[nh, (n+1/3)h],$ \begin{equation} \label{inter:1} l^h(x,t,v){}={}l^n(x-(t-nh)3v,v); \end{equation} for $t\in[(n+1/3)h,(n+2/3)h],$ \begin{equation} \label{inter:2} l^h(x,t,v){}={}\left( 1-3\frac{t-(n+1/3)h}{h}\right)l^n(x-hv,v){}+{}3\frac{t-(n+1/3)h}{h}\check{l}^n(x,v); \end{equation} for $t\in[(n+2/3)h,(n+1)h],$ \begin{equation} \label{inter:3} l^h(x,t,v){}={}\left( 1-3\frac{t-(n+2/3)h}{h}\right)\check{l}^n(x,v){}+{}3\frac{t-(n+2/3)h}{h}l^{n+1}(x,v). \end{equation} Repeating the convergence arguments of theorem \ref{theorem:1} we find that $l^h$ converges (uniformly on compact sets of $\mathbb{R}^3\times[0,T]\times\mathbb{R}^3$ to a Lipschitz continuous function $l,$ which is a classical solution of \eqref{model:2}. \end{proof} \end{section} \begin{section}{Appendix} \subsubsection{Lemmas from the theory of Sobolev's spaces} \begin{lemma} \label{Sobolev:1} Let $f$ be a measurable function such that for some constant $\bar{f},$ $f-\bar{f}\in H^3(\mathbb{R}^3).$ There is $C>0,$ independent of $f,$ and $\bar{f},$ such that \[ \mbox{\rm ess\,sup}_x|\nabla{} f|\leq{}C\|f-\bar{f}\|_{H^3(\mathbb{R})^3}. \] \end{lemma} \begin{lemma} \label{Sobolev:2} Let $f\in H^1(\mathbb{R}^3).$ Then, for all $h\in\mathbb{R}^3,$ \[ \int |f(x+h)-f(x)|^2\,dx{}\leq{}|h|^2\int|\nabla{} f|^2\,dx. \] \end{lemma} \subsubsection{Equivalence of norms in $v$--variable} By assumption on the linear independence of polynomials $l_i(v)$ spanning $E^*,$ restriction of $E^*$ to any ball is a $k+1$--dimensional vector space. Any two norms on $E^*$ are equivalent to $\|l\|{}={}(\sum_i \gamma_i^2)^{1/2},$ for $l=\sum_i \gamma_il_i(v).$ In particular, if $B_R$ and $B_r$ are two balls, norms \[ \|l\|_{L^p(B_R)},\,\|l\|_{L^p(B_r)},\quad p\in[1,+\infty], \] are equivalent. This property implies the following lemma. \begin{lemma} Consider set \[ \bar{E}^*{}={}\left\{ \sum \gamma_i(x)l_i(v)\,:\, \gamma_i\in H^m(\mathbb{R}^3),\,i=1..k\right\}, \] for some $m\geq 0.$ For any $p,q\in[1,+\infty],$ norms \[ \|l\|_{L^p(B_R;H^m(\mathbb{R}^3))},\, \|l\|_{L^q(B_R;H^m(\mathbb{R}^3))}, \] on $\bar{E}^*,$ are equivalent. For any $p\in[1,+\infty],$ norms \[ \|l\|_{L^p(B_R;H^m(\mathbb{R}^3))},\, \|l\|_{L^p(B_r;H^m(\mathbb{R}^3))}, \] on $\bar{E}^*,$ are equivalent. \end{lemma} \end{section}
bc95a429089ebc43af16794150486791ed2cc8e3
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Suppose that $\ensuremath{\textup{\textsf{F}}} = \ensuremath{\textup{\textsf{F}}}_q$ is a finite field with size~$q$. Let $m$ and~$n$ be integers such that $0 \le m \le n$. The paper that introduced Wiedemann's algorithm~\cite{wied86} also includes a proof of the following claim --- which concerns an $n \times n$ matrix obtained by appending an additional set of row vectors to a matrix $A \in \matgrp{\ensuremath{\textup{\textsf{F}}}}{m}{n}$ with maximal rank:\footnote{Wiedemann attributes much of the proof of this claim to an anonymous referee who is thanked for allowing this work to be included.} \begin{quote} \textbf{Theorem~$\text{\textbf{1}}'$ [Wiedemann]}: \emph{Numbers~$\epsilon > 0$ and~$c_1$ exist, both independent of~$q$, with the following property: For any integers $n > m \ge 0$ a random procedure exists for generating $n-m$ row vectors with length~$n$ such that if $A$ is an $m \times n$ matrix of rank~$m$, then with probability at least~$\epsilon$, the resulting $n \times n$ matrix is nonsingular and the total Hamming weight of the generated rows is at most $1 + c_1 n \log n$.} \end{quote} Unfortunately the unknown constants~$\epsilon$ and~$c_1$ are neither supplied nor estimated. Furthermore, it seems that if the proof in~\cite{wied86} is applied without change in order to determine these values then either~$c_1$ must be so large or $\epsilon$ so tiny that the result is of limited practical interest. This is, somewhat, rectified in Section~\ref{sec:modified_proof}: While the outline of Wiedemann's argument is maintained, along with the details of several steps, several other components are revised or replaced entirely in order to remove unnecessary bounds on various parameters and to simplify the estimation of the unknown parameters~$\epsilon$ and~$c_1$. The bound~$\epsilon$ is also increased by adding another $\ell$ rows to the resulting matrix~$B \in \matgrp{\ensuremath{\textup{\textsf{F}}}}{(n+\ell)}{n}$ rows; here, $\ell$ depends on the size of the field~$\ensuremath{\textup{\textsf{F}}}$. In particular, the following result is obtained. \begin{theorem} \label{thm:constants_for_conditioner} Let $\ensuremath{\textup{\textsf{F}}} = \ensuremath{\textup{\textsf{F}}}_q$ be the finite field with size~$q$. Let $m$ and~$n$ be integers such that $n \ge m \ge 0$. Let $\ell$ be a nonnegative integer and let $\sigma$, $\tau$ and $\upsilon$ be positive constants (depending on~$q$, but independent of~$n$ and~$m$) as given in Table~\ref{fig:summary_table} on page~\pageref{fig:summary_table}. A random procedure exists for generating $n - m + \ell$ rows with length~$n$ such that if $A$ is an $m \times n$ matrix of rank~$m$, then an additional $m + \ell$ rows (each with length~$n$) are produced, and the expected number of nonzero entries in these rows is $\sigma n \ln n + \tau n$ if $n - m \ge \upsilon$, and at most $\frac{q-1}{q} (n-m+\ell) n \le \frac{q-1}{q} (\upsilon + \ell) n$, otherwise. If $q \le n^2$ then the matrix $B \in \ensuremath{\textup{\textsf{F}}}^{(n+\ell) \times n}$ obtained from~$A$ by adding these rows has maximal rank~$n$ with probability at least~$\frac{9}{10}$. If $q > n^2$ then this matrix has maximal rank with probability at least $\frac{9}{10} - \frac{9}{10n}$. \end{theorem} The probability bounds listed above are quite arbitrary. The parameter~$\sigma$ does not depend on this probability. Formulas for~$\ell$, $\tau$ and~$\upsilon$, depending on the field size~$q$ and an arbitrarily small failure probability~$\epsilon$, are given in Section~\ref{sec:modified_proof}. \begin{figure}[t!] \begin{center} \begin{tabular}{ccccc|ccccc} $q$ & $\ell$ & $\sigma$ & $\tau$ & $\upsilon$ & $q$ & $\ell$ & $\sigma$ & $\tau$ & $\upsilon$ \\[2pt] \hline $2$\rule{0pt}{12pt} & $8$ & $\frac{43}{2}$ & $17$ & $41$ & $13$ & $3$ & $\frac{120}{13}$ & $6$ & $150$ \\ $3$\rule{0pt}{12pt} & $5$ & $16$ & $11$ & $55$ & $16$--$19$ & $2$ & $\frac{9(q-1)}{q}$ & $5$ & $194$ \\ $4$\rule{0pt}{12pt}& $4$ & $\frac{225}{16}$ & $9$ & $65$ & $23$--$29$ & $2$ & $\frac{8(q-1)}{q}$ & $4$ & $285$ \\ $5$\rule{0pt}{12pt} & $4$ & $\frac{64}{5}$ & $8$ & $75$ & $31$--$43$ & $2$ & $\frac{15(q-1)}{2q}$ & $4$ & $381$ \\ $7$\rule{0pt}{12pt} & $3$ & $\frac{78}{7}$ & $7$ & $96$ & $47$---$59$ & $2$ & $\frac{7(q-1)}{q}$ & $4$ & $577$ \\ $8$\rule{0pt}{12pt} & $3$ & $\frac{21}{2}$ & $6$ & $108$ & $61$--$71$ & $2$ & $\frac{27(q-1)}{4q}$ & $4$ & $783$ \\ $9$\rule{0pt}{12pt} & $3$ & $\frac{88}{9}$ & $6$ & $124$ & $73$--$83$ & $2$ & $\frac{33(q-1)}{5q} n$ & $4$ & $996$ \\ $11$\rule{0pt}{12pt} & $3$ & $\frac{105}{11}$ & $6$ & $136$ & $\ge 89$ & $2$ & $\frac{13(q-1)}{2q}$ & $4$ & $1213$ \end{tabular} \end{center} \caption{Bounds Established in Section~\ref{sec:modified_proof}} \label{fig:summary_table} \end{figure} A second result, which is also proved in Section~\ref{sec:modified_proof}, establishes that the constant~$\sigma$, mentioned above, can be made arbitrarily close to~$6 \left(1 - \frac{1}{q}\right)$, provided that the minimum field size~$q$ and the constant~$\upsilon$ are both increased --- at the cost of increasing the constants~$\ell$, $\tau$ and $\upsilon$ (but not~$\sigma$) that are listed. \begin{theorem} \label{thm:second_constants_for_conditioner} Let $N$ be an integer such that $N \ge 18$. Let $\ensuremath{\textup{\textsf{F}}}_q$ be a finite field with size $q \ge 16N+9$ Let $m$ and~$n$ be integers such that $n \ge m \ge 0$. Let $\sigma = \left(1 - \frac{1}{q}\right)\ \cdot \left (6 + \frac{3}{N}\right)$, $\tau = 1$, and $\upsilon = \lceil (2N+1) \ln (2N+1) + \frac{167}{5} (2N+1) \rceil$. A random procedure exists for generating $n-m$ rows with length~$n$ such that if $A$ is an $m \times n$ matrix of rank~$m$, then an additional $m$ rows (each with length~$n$) are produced, and the expected number of nonzero entries in this row is $\sigma n \ln n + \tau n$ if $n - m \ge \upsilon$, and at most $\frac{q-1}{q} (n - m) n \le \frac{q-1}{q} \upsilon n$, otherwise. If $q \le n^2$ then the matrix $B \in \matring{\ensuremath{\textup{\textsf{F}}}}{n}$ obtained from~$A$ by adding these rows is nonsingular with probability at least $\frac{8}{9}$. If $q > n^2$ then this matrix is nonsingular with probability at least \mbox{$\frac{8}{9} - \frac{8}{9n}$}. \end{theorem} Once again, the probability bounds here are quite arbitrary, and probability bounds that are closer to one can be obtained by applications of the same techniques, at the cost of increasing the values of the constants~$\tau$ and~$\upsilon$. This is work in progress. Future versions of this report will document progress in establishing that this yields an efficient matrix preconditioner, to bound the number of nontrivial nilpotent blocks of a conditioned matrix without lowering matrix rank, for matrices over small finite fields. \section{A Modified Proof of Wiedemann's Result} \label{sec:modified_proof} This section describes modifications to Wiedemann's argument needed to establish Theorems~\ref{thm:constants_for_conditioner} and~\ref{thm:second_constants_for_conditioner}. \subsection{Getting Started --- and Improving Reliability by Adding Rows} \label{ssec:getting_started} Suppose that $m$ and~$n$ are positive integers such that $0 \le m \le n$ and $A \in \matgrp{\ensuremath{\textup{\textsf{F}}}}{m}{n}$ is a matrix with maximal rank~$m$, where $\ensuremath{\textup{\textsf{F}}} = \ensuremath{\textup{\textsf{F}}}_q$ is a finite field with size~$q$. Following Wiedemann's argument, let us begin by assuming that $q \le n^2$ and suppose, as well, that \begin{equation} \label{eq:definition_of_k_and_c2} n = m + k + c_2 \end{equation} where $c_2$ will be defined later. Suppose first that $k \le c_3 \ln n$, where $c_3$ is another constant to be chosen later, and that the remaining $n-m+\ell$ rows of an $(n + \ell) \times n$ matrix~$B$ are chosen uniformly and independently from~$\matgrp{\ensuremath{\textup{\textsf{F}}}}{1}{n}$. The following lemma, which is easily proved, bounds the probability that the rank of~$B$ is less than~$n$ in this case. \begin{lemma} \label{lem:dense_selection_of_vectors} Let $\ell$ be a nonnegative integer and let $B \in \matgrp{\ensuremath{\textup{\textsf{F}}}}{(n + \ell)}{n}$ be a matrix produced by appending another $n-m+\ell$ rows, selected uniformly and independently from~$\matgrp{\ensuremath{\textup{\textsf{F}}}}{1}{n}$, to~$A$. Then the probability that the rank of~$B$ is less than~$n$ is at most $q^{-\ell}$. Furthermore, if $q \ge 3$ then the top $n \times n$ submatrix of~$B$ is nonsingular with probability at least $\frac{1}{q-1}$. \end{lemma} \begin{proof} Permuting the columns of~$A$ (and~$B$) as needed we may assume without loss of generality that the principal $m \times m$ submatrix of~$A$ is nonsingular. Let us continue by choosing the entries in the leftmost $m$~columns of the $n-m+\ell$ rows that are to be appended to~$A$. Regardless of the choice of these entries, this completes an $(n+\ell) \times m$ submatrix of~$B$, including the leftmost $m$ columns, that must also have maximal rank~$m$. The remaining entries of the top~$m$ rows of~$B$ are, of course, fixed: They are entries of~$A$. The entries in the lower columns may now be chosen freely and (viewing the selection of these entries in column order, instead of row order) a standard argument establishes that, for $1 \le i \le n-m$, if the leftmost $m+i-1$ columns of~$B$ are linearly independent then the probability that the $m+i^{\text{th}}$ column of~$B$ is a linear combination of these columns is at most $q^{m+i-n-\ell-1}$. It now follows that $B$ has rank less than~$n$ with probability at most \[ \sum_{i=1}^{n-m} q^{m+i-n-\ell-1} < \sum_{j \ge 0} q^{-(\ell+1)-j} = \frac{q^{-(\ell+1)}}{1-q^{-1}} \le q^{-\ell}. \] Finally, if $q \ge 3$ then, setting $\ell = 0$, one can see the (top) $n \times n$ submatrix obtained by appending these rows is singular with probability at most $\frac{q^{-1}}{1 - q^{-1}} = \frac{1}{q-1}$, as claimed. \end{proof} Since $n- m = k + c_2$, the expected number of nonzero entries in these rows is \[ \left( \frac{q-1}{q} \right) (k + c_2 + \ell)n < \left( \frac{q-1}{q} \right) \cdot \left( c_3 n \ln n + (c_2 + \ell) n \right) \] when $k \le c_3 \ln n$. With that noted let us suppose, instead, that $k > c_3 \ln n$. Following Wiedemann once again, let \begin{equation} \label{eq:definition_of_z} z = 1 - \frac{c_3 \ln n}{k} > 0. \end{equation} Suppose that, for the initial $k$~rows, each entry is set to zero with probability~$z$. The remaining unset entries are then chosen uniformly and independently from~$\ensuremath{\textup{\textsf{F}}}$. The expected number of nonzero entries in these rows is then \[ \left( \frac{q-1}{q} \right) c_3 n \ln n. \] If the entries of another $c_2 + \ell$ rows are chosen uniformly and independently from~$\ensuremath{\textup{\textsf{F}}}$ then the expected number of nonzero entries in these rows is \[ \left( \frac{q-1}{q} \right) (c_2 + \ell) n. \] Consequently the expected number of nonzero entries in all these rows is less than \[ \left( \frac{q-1}{q} \right) \cdot \left( c_3 n \ln n + (c_2 + \ell)n \right) \] in this case as well. The bulk of the rest of Wiedemann's argument concerns the derivation of an upper bound for the probability that $B$ has rank less than~$n$ if $k > c_3 \ln n$ and the rows of~$B$ are chosen in this way. Following Wiedemann, let $\rho$ be the probability that the rows of~$A$, together with the first $k$ (sparse) rows generated using the above process, are linearly dependent --- that is, the probability that the space spanned by these vectors has dimension less than $m+k$. Wiedemann shows that \begin{equation} \label{eq:bound_for_rho} \rho \le \rho_0 + \rho_1 \end{equation} where \begin{equation} \label{eq:definition_of_rho0} \rho_0= \sum_{1 \le j \le k\beta_0} \binom{k}{j} (q-1)^j \left( q^{-1} + \frac{(q-1)}{q} (nq)^{-c_4 \beta} \right)^{n-m}, \end{equation} and \begin{equation} \label{eq:definition_of_rho_1} \rho_1 = \sum_{k \beta_0 < j \le k} \binom{k}{j} (q-1)^j \left( q^{-1} + \frac{(q-1)}{q} (nq)^{-c_4 \beta} \right)^{n-m}, \end{equation} when \begin{equation} \label{eq:definition_of_c4} c_4 = \frac{c_3}{3}, \end{equation} and when \begin{equation} \label{eq:definition_of_beta} \beta = \frac{j}{k} \end{equation} in the above expressions, and where $\beta_0$ is yet another constant to be defined later. A useful bound for $\rho_1$ is next obtained: Assuming that \begin{equation} \label{eq:constraints_on_c2_and_c4} c_2 \ge \frac{3}{2} \ge \log_2 e \qquad \text{and} \qquad c_4 \ge \frac{2}{\beta_0}, \end{equation} Wiedemann establishes that \begin{equation} \label{eq:bound_for_rho1} \rho_1 \le 2 q^{-c_2}. \end{equation} Wiedemann continues by observing that \begin{equation} \label{eq:first_bound_for_rho0} \rho_0 \le \sum_{1 \le j < k \beta_0} \left( \frac{1}{\beta^{\beta} (1-\beta)^{1-\beta}} f(q) \right)^k, \end{equation} where \begin{equation} \label{eq:definition_of_f} f(x) = (x-1)^{\beta} \left( x^{-1} + (1-x^{-1}) (nx)^{-c_4 \beta} \right). \end{equation} \subsection{Getting to the Next Step by a Different Route} \label{ssec:next_step} Wiedemann continues by using the above to establish that \begin{equation} \label{eq:big_goal} \rho_0 \le \sum_{1 \le j < k \beta_0} \beta^{k \beta}. \end{equation} Unfortunately, Wiedemann's involves a Taylor series approximation that seems only to be accurate for a limited range of values, and might suggest that either $\beta$ must be tiny or $c_4$ must be huge in order for it to be applicable. The argument that follows is, therefore, quite different from given by Wiedemann~\cite{wied86}. With that noted, consider the equation at lines~\eqref{eq:first_bound_for_rho0} and~\eqref{eq:definition_of_f} once again. These imply that \begin{equation} \label{eq:what_is_known_1} \begin{split} \rho_0 &\le \sum_{1 \le j < k \beta_0} \left(\frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{q} + \frac{q-1}{q} (nq)^{-c_4 \beta} \right) \right)^k \\ &\le \sum_{1 \le j < k \beta_0} \left(\frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{q} + \frac{q-1}{q} \left( \frac{\beta}{q} \right)^{c_4 \beta} \right)\right)^k \end{split} \end{equation} since $n^{-1} \le k^{-1} \le \frac{j}{k} = \beta$. It follows from this that \begin{equation} \label{eq:what_is_known_2} \rho_0 < \sum_{1 \le j < k \beta_0} \left(\frac{1}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{(q-1)^{\beta}}{q} + \frac{(q-1) q^{(1-c_4)\beta}}{q} \beta^{c_4 \beta} \right) \right)^k. \end{equation} \subsubsection{Bounding Terms in $\boldsymbol{\rho_0}$ When $\boldsymbol{\beta}$ is Extremely Small} \label{ssec:beta_is_tiny} Note that $\displaystyle{\lim_{\beta \rightarrow 0^{+}}} \beta^{\beta} = \displaystyle{\lim_{\beta \rightarrow 0^{+}}} (1-\beta)^{1-\beta} = 1$, and $\beta^{\beta} < 1$ when $0 < \beta < 1$. The bounds given at lines~\eqref{eq:what_is_known_1} and~\eqref{eq:what_is_known_2} can be simplified by establishing a lemma like the following, allowing the factor $(1-\beta)^{-(1-\beta)}$ to be replaced by a factor $\beta^{\gamma \beta}$, for a negative constant~$\gamma$, when $\beta$ is small. \begin{lemma} \label{lem:getting_rid_of_pesky_term} Consider the relationship between $(1-x)^{-(1-x)}$ and $x^{\gamma x}$, for a negative constant~$\gamma$, when $x$ is small and positive. \begin{enumerate} \renewcommand{\labelenumi}{\text{\textup{(\alph{enumi})}}} \setlength{\itemsep}{0pt} \item If $0 < x \le \frac{5}{43}$ then $(1-x)^{-(1-x)} \le x^{\gamma x}$, with $\gamma = -\frac{11}{25}$. \item If $0 < x \le \frac{1}{5}$ then $(1-x)^{-(1-x)} \le x^{\gamma x}$, with $\gamma = -\frac{23}{40}$. \end{enumerate} \end{lemma} \begin{proof} Consider the function \[ g(x) = \gamma x \ln x + (1-x) \ln (1-x) \] when $\gamma$ is a negative constant and $0 < x < 1$. Since $e^{g(x)} = \frac{x^{\gamma x}}{(1-x)^{-(1-x)}}$ and $(1-x)^{-(1-x)} > 0$, $(1-x)^{-(1-x)} \le x^{\gamma x}$ (for $0 < x < 1$) if and only if $g(x) \ge 0$. Now \begin{align*} \lim_{x \rightarrow 0^{+}} g(x) &= \lim_{x \rightarrow 0^{+}} \frac{\gamma \ln x}{x^{-1}} + \lim_{x \rightarrow 0^{+}} (1-x) \ln (1-x) \\ &= \lim_{x \rightarrow 0^{+}} \frac{\gamma \ln x}{x^{-1}} \\ &= \lim_{x \rightarrow 0^{+}} \frac{\gamma x^{-1}}{-x^{-2}} \tag{by l'H\^{o}pital's Rule} \\ &= \lim_{x \rightarrow 0^{+}} -\gamma x = 0. \end{align*} It is easily checked that $g'(x) = \gamma \ln x - \ln (1-x) + \gamma - 1$. Thus $\displaystyle{\lim_{x \rightarrow{0^{+}}}} g'(x) = +\infty$, so that $g'(x) > 0$ when $x$ is small and positive. It follows from the above that $g(x) > 0$ when $x$ is small and positive, as well. Note next that $g''(x)= \frac{\gamma}{x} + \frac{1}{1-x}$, so that $\displaystyle{\lim_{x \rightarrow 0^{+}}} g''(x) = -\infty$, since $\gamma < 0$, and $g''(x) < 0$ when $x$ is small and positive. On the other hand, $g'''(x) = -\frac{\gamma}{x^2} + \frac{1}{(1-x)^2} > 0$ when $0 < x < 1$. It now follows that $g(x) \ge 0$ for $0 < x \le \delta$ if $g(\delta) \ge 0$ and $g'(\delta) < 0$: Since $g'''(x) > 0$ for all $x$ such that $0 < x < 1$, this implies that either \begin{enumerate} \renewcommand{\labelenumi}{\text{\textup{\roman{enumi}.}}} \setlength{\itemsep}{0pt} \item $g(x) \ge 0$ for all $x$ such that $0 < x < 1$, \item there exists a value $\Gamma$ such that $0 < \Gamma < 1$, $g'(x) \ge 0$ when $0 < x \le \Gamma$ and $g'(x) < 0$ when $\Gamma < x < 1$, or \item there exist values~$\Gamma_1$ and~$\Gamma_2$ such that $0 < \Gamma_1 < \Gamma_2 < 1$, $g'(x) \ge 0$ when $0 < x \le \Gamma_1$, $g'(x) \le 0$ when $\Gamma_1 \le x \le \Gamma_2$, and $g'(x) > 0$ when $\Gamma_2 < x < 1$. \end{enumerate} The claim is certainly trivial in the first case. In the second case it necessarily follows that $\Gamma < \delta$, while it follows that $\Gamma_1 < \delta < \Gamma_2$ in the third case. In each of the last two cases, the function~$g$ is either nondecreasing or has a local maximum in the interval $0 < x \le \delta$. In either case, it is minimized at one or the other of this interval's endpoints. Since $\displaystyle{\lim_{x \rightarrow 0^{+}}} g(x) = 0$, it therefore suffices to confirm that $g(\delta) \ge 0$ in order to establish that $g(x) \ge 0$ when $0 < x \le \delta$. Part~(a) of the claim can now be established by choosing $\gamma = -\frac{11}{25}$ and $\delta = \frac{5}{43}$; then $g'(\delta) = -\frac{11}{25} \ln \frac{5}{43} - \ln \frac{38}{43} - \frac{36}{25} < -0.3 < 0$ and $g(\delta) = -\frac{11}{215} \ln \frac{5}{43} + \frac{38}{43} \ln \frac{38}{43} > 0.0008 > 0$, as required. Part~(b) of the claim can be established by choosing $\gamma = -\frac{23}{40}$ and $\delta = \frac{1}{5}$; then $g'(\delta) = \frac{23}{40} \ln 5 - \ln \frac{4}{5} - \frac{63}{40} < -0.4 < 0$ and $g(\delta) = \frac{23}{200} \ln 5 + \frac{4}{5} \ln \frac{4}{5} > 0.006 > 0$. \end{proof} Suppose, now, that $\gamma \in \ensuremath{\mathbb{Q}}$ is a negative constant such that $(1 - \beta)^{-(1-\beta)} \le \beta^{\gamma \beta}$ when $0 < \beta \le \Delta$ for a positive constant~$\Delta$; it follows by the above lemma that one might choose $\gamma = -\frac{11}{25}$ if $\Delta = \frac{5}{43}$, and that one might choose $\gamma = -\frac{23}{40}$ if $\Delta = \frac{1}{5}$. It would follow from the inequality at line~\eqref{eq:what_is_known_2} that \begin{equation} \label{eq:what_is_known_3} \rho_0 \le \sum_{1 \le j < k \beta_0} \left(\beta^{(\gamma-1)\beta} \left( \frac{(q-1)^{\beta}}{q} + \frac{(q-1)q^{(1-c_4)\beta}}{q} \beta^{c_4 \beta} \right)\right)^k \end{equation} This can be further simplified by bounding $(q-1)^{\beta}$ by $\beta^{\delta \beta}$ for a small positive constant~$\delta$: \begin{lemma} \label{lem:getting_rid_of_power_of_q} Consider the relationship between $\zeta^x$ and $x^{\delta x}$ when $\zeta$ is a positive constant. If $\zeta = 2$ and $0 < x \le \frac{1}{5}$ then $\zeta^x \le x^{\delta x}$ when $\delta = -\frac{9}{20}$. \end{lemma} \begin{proof} Consider the function \[ h(x) = \delta x \ln x - x \ln \zeta \] when $\delta$ is a negative constant and $\zeta$ is a positive one. Since $e^{h(x)} = \frac{x^{\delta x}}{\zeta^x}$ and $\zeta^x > 0$ when $x > 0$, $\zeta^x \le x^{\delta x}$ (for positive~$x$) if and only if $h(x) \ge 0$. Now note that\\[-33pt] \begin{align*} \lim_{y \rightarrow 0^{+}} h(y) &= \lim_{y \rightarrow 0^{+}} \frac{\delta \ln y}{y^{-1}} - 0\\ &= \lim_{y \rightarrow 0^{+}} \frac{\delta y^{-1}}{-y^{-2}} \tag{by l'H\^{o}pital's Rule} \\ &= \lim_{y \rightarrow 0^{+}} -\delta y = 0. \end{align*} It is easily checked that $h'(x) = \delta \ln x + \delta - \ln \zeta$. Now $\displaystyle{\lim_{y \rightarrow 0^{+}}} h'(y)= +\infty$, since $\delta < 0$, so that $h'(x) > 0$ for sufficiently small positive~$x$. It follows by the above that $h(x) > 0$ when $x$ is small and positive as well. Since $h''(x) = \frac{\delta}{x} < 0$ when $x > 0$ it now follows, for any positive constant~$\rho$, that if $h(\rho) \ge 0$ then $h(x) \ge 0$ as well when $0 < x \le \rho$ --- for it follows from that this that either $h(x) \ge 0$ for all positive~$x$ --- in which case the claim is trivial --- or there exists some constant $\Gamma$ such that $h(x) \ge 0$ when $0 < x \le \Gamma$, and such that $h(x) < 0$ when $x > \Gamma$. If $h(\frac{1}{5}) > \frac{1}{2}$, then $\Gamma$ must be greater than~$\frac{1}{5}$ in this second case. With that noted, the claim can now be established by setting $\zeta=2$, $\delta = -\frac{9}{20}$, $\rho = \frac{1}{5}$, and confirming that $h(\rho) = \frac{9}{100} \ln 5 - \frac{1}{5} \ln 2 > 0.006$. \end{proof} Suppose next that $\delta \in \ensuremath{\mathbb{Q}}$ is constant that is less than or equal to~$0$ such that $(q-1)^{\beta} \le \beta^{\delta \beta}$ when $0 < x \le \Delta$; one can certainly choose $\delta = 0$ if $q = 2$, and it follows by the above lemma that if $q=3$ then one can choose $\delta = -\frac{9}{20}$ when $\Delta = \frac{1}{5}$. It is already necessary for the argument being developed that $c_4 \ge 1$, so that (when $q \ge 2$) $q^{(1-c_4)\beta} \le 1$. It would therefore follow from the bound at line~\eqref{eq:what_is_known_3} that \begin{equation} \label{eq:what_is_known_4} \rho_0 \le \sum_{1 \le j < k \beta_0} \left(\beta^{(\gamma -1)\beta} \left( \frac{\beta^{\delta \beta}}{q} + \frac{(q-1)}{q} \beta^{c_4 \beta}\right)\right)^k. \end{equation} Once again suppose that --- as in Wiedemann's original argument --- we wish to show that \[ \rho_0 \le \sum_{1 \le j < k \beta_0} \beta^{k\beta}. \] Then it follows from the above that it is sufficient to show that \begin{equation} \label{eq:what_is_needed_1} \beta^{(\gamma - 1) \beta} \left( \frac{\beta^{\delta \beta}}{q} + \frac{(q-1)}{q} \beta^{c_4 \beta} \right) \le \beta^{\beta} \qquad \text{when $0 < \beta = \frac{j}{k} \le \beta_0$,} \end{equation} that is, that $f_1(\beta, q, c, \gamma, \delta) \ge 0$, when $0 < \beta \le \beta_0$ and $c = c_4$, for \begin{equation} \label{eq:first_general_approximation} f_1(\beta, q, c, \gamma, \delta) = \beta^{\beta} - \beta^{\beta (\gamma-1)} \left( \frac{\beta^{\beta \delta}}{q} + \frac{(q-1)}{q} \beta^{\beta c} \right). \end{equation} Since ${\displaystyle{\lim_{\beta \rightarrow 0^{+}}}} \beta^{\beta} = 1$, \[ \lim_{\beta \rightarrow 0^{+}} f_1(\beta, q, c, \gamma, \delta) = 1 - \left( {\textstyle{\frac{1}{q}}} + {\textstyle{\frac{q-1}{q}}} \right) = 0. \] \begin{lemma} \label{lem:correctness_of_first_approximation} Suppose that $\gamma = \frac{a}{d}$ and $\delta = \frac{b}{d}$ for non-positive integers~$a$ and~$b$ and a positive integer~$d$. Suppose, as well, that \begin{equation} \label{eq:new_requirement_for_c} c > \frac{2q}{q-1} - \frac{\delta}{q-1} - \frac{q \gamma}{q-1}. \end{equation} If $\Delta$ and~$\widehat{\Delta}$ are positive constants such that $0 < \Delta < \widehat{\Delta} \le \frac{1}{e}$, $f_1(\Delta, q, c, \gamma, \delta) \ge 0$ and $f_1(\widehat{\Delta}, q, c, \gamma, \delta) < 0$, then $f_1(\beta, q, c, \gamma, \delta) \ge 0$ for $0 < \beta \le \Delta$. \end{lemma} \begin{proof} Let $z = \beta^{\beta}$, so that $f_1(\beta, q, c, \gamma, \delta) = p_0(z)$, where \[ p_0(z) = z - \frac{1}{q} z^{\gamma + \delta - 1} - \frac{q-1}{q} z^{c+\gamma - 1}. \] It follows that (differentiating with respect to~$z$) \[ p_0'(z) = 1 - \frac{1}{q}(\gamma + \delta - 1) z^{\gamma + \delta - 2} - \frac{(q-1)}{q}(c+\gamma -1) z^{c+ \gamma - 2}, \] so that \begin{align*} \lim_{z \rightarrow 1^{-}} p_0'(z) &= 1 - \frac{1}{q}(\gamma + \delta - 1) - \frac{(q-1)}{q} (c + \gamma - 1) \\ &= 2 - \frac{\delta}{q} - \gamma - \frac{(q-1)}{q} c \\ &= \frac{q-1}{q} \left( \frac{2q}{q-1} - \frac{\delta}{q-1} - \frac{q \gamma}{q-1} - c \right) \\ &< 0, \end{align*} by the inequality at line~\eqref{eq:new_requirement_for_c}. Thus $p_0$ is a decreasing function as $z$ approaches~$1$ from below and, since ${\displaystyle{\lim_{z \leftarrow 1^{-}}}} p_0(z) = 0$, it follows that $p_0(z) > 0$ when $z$ is less than and sufficiently close to~$1$. Now, since $\beta^{\beta} < 1$ when $0 < \beta < 1$ and ${\displaystyle{\lim_{\beta \rightarrow 0^{+}}}} \beta^{\beta} = 1$, this implies that $f_1(\beta) > 0$ when $\beta$ is positive, and sufficiently small, as well. Recall that $\gamma = \frac{a}{d}$ and $\delta = \frac{b}{d}$ where $a, b, d \in \ensuremath{\mathbb{Z}}$, $a \le 0$, $b \le 0$, and $d > 0$. Thus $\beta^{(1-\gamma-\delta)\beta} f_1(\beta, q, c, \gamma, \delta) = p_1(y)$, where $y = \beta^{\beta/d}$ and \[ p_1(y) = -\frac{(q-1)}{q} y^{cd - b} + y^{2d-a-b} -\frac{1}{q} \in \ensuremath{\mathbb{Q}}[y], \] so that (for $y$ as above) $f_1(\beta, q, c, \gamma, \delta) \ge 0$ if and only if $p_1(y) \ge 0$ --- and (by the above) $p_1(y) > 0$ if $y$ is less than and sufficiently close to~$1$. Now --- regardless of the relationship between $c$ and~$\gamma$ --- there are at most two changes in sign of the nonzero coefficients of this polynomial, when listed by decreasing powers of~$y$. It follows that if $0 < \Delta < \widehat{\Delta} \le \frac{1}{e}$ (so that the function $g(\beta) = \beta^{\beta/d}$ is decreasing, and injective, over the interval $0 < \beta \le \widehat{\Delta}$), $f_1(\Delta) \ge 0$ and $f_1(\widehat{\Delta}) < 0$, then $f_1(\beta) \ge 0$ for $0 < \beta \le \Delta$ --- for, otherwise, the polynomial~$p_1$ would necessarily have at least four positive roots in the interval $\widehat{\Delta}^{\widehat{\Delta}/d} \le y \le 1$, contradicting Descarte's rule of signs. \end{proof} It therefore suffices to check the condition at line~\eqref{eq:new_requirement_for_c} and to confirm that $f_1(\Delta) \ge 0$ and $f_1(\widehat{\Delta}) < 0$, for $0 < \Delta < \widehat{\Delta} \le \frac{1}{e}$, in order to establish that $f_1(\beta) \ge 0$ when $0 < \beta \le \Delta$ --- so that \begin{equation} \label{eq:first_target} \sum_{1 \le j \le k \Delta} \left( \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1 - \beta}} \left( \frac{1}{q} + \frac{q-1}{q} (nq)^{-c_4 \beta} \right) \right)^k \le \sum_{1 \le j \le k \Delta} \beta^{k\beta}. \end{equation} \subsubsection{Bounding Terms in $\boldsymbol{\rho_0}$ When $\boldsymbol{\beta}$ is Larger} \label{ssec:beta_is_moderate} The process described in Subsection~\ref{ssec:beta_is_tiny} can only be used to establish the inequality at line~\eqref{eq:first_target}, above, for small positive constants~$\Delta$ that are generally much smaller than the desired bound $\beta_0 = \frac{2}{c_4}$. However, a complementary process --- which in turn, does not seem to be useful to establish the above inequalities when $\beta$ is extremely close to zero --- can (at least, sometimes) be used to establish that these inequalities hold for larger~$\Delta$ as well. Once again, recall that \[ \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{q} + \frac{q-1}{q} (nq)^{-c_4 \beta} \right) \le \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{q} + \frac{q-1}{q} \left( \frac{\beta}{q} \right)^{c_4 \beta} \right). \] Suppose, now, that $\eta$ is a constant such that $0 < \eta < 1$. Then it certainly follows that \[ \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{q} + \frac{q-1}{q} (nq)^{-c_r \beta} \right) \le \beta^{\beta} \] when $\Gamma_L \le \beta \le \Gamma_H$, for positive constants~$\Gamma_L$ and~$\Gamma_H$, if \begin{equation} \label{eq:first_new_relationship} \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \cdot \frac{1}{q} \le \eta \beta^{\beta} \end{equation} and \begin{equation} \label{eq:second_new_relationship} \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \cdot \frac{q-1}{q} \left(\frac{\beta}{q}\right)^{c_4 \beta} \le (1 - \eta) \beta^{\beta} \end{equation} when $\Gamma_L \le \beta \le \Gamma_H$ as well. Since $q > 1$ and $0 < \beta < 1$, the inequality at line~\eqref{eq:first_new_relationship} holds if and only if \[ \frac{\eta \beta^{2\beta} (1-\beta)^{1-\beta} q}{(q-1)^\beta} \ge 1. \] Considering logarithms, one can see that this is the case for $\Gamma_L \le \beta \le \Gamma_H$ if and only if $F_1(\beta) \ge 0$ for $\Gamma_L \le \beta \le \Gamma_H$, where \begin{equation} \label{eq:defiinition_of_F1} F_1(x) = 2x\ln x + (1-x) \ln (1-x) - x \ln (q-1) + \ln q + \ln \eta. \end{equation} Similarly, the inequality at line~\eqref{eq:second_new_relationship} is satisfied if and only if \[ \frac{(1 - \eta) \beta^{(2 - c_4)\beta} (1-\beta)^{1-\beta} q^{c_4\beta + 1}}{(q-1)^{\beta+1}} \ge 1. \] Considering logarithms, one can see that this is the case for $\Gamma_L \le \beta \le \Gamma_H$ if and only if $F_2(\beta) \ge 0$ for $\Gamma_L \le \beta \le \Gamma_2$, where \begin{equation} \label{eq:definition_of_F2} F_2(x) = (2-c_4)x\ln x + (1-x)\ln(1-x) + (c_4x+1) \ln q - (x+1)\ln(q-1) + \ln (1-\eta). \end{equation} Note next that \[ F_1'(x) = 1 + 2 \ln x - \ln (1-x) - \ln (q-1) \] --- which is independent of~$\eta$ --- and \[ F_1''(x) = \frac{2}{x} + \frac{1}{1-x} \] --- which is positive if $0 < x < 1$. Consequently it it suffices to check that $F_1'(\beta_0) < 0$ in order to establish that $F_1'(x) < 0$ for $0 < x \le \beta_0$, so that $F_1(x)$ is decreasing over the interval $\Delta < x \le \beta_0$. If this is the case, and $\Delta \le \Gamma_L < \Gamma_H \le \beta_0$, then it suffices to check that $F_1(\Gamma_H) \ge 0$ in order to confirm that $F_1(\beta) \ge 0$ for $\Gamma_L \le \beta \le \Gamma_H$ as well. Note as well that \[ F_2'(x) = (1-c_4) + (2-c_4)\ln x - \ln (1-x) + c_4 \ln q - \ln (q-1) \] --- which is also independent of~$\eta$ --- and that \[ F_2''(x) = \frac{2- c_4}{x} + \frac{1}{1-x} \] --- which is negative (for $c_4 > 2$) if $0 < x < \frac{2-c_4}{1-c_4} = 1 + \frac{1}{1-c_4}$, zero if $x = 1 + \frac{1}{1-c_4}$, and positive if $x > 1 + \frac{1}{1-c_4}$. Consequently if $\beta_0 \le 1 + \frac{1}{1-c_4}$ then it suffices to check that $F2'(\beta_0) \ge 0$ in order to establish that $F_2$ is increasing over the interval $\Delta \le x \le \beta_0$. If $\Delta \le 1 + \frac{1}{1 - c_4} \le \beta_0$ then it suffices to check that $F_2'\left(1 + \frac{1}{1-c_4}\right) \ge 0$ in order to confirm that $F_2$ is increasing over this interval. If $1 + \frac{1}{1 -c_4} \le \Delta$ then it suffices to check that $F_2'(\Delta) \ge 0$ in order to confirm this. In any case, if this has been confirmed and $\Delta \le \Gamma_L < \Gamma_H \le \beta_0$, then it suffices to check that $F_2(\Gamma_L) \ge 0$ in order to confirm that $F_2(\beta) \ge 0$ for $\Gamma_L \le \beta \le \Gamma_H$. The desired inequality can now be established, for $\Delta \le \beta \le \beta_0$, by breaking this interval into one or more subintervals, and using the above process with various choices of~$\eta$ to confirm that $F_1$ and~$F_2$ are both non-negative over each subinterval. \subsubsection{Establishing That $f$ is a Decreasing Function} \label{ssec:f_is_decreasing} Once again, consider the function \[ f(x) = (x-1)^{\beta} (x^{-1} + (1 - x^{-1}) (nx)^{-c_4 \beta}) \] as defined at line~\eqref{eq:definition_of_f}. Wiedemann establishes that if $0 < \beta \le \beta_0 < {\textstyle{\frac{1}{4}}}$, $n \ge 1$, $q \ge 2$ and $c_4 > {\textstyle{\frac{4}{\ln 2}}}$, then $f$ is a non-increasing function, so that $f(q) \le f(2)$ for $q \ge 2$ --- as needed to establish that results like the above hold for larger finite fields as well. Unfortunately, this argument requires both $c_4$ and~$c_3$ to assume larger values than are either necessary or desirable. However, Wiedemann's argument can be modified in a straightforward way to establish the following. \begin{lemma} \label{lem:f_is_decreasing} Suppose that $n \ge 1$, $q \ge 16$, $0 < \beta \le \beta_0 \le \frac{12}{13}$, and $c_4 \ge \frac{2}{\beta_0} \ge \frac{13}{6}$. Then $f$ is a decreasing function of~$x$ over the interval $x \ge q$. \end{lemma} \begin{proof} As Wiedemann notes, if $f$ is as defined at line~\eqref{eq:definition_of_f} then \begin{multline*} f'(x) = \beta (x-1)^{\beta-1} (x^{-1} + (1-x^{-1}) (nx)^{-c_4 \beta}) \\ + (x-1)^{\beta}(-x^{-2} + x^{-2}(nx)^{-c_4\beta} -c_4 \beta(x^{-1}-x^{-2}) (nx)^{-c_4\beta}) \end{multline*} so that \[ x^2(x-1)^{-\beta}f'(x) = g(x) + h(x), \] where \[ g(x) = (\beta x - c_4 \beta x + c_4 \beta) (nx)^{-c_4 \beta} \] and \[ h(x) = \frac{\beta x}{x-1} - 1 + (nx)^{-c_4 \beta} = \beta + \beta(x-1)^{-1} - 1 + (nx)^{-c_4 \beta}. \] Consequently, for $x > 1$, if $g(x) < 0$ and $h(x) < 0$ then $f'(x) < 0$ as well. Now, since $\beta (nx)^{-c_4 \beta} > 0$ when $n$, $x$, $c_4$ and~$\beta$ are all positive, it suffices to show that $q - c_4 q + c_4 < 0$ in order to establish that $g(q) < 0$, and $q - c_4 q + c_4 < 0$ if and only if $q > \frac{c_4}{c_4 - 1} = 1 + \frac{1}{c_4-1}$. Since $c_4 \ge \frac{13}{6} > 2$, $1 + \frac{1}{c_4 - 1} < 2$, so that $g(q) < 0$ when $q \ge 16$, as desired. Consider the function~$h$ when $n$, $q$, $\beta$, $\beta_0$ and~$c_4$ are as above. This function is certainly decreasing with both~$x$ and~$c_4$. It therefore suffices to set $x = q = 16$ and $c_4 = \frac{13}{6}$ and show that \[ H(\beta) = h(16) = {\textstyle{\frac{16}{15}}} \beta - 1 + (16n)^{-\frac{13}{6} \beta} < 0 \] when $0 < \beta \le \frac{12}{13}$ in order to establish that $h(x) < 0$, for $\beta$ in this range, and for $q$ and~$n$ as above. Now it is easily checked that ${\displaystyle{\lim_{\beta \rightarrow 0^{+}}}} H(\beta) = 0 - 1 + 1 = 0$. Considered as a function of~$\beta$ (and differentiating with respect to~$\beta$), $H'(\beta) = {\textstyle{\frac{16}{15}}} - {\textstyle{\frac{13}{6}}} \ln (16n) \cdot (16n)^{-\frac{13}{6} \beta}$, so that ${\displaystyle{\lim_{\beta \rightarrow 0^{+}}}} H'(\beta) = \frac{16}{15} - \frac{13}{6} \ln(16n) \le \frac{16}{15} - \frac{13}{6} \ln 16 < -4$. Thus both $H'(\beta)$ and~$H(\beta)$ are negative when $\beta$ is positive and sufficiently small. Note next that $H''(\beta) = \frac{169}{36} \ln(16n)^2 \cdot (16n)^{-\frac{13}{6} \beta} > 0$ whenever $\beta > 0$, so that $H'(\beta)$ is a strictly increasing function of~$\beta$. This admits (only) two possibilities: Either $H(\beta) < 0$ for all $\beta > 0$ --- which certainly establishes the desired result --- or there exists a positive value $\Delta$ such that $H(\beta) < 0$ when $0 < \beta < \Delta$, $H(\Delta) = 0$, and $H(\beta) > 0$ when $\beta > \Delta$. In either case it now suffices to check that $H(\beta) < 0$ when $\beta$ has the maximum value of interest, that is, when $\beta = \frac{12}{13}$. It therefore remains only to note that \[ H\left(\frac{12}{13}\right) = \frac{64}{65} -1 + (16n)^{-2} \le -\frac{1}{65} + \frac{1}{256} < -0.01 < 0 \] in order to complete the proof. \end{proof} \subsubsection{Application of These Processes} \label{ssec:bounding_rho0} The processes described in Subsections~\ref{ssec:beta_is_tiny} and~\ref{ssec:beta_is_moderate} and the result established in Section~\ref{ssec:f_is_decreasing} can now be applied to establish the following. \begin{lemma} \label{lem:bound_when_q=2} If $q = 2$, $0 < \beta \le \frac{6}{43}$ and $c_4 \ge \frac{43}{3}$ then \[ \frac{1}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( {\textstyle{\frac{1}{2}}} + {\textstyle{\frac{1}{2}}} (2n)^{-c_4 \beta} \right) \le \beta^{\beta}. \] \end{lemma} \begin{proof} To begin, let us use the process described in Subsection~\ref{ssec:beta_is_tiny} to establish the above inequality when $0 < \beta \le \frac{5}{43}$. It follows by part~(a) of Lemma~\ref{lem:getting_rid_of_pesky_term} that $(1-x)^{-(1-x)} \le x^{\gamma x}$ when $0 < x \le \frac{5}{43}$ and $\gamma = -\frac{11}{25}$, so that $\gamma$ can be set to have this value when this process is applied. Since $(q-1)^x = 1^x = 1$ when $0 < x \le \frac{5}{43}$, $(q-1)^x \le x^{\delta x}$ in this range when $\delta = 0$, so this value will be used for this constant. In this case \[ c_4 - \left( \frac{2q}{q-1} - \frac{\delta}{q-1} - \frac{q\gamma}{q-1} \right) = \frac{89}{15} > 0, \] so that the condition at line~\eqref{eq:new_requirement_for_c} is satisfied. Since $0 < \frac{5}{43} < \frac{7}{43} \le \frac{1}{e}$, it now suffices to note that $f_1(\frac{5}{43}, 2, \frac{43}{3}, -\frac{11}{25}, 0) > 0.04$ and $f_1(\frac{7}{43}, 2, \frac{43}{3}, -\frac{11}{25}, 0) < - 0.03$ --- for it then follows by Lemma~\ref{lem:correctness_of_first_approximation} that the inequality in the claim is satisfied when $0 < \beta \le \frac{5}{43}$. The process described in Subsection~\ref{ssec:beta_is_moderate} can now be used to establish the above inequality when $\frac{5}{43} \le \beta \le \frac{6}{43}$, completing the proof. Since $F_1'\left(\frac{6}{43}\right) < -2.7$ the function~$F_1$ is decreasing over this interval, for every choice of~$\eta$. Since $\frac{6}{43} < \frac{37}{40} = 1 + \frac{1}{1 - (43/3)}$ and $F_2'\left(\frac{6}{43}\right) > 21$, the function $F_2$ is increasing over this interval for every choice of~$\eta$. It now suffices to confirm that if $\eta = \frac{99}{100}$ then $F_1\left(\frac{6}{43}\right) > 0.004$ and $F_2\left(\frac{5}{43}\right) > 0.2$, so that $F_1$ and $F_2$ are both non-negative over the interval $\frac{5}{43} \le \beta \le \frac{6}{43}$, as desired. \end{proof} It now follows that \begin{equation} \label{eq:fourth_bound_for_rho0_when_q=2} \rho_0 \le \sum_{1 \le j < k \beta_0} \beta^{k\beta} \quad \text{if $q = 2$, $\beta_0 = \frac{6}{43}$, and $c_4 \ge \frac{43}{3}$.} \end{equation} In the above lemma the upper limit, $\beta_0 = \frac{6}{43}$ for~$\beta$, has been chosen so that $\beta_0 = \frac{2}{c_4}$ --- in order to match the constraint between~$\beta_0$ and~$c_4$ as shown at line~\eqref{eq:constraints_on_c2_and_c4}. A plot of the function $\beta^{\beta} - \frac{1}{\beta^{\beta} (1-\beta)^{1-\beta}} \left(\frac{1}{2} + \frac{1}{2} \left(\frac{\beta}{2}\right)^{\frac{43}{3}\beta}\right)$, for $0 < \beta \le \frac{6}{43}$, is shown in Figure~\ref{fig:plot1}. \begin{figure}[t!] \begin{center} \scalebox{0.4}{\includegraphics{plot_1.eps}} \end{center} \caption{Plot of $\beta^{\beta} - \frac{1}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{2} - \frac{1}{2} \left( \frac{\beta}{2}\right)^{\frac{43}{3}\beta}\right)$ when $0 < \beta \le \frac{6}{43}$} \label{fig:plot1} \end{figure} As this may suggest, the above result result can be improved slightly --- but not by very much: The inequality at line~\eqref{eq:fourth_bound_for_rho0_when_q=2}, above, is not satisfied if $c_4$ is decreased to~$14$ and $\beta_0$ increased to~$\frac{1}{7}$. In order to see one more example of this process let us consider the case that $q \ge 3$. An application of the technique described above establishes the following. \begin{lemma} \label{lem:bound_when_q=3} If $q = 3$, $0 < \beta \le \frac{1}{4}$ and $c_4 \ge 8$ then \[ \frac{2^{\beta}}{\beta^{\beta}(1-\beta)^{1-\beta}} \left( {\textstyle{\frac{1}{3}}} + {\textstyle{\frac{2}{3}}} (3n)^{-c_4\beta}\right) \le \beta^{\beta}. \] \end{lemma} \begin{proof} To begin, let us use the process described in Subsection~\ref{ssec:beta_is_tiny} to establish the above inequality when $0 < \beta \le \frac{1}{5}$. It follows by part~(b) of Lemma~\ref{lem:getting_rid_of_pesky_term} that $(1-x)^{-(1-x)} \le x^{\gamma x}$ when $0 < x \le \frac{1}{5}$ and $\gamma = -\frac{23}{40}$, so that $\gamma$ can be set to have this value when this process is applied. It follows by Lemma~\ref{lem:getting_rid_of_power_of_q} that if $q=3$ and $\delta = -\frac{9}{20}$ then $(q-1)^x \le x^{\delta x}$ when $0 < x \le \frac{1}{5}$, so $\delta$ can be set to be $-\frac{9}{20}$ in this argument. In this case \[ c_4 - \left( \frac{2q}{q-1} - \frac{\delta}{q-1} - \frac{q\gamma}{q-1} \right) = \frac{313}{80} > 0, \] so that the condition at line~\eqref{eq:new_requirement_for_c} is satisfied. Since $0 < \frac{1}{5} < \frac{1}{4} \le \frac{1}{e}$, it now suffices to note that $f_1(\frac{1}{5}) > 0.0008$ and $f_1(\frac{1}{4}) < - 0.03$ --- for it then follows by Lemma~\ref{lem:correctness_of_first_approximation} that the inequality in the claim is satisfied when $0 < \beta \le \frac{1}{5}$. The process described in Subsection~\ref{ssec:beta_is_moderate} can now be used to establish the above inequality when $\frac{1}{5} \le \beta \le \frac{1}{4}$, completing the proof. Since $F_1'\left(\frac{1}{4}\right) < -2.1$ the function~$F_1$ is decreasing over this interval, for every choice of~$\eta$. Since $\frac{1}{4} < \frac{6}{7} = 1 + \frac{1}{1 - 8}$ and $F_2'\left(\frac{1}{4}\right) > 9$, the function $F_2$ is increasing over this interval for every choice of~$\eta$. It now suffices to confirm that if $\eta = \frac{39}{40}$ then $F_1\left(\frac{6}{25}\right) > 0.01$ and $F_2\left(\frac{1}{5}\right) > 0.08$, so that $F_1$ and $F_2$ are both non-negative over the interval $\frac{1}{5} \le \beta \le \frac{6}{25}$. It then suffices to confirm that if $\eta = \frac{123}{125}$ then $F_1\left(\frac{1}{4}\right) > 0.0002$ and $F_2\left(\frac{6}{25}\right) > 0.05$, so that $F_1$ and $F_2$ are both non-negative over the interval $\frac{6}{25} \le x \le \frac{1}{4}$, as needed to complete the proof. \end{proof} It now follows that \begin{equation} \label{final_bound_for_rho0_when_q=3} \rho_0 \le \sum_{1 \le j < k \beta_0} \beta^{k\beta} \quad \text{if $q = 3$, $\beta_0 = \frac{1}{4}$, and $c_4 \ge 4$.} \end{equation} A plot of the function $\beta^{\beta} - \frac{2^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{3} + \frac{2}{3} \left(\frac{\beta}{3}\right)^{8 \beta}\right)$, when $0 < \beta \le \frac{1}{4}$, is shown in Figure~\ref{fig:plot2}. \begin{figure}[t!] \begin{center} \scalebox{0.4}{\includegraphics{plot_2.eps}} \end{center} \caption{Plot of $\beta^{\beta} - \frac{1}{\beta^{\beta} (1-\beta)^{1-\beta}} \left(\frac{1}{3} + \frac{2}{3} \left(\frac{\beta}{3}\right)^{8\beta}\right)$ when $0 < \beta \le \frac{1}{4}$} \label{fig:plot2} \end{figure} Once again, this suggests that the above result cannot be improved by very much. Appendix~\ref{sec:continuation_of_argument} include details of analyses for additional field sizes as well --- as summarized in Figure~\ref{fig:choices_of_c4} on page~\pageref{fig:choices_of_c4}. A Maple worksheet, that can be used to check these details, is available online at \begin{center} \url{http://www.cpsc.ucalgary.ca/~eberly/Research/sparse_conditioner.mw}. \end{center} \begin{figure}[t!] \begin{center} \begin{tabular}{ccc|ccc|ccc} $q$ & $c_4$ & $\beta_0$ & $q$ & $c_4$ & $\beta_0$ & $q$ & $c_4$ & $\beta_0$ \\[2pt] \hline $2$ \rule{0pt}{12pt} & $\frac{43}{3}$ & $\frac{6}{43}$ & $9$ & $\frac{11}{3}$ & $\frac{6}{11}$ & $47$--$59$ & $\frac{7}{3}$ & $\frac{6}{7}$ \\ $3$ \rule{0pt}{12pt} & $8$& $\frac{1}{4}$ & $11$ & $\frac{7}{2}$ & $\frac{4}{7}$ & $61$--$71$ & $\frac{9}{4}$ & $\frac{8}{9}$ \\ $4$ \rule{0pt}{12pt} & $\frac{25}{4}$ & $\frac{8}{25}$ & $13$ & $\frac{10}{3}$ & $\frac{3}{5}$ & $73$--$83$ & $\frac{11}{5}$ & $\frac{10}{11}$ \\ $5$ \rule{0pt}{12pt} & $\frac{16}{3}$ & $\frac{3}{8}$ & $16$--$19$ & $3$ & $\frac{2}{3}$ & $\ge 89$ & $\frac{13}{6}$ & $\frac{12}{13}$ \\ $7$ \rule{0pt}{12pt} & $\frac{13}{3}$ & $\frac{6}{13}$ & $23$--$29$ & $\frac{8}{3}$ &$\frac{3}{4}$ & \\ $8$ \rule{0pt}{12pt} & $4$ & $\frac{1}{2}$ & $31$--$43$ & $\frac{5}{2}$ & $\frac{4}{5}$ & & \end{tabular} \end{center} \caption{Choices of~$c_4$ and~$\beta_0$ for Various Field Sizes} \label{fig:choices_of_c4} \end{figure} It follows from this that \[ \rho_0 \le \sum_{1 \le j < k \beta_0} \beta^{k\beta} \] for each of the choices of~$q$, $\beta_0$ and~$c_4$ given in Figure~\ref{fig:choices_of_c4}. \subsection{Asymptotic Results: Choice of Field Size} \label{ssec:asymptotics_1} The objective of this next subsection is to identify bounds on the sizes of primes allowing the inequality at line~\eqref{eq:big_goal} to be established when $c_4$ is closer to~$2$. Suppose, in particular, that $N$ is an integer such that $N \ge 18$ and that \begin{equation} \label{eq:asymptotic_constraints_on_constants} c_4 = 2 + {\textstyle{\frac{1}{N}}} \qquad \text{and} \qquad q \ge 16N + 9. \end{equation} Consider now the function \begin{equation} \label{eq:definition_of_g} \begin{split} g(\beta) &= \frac{\left( \frac{(q-1)^{\beta}}{\beta^{\beta} (1-\beta)^{1-\beta}} \right) \left( \frac{1}{q} + \left(\frac{q-1}{q}\right) \left( \frac{\beta}{q} \right)^{c\beta} \right)}{\beta^{\beta}} \\ &= \frac{(q-1)^{\beta}}{\beta^{2\beta} (1-\beta)^{1-\beta}} \left( \frac{1}{q} + \left(\frac{q-1}{q}\right) \left( \frac{\beta}{q} \right)^{c_4 \beta} \right) \end{split} \end{equation} noting --- by the inequality at line~\eqref{eq:what_is_known_1} --- that the inequality at line~\eqref{eq:big_goal} is satisfied if $g(\beta) \le 1$ when $0 < \beta \le \beta_0 = \frac{2}{c_4}$. Consider, as well the function \begin{equation} \label{eq:definition_of_h} \begin{split} h(\beta) &= \ln g(\beta) \\ &= \beta \ln (q-1) - 2 \beta \ln \beta - (1 - \beta) \ln (1-\beta) \\ &\hspace*{1 true in} + \ln \left( \frac{1}{q} + \left( \frac{q-1}{q} \right) \left(\frac{\beta}{q} \right)^{c_4 \beta} \right) \end{split} \end{equation} observing that \begin{equation} \label{eq:definition_of_hprime} h'(\beta) = \ln (q-1) - 2 \ln \beta + \ln (1-\beta) - 1 + \frac{H(\beta)}{H(\beta)+1} (c_4 + c_4 \ln \beta - c_4 \ln q) \end{equation} where \begin{equation} \label{eq:definition_of_H} H(\beta) = (q-1) \left( \frac{\beta}{q}\right)^{c_4 \beta}. \end{equation} \begin{lemma} \label{lem:first_asymptotic_interval} If $0 < \beta \le \frac{1}{50 c_4 \ln q}$ then $g(\beta) \le 1$. \end{lemma} \begin{proof} It is easily checked that $\displaystyle{\lim_{\beta \rightarrow 0^+} g(\beta)} = 1$ and $\displaystyle{\lim_{\beta \rightarrow 0^+} h(\beta)} = 0$. The claim can therefore be established by showing that $h'(\beta) < 0$ when $0 < \beta \le \frac{1}{50 c_4 \ln q}$. Consider the above function $H(\beta)$, recalling as well that \begin{equation} \label{eq:log_approximation} \frac{x}{1+x} \le \ln (1+x) \le x, \end{equation} for any real number~$x$ such that $x > -1$. Replacing $x$ with $-x$ (where $x < 1$), one has that \[ - \frac{x}{1-x} \le \ln (1-x) \le -x \] as well, and replacing $x$ with $1-x$ (where, once again $x < 1$) one has that \[ 1 - \frac{1}{x} \le \ln x \le x-1. \] It follows from this that (for $0 < x < 1$) \[ x - 1 \le x \ln x \le x^2 - x, \] so that \[ e^{x-1} \le x^x \le e^{x^2 - x} \le 1 \] when $0 < x \le 1$. It follows from the definition of $H(\beta)$ at line~\eqref{eq:definition_of_H} that \begin{align*} H(\beta) &\ge (q-1) \frac{e^{(\beta-1) c_4}}{q^{c_4 \beta}} \\ &\ge (q-1) \frac{e^{-c_4}}{q^{c_4 \beta}} \\ &= (1 - q^{-1}) e^{-c_4} q^{1 - c_4 \beta} \\ &\ge (1 - q^{-1}) e^{-\frac{37}{18}} q^{1- c_4 \beta} \tag{since $N \ge 18$, so that $c_4 \le \frac{37}{18}$} \\ &\ge \frac{296}{297} \frac{e^{-\frac{37}{18}}}{q^{c_4 \beta}} q \tag{since $q = 16N + 9 \ge 297$}\\ &\ge \frac{296}{297} \frac{e^{-\frac{37}{18}}}{q^{\frac{1}{50 \ln q}}} q & \tag{since $\beta \le \frac{1}{50 c_4 \ln q}$} \\ &= \frac{296}{297} e^{-\frac{37}{18} - \frac{1}{50}} q\\ &> \frac{q}{8} \\ &\ge 2N+1 \tag*{(since $q \ge 16N + 9$).} \\ \end{align*} It follows that \[ \frac{H(\beta)}{H(\beta)+1} = 1 - (H(\beta)+1)^{-1} \ge 1 - (2N+1)^{-1} = \frac{2N}{2N+1}. \] Note, as well that $c_4 + c_4 \ln \beta - c_4 \ln q = c_4 \left( 1 + \ln \textstyle{\frac{\beta}{q}} \right) < c_4 \left( 1 + \ln \textstyle{\frac{1}{q}} \right) < 0$. It now follows by the equation at line~\eqref{eq:definition_of_hprime} that \begin{align*} h'(\beta) &= \ln (q-1) - 2 \ln \beta + \ln (1-\beta) -1 + \frac{H(\beta)}{H(\beta) +1} (c_4 + c_4 \ln \beta - c_4 \ln q) \\ &\le \ln (q-1) - 2 \ln \beta + \ln (1-\beta) - 1 + \textstyle{\frac{2N}{2N+1}} (c_4 + c_4 \ln \beta - c_4 \ln q) \\ &= \ln (q-1) - 2 \ln \beta + \ln (1-\beta) - 1 + 2(1+ \ln \beta - \ln q) \tag{since $c_4 = 2 + \frac{1}{N} = \frac{2N+1}{N}$} \\ &< - \ln q + \ln (1 - \beta) + 1 < 0, \end{align*} as required. \end{proof} A different approach is required for larger values of~$\beta$ because $h'(\beta)$ is eventually positive. Note that $g(\beta) = g_1(\beta) + g_2(\beta)$, where \begin{equation} \label{eq:definition_of_g1} g_1(\beta) = \frac{(q-1)^{\beta}}{\beta^{2\beta} (1-\beta)^{1-\beta}} \cdot \frac{1}{q} = q^{-1} (q-1)^{\beta} \cdot \frac{(1-\beta)^{\beta-1}}{\beta^{2\beta}} \end{equation} and \begin{equation} \label{eq:definition_of_g2} g_2(\beta) = \frac{(q-1)^{\beta}}{\beta^{2\beta} (1-\beta)^{1-\beta}} \cdot \left( \frac{q-1}{q} \right) \left( \frac{\beta}{q} \right)^{c_4 \beta} =\left( 1-\frac{1}{q}\right) \frac{\beta^{\beta/N}}{ \left( \frac{q^{c_4-1}}{1-\beta} \right)^{\beta} (1- \beta)}. \end{equation} \begin{lemma} \label{lem:second_asymptotic_interval} If $\frac{1}{50 c_4 \ln q} \le \beta \le \frac{7}{8}$ then $g(\beta) \le 1$. \end{lemma} \begin{proof} Consider first the function \[ k(x) = \frac{(1-x)^{x-1}}{x^{2x}} \] and \[ \ell(x) = \ln k(x) = (x-1) \ln (1-x) - 2x \ln x, \] noting that \[ \ell'(x) = \ln (1-x) - 2 \ln x - 1 \] and \[ \ell''(x) = - \frac{1}{1-x} - \frac{2}{x}, \] so that $\ell''(x) < 0$ when $0 < x < 1$. Note that $\ell'(x) = 0$ when \[ \frac{1-x}{x^2} = e, \] that is, when $e x^2 + x - 1 = 0$. Applying the quadratic equation, one can see that the functions $k(x)$ and~$\ell(x)$ are both increasing when $0 < x < \frac{-1 + \sqrt{1+4e}}{2e}$, and decreasing when $\frac{1 + \sqrt{1+4e}}{2e} < x < 1$. In particular, $k(x) \le k\left(\frac{-1 + \sqrt{1 + 4e}}{2e}\right) < 3$ for $0 < x < 1$. This be used to obtain upper bounds for the function~$g_1$, shown at line~\eqref{eq:definition_of_g1}, above, over various intervals. Suppose, in particular, that $k(\beta) \le \delta$ when $\Delta_L < \beta \le \Delta_H$ for constants $\delta$, $\Delta_L$ and~$\Delta_H$ such that $0 < \Delta_L < \Delta_H < 1$. It follows that if $\Delta_L < \beta \le \Delta_H$ then \[ g_1(\beta) \le \delta q^{-1} (q-1)^{\beta} \le {\textstyle{\frac{\delta}{297}}} 296^{\beta} \quad \text{when $\beta \le \textstyle{\frac{7}{8}}$} \] since $q \ge 297$; this upper bound for $g_2(\beta)$ is increasing with $\beta$. Note, as well, that, since $\frac{x}{1+x} \le \ln (1+x) \le x$ for any real number~$x$ such that $x > -1$, $-\frac{x}{1-x} \le \ln (1-x) \le -x$ for any real number~$x$ such that $0 < x < 1$. Consequently $-x \le (1-x) \ln (1-x) \le -x + x^2$, so that \[ e^{-x} \le (1-x)^{1-x} \le e^{-x+x^2}, \] and (replacing $x$ with $1-x$, and applying the bounds for $(1-x)^{1-x}$) \[ e^{1-x} \le x^x \le e^{-x+x^2} \] when $0 < x < 1$. Thus, if $g_2(\beta)$ is as given at line~\eqref{eq:definition_of_g2}, above, then \begin{align*} \label{eq:bound_for_g2} g_2(\beta) &= \left( 1 - \frac{1}{q} \right) \frac{\beta^{\beta/N}}{\left( \frac{q^{c_4-1}}{1-\beta}\right)^{\beta} (1-\beta)} \\ &\le \left( 1 - \frac{1}{q} \right) \frac{1}{q^{(c_4 -1)\beta} (1-\beta)^{1-\beta}} \tag{since $\beta^{\beta} \le e^{-\beta + \beta^2} \le 1$} \\ &\le \left( 1 - \frac{1}{q} \right) \left( \frac{e}{q^{c_4-1}} \right)^{\beta} & \tag{since $(1-\beta)^{1-\beta} \ge e^{-\beta}$} \\ &< \left( 1 - \frac{1}{q} \right) \left( \frac{e}{q} \right)^{\beta} \tag{since $c_4 > 2$} \\ &\le {\textstyle{\frac{296}{297}}} \left( {\textstyle{\frac{e}{297}}} \right)^{\beta} \tag*{(since $q \ge 297$).} \end{align*} Since $q > e$, this upper bound for $g_2(\beta)$ is certainly decreasing as $\beta$ increases. It follows by Lemma~\ref{lem:first_asymptotic_interval} that $g(\beta) = g_1(\beta) + g_2(\beta) < 1$ when $\beta = \frac{1}{50 c_4 \ln q}$. Now \begin{align*} g_2 \left( {\textstyle{\frac{1}{50 c_4 \ln q}}} \right) &\le \left( {\textstyle{1 - \frac{1}{q}}} \right) \left( {\textstyle{\frac{e}{q}}} \right)^{\frac{1}{50 c_4 \ln q}} \\ &\le \left( {\textstyle{1 - \frac{1}{q}}} \right) \left( e^{\frac{1}{50 c_4}} \right)^{\frac{1}{\ln q} - 1} \\ &\le \left( e^{\frac{1}{50 c_4}} \right)^{\frac{1}{\ln q}-1}. \\ \end{align*} Since $e^{\frac{1}{50 c_4}} > 1$, $q \ge 297$, and the above exponent $\frac{1}{\ln q}- 1$ decreases as $q$ increases, it now follows that \[ g_2\left({\textstyle{\frac{1}{50 c_4 \ln q}}} \right) \le \left( e^{\frac{1}{50 c_4}} \right)^{\frac{1}{\ln 297} - 1}. \] Now, since $N \ge 18$, $c_4 \le \frac{37}{18}$ and $\frac{1}{50 c_4} \ge \frac{9}{925}$. Since the exponent in the above expression is negative, it now follows that \[ g_2\left({\textstyle{\frac{1}{50 c_4 \ln q}}} \right) \le \left( e^{\frac{9}{925}} \right)^{\frac{1}{\ln 297} - 1} < \frac{397}{400}. \] Since $g_2$ is decreasing with $\beta$ and $g_1$ is increasing with~$\beta$, it now suffices to choose a value~$\gamma$ such that $\frac{1}{50 c_4 \ln q} < \gamma \le \frac{7}{8}$, and $g_1(\gamma) \le \frac{3}{400}$ in order to conclude that $g(\beta) = g_1(\beta) + g_2(\beta) \le 1$ when $\frac{1}{50 c_4 \ln q} \le \beta \le \gamma$. Suppose now that $\gamma \le \frac{1}{8}$ as well; then $k(\gamma) \le k\left(\frac{1}{8}\right) \le 2$ so that $g_1(\beta) \le \frac{2}{297} 296^{\beta}$. It follows from that that if $\gamma = \frac{1}{75}$ then $\frac{1}{50 c_4 \ln q} < \frac{1}{75} < \frac{1}{8}$ and $g_1(\gamma) < \frac{3}{400}$, as required to conclude that $g(\beta) < 1$ when $\frac{1}{50 c_4 \ln q} \le \beta \le \frac{1}{75}$. Note next that $g_2\left(\frac{1}{75}\right) \le \frac{296}{297} \left( \frac{e}{297} \right)^{\frac{1}{75}} < \frac{19}{20}$. Since $g_2$ is decreasing with~$\beta$ and $g_1$ is increasing with $\beta$, it suffices to choose a value~$\widehat{\gamma}$ such that $\frac{1}{75} < \widehat{\gamma} \le \frac{7}{8}$ and $g(\widehat{\gamma}) < \frac{1}{20}$ in order to conclude that $g(\beta) \le 1$ when $\frac{1}{75} \le \beta \le \widehat{\gamma}$ as well. As noted above, $k(\beta) < 3$ when $0 < \beta < 1$, and it follows that $g_1(\beta) \le \frac{3}{297} 296^{\beta}$ for all such~$\beta$. It follows from this that if $\widehat{\gamma} = \frac{1}{4}$ then $g_1(\gamma) < \frac{1}{20}$, as needed to conclude that $g(\beta) \le 1$ when $\frac{1}{75} \le \beta \le \frac{1}{4}$. Now note that $g_2(\left(\frac{1}{4}\right) \le \frac{296}{97} \left( \frac{e}{297} \right)^{\frac{1}{4}} < \frac{1}{3}$. Since $g_2$ is decreasing with~$\beta$ and $g_1$ is increasing with~$\beta$ it suffices to choose a value $\widetilde{\gamma}$ such that $\frac{1}{4} < \widetilde{\gamma} \le \frac{7}{8}$ and $g(1)(\widetilde{\gamma}) \le \frac{2}{3}$ in order to conclude that $g(\beta)\le1$ when $\frac{1}{4} \le \beta \le \widetilde{\gamma}$. Once again, $g_1(\widetilde{\gamma}) \le \frac{3}{297} 296^{\widetilde{\gamma}}$, and this suffices to set $\widetilde{\gamma} = \frac{2}{3}$ in order to ensure that $g_1(\widetilde{\gamma}) < \frac{2}{3}$, as needed. Now $g_2(\left(\widetilde{\beta}\right) = g_2\left(\frac{2}{3}\right) < \frac{1}{20}$, so it suffices to choose $\overline{\gamma}$ such that $\frac{2}{3} < \overline{\gamma} \le \frac{7}{8}$ and $g_1(\overline{\gamma}) \le\frac{19}{20}$ in order to ensure that $g(\beta) \le 1$ when $\frac{2}{3} \le \beta \overline{\gamma}$. Now, $k(x) \le \frac{5}{2}$ when $\frac{2}{3} \le x \le \frac{7}{8}$, so that $g_1(x) \le \frac{5}{2} \cdot \frac{1}{297} \cdot 296^x$ for all $x$ in this range, and this can be used to establish that one can set $\overline{\beta} = \frac{4}{5}$ in order to ensure that the desired conditions are met. Finally, $g_2\left(\frac{4}{5}\right) < \frac{23}{1000}$ and $k(x) < \frac{99}{50}$ when $\frac{4}{5} \le x \le \frac{7}{8}$. This can be used to establish that $g_1(\left(\frac{7}{8}\right) < \frac{977}{1000}$, as needed to establish that $g(\beta) \le 1$ when $\frac{4}{5} \le \beta \le \frac{7}{8}$ and complete the proof of the claim. \end{proof} Once again the functions~$h(\beta)$, $h'(\beta)$ and~$H(\beta)$, shown at lines~\eqref{eq:definition_of_h}--\eqref{eq:definition_of_H}, are of use to prove the desired result for larger values of~$\beta$. \begin{lemma} \label{lem:third_asymptotic_interval} If $\frac{7}{8} \le \beta \le \frac{2N}{2N+1}$ then $g(\beta) \le 1$. \end{lemma} \begin{proof} Consider the functions~$h(\beta)$, $h'(\beta)$ and~$H(\beta)$. Note first that if $\frac{7}{8} \le \beta \le \frac{2N}{2N+1}$ then \begin{align*} H(\beta)&= (q-1) \left({\textstyle{\frac{\beta}{q}}}\right)^{c_4 \beta} \\ &\le (q-1) \left({\textstyle{\frac{1}{q}}}\right)^{c_4 \beta}\tag{since $\beta < 1$, $q >0$,and $c_4 \beta > 1$} \\ &\le q^{1 - c_4 \beta} \\ &\le q^{-\frac{3}{4}} \tag*{(since $c_4 \ge 2$ and $\beta \ge \frac{7}{8}$, so that $1 - c_4 \beta \le - \frac{3}{4}$).} \end{align*} Since $H(\beta) \ge 0$ as well, it follows that \[ \frac{H(\beta)}{H(\beta)+1} \le H(\beta) \le q^{-\frac{3}{4}} \] as well. Since $c_4 = 2 + \frac{1}{N} \le 2 + \frac{1}{18} = \frac{37}{18}$, \[ c_4 \frac{H(\beta)}{H(\beta)+1} \le \frac{37}{18} \cdot q^{-\frac{3}{4}}. \] Since $1 + \ln \beta - \ln q < 0$ it now follows that \begin{align*} h'(\beta) &= \ln (q-1) - 2 \ln \beta + \ln (1-\beta) - 1 + \frac{H(\beta)}{H(\beta)+1} (c_4 + c_4 \ln \beta - c_4 \ln q) \\ &\ge \ln (q-1) - 2 \ln \beta + \ln (1-\beta) - 1 + \frac{37}{18} \cdot q^{-\frac{3}{4}} (1 + \ln \beta + \ln q) \\ &\ge \ln (16N+8) - 2 \ln \beta + \ln \left(\frac{1}{2N+1}\right) - 1 + \frac{37}{18} \cdot q^{-\frac{3}{4}} (1 + \ln \beta + \ln q) \tag{since $q-1 \ge 16N+8$ and $1 - \beta \ge \frac{1}{2N+1}$} \\ &= \ln 8 - 2 \ln \beta - 1 + \frac{37}{18} \cdot q^{-\frac{3}{4}}(1 + \ln \beta + \ln q) \\ &= (\ln 8 - 1 + {\textstyle{\frac{37}{18}}} \cdot q^{-\frac{3}{4}} (1 + \ln q)) - (2 - {\textstyle{\frac{37}{18}}} \cdot q^{-\frac{3}{4}}) \ln \beta \\ &\ge {\textstyle{\frac{9}{10}}} - (2 - {\textstyle{\frac{37}{18}}} \cdot q^{-\frac{3}{4}}) \ln \beta \tag{since $\ln 8 - 1 + {\textstyle{\frac{37}{18}}} \cdot q^{-\frac{3}{4}} (1 + \ln q) \ge \ln 8 - 1 + {\textstyle{\frac{37}{18}}} \cdot 267^{-\frac{3}{4}} (1 + \ln 267) \le {\textstyle{\frac{9}{10}}}$} \\ &\ge {\textstyle{\frac{9}{10}}} \tag{since $\ln \beta < 0$ and $2 - {\textstyle{\frac{37}{18}}} \cdot q^{-\frac{3}{4}} \ge 2 - {\textstyle{\frac{37}{18}}} \cdot 267^{-\frac{3}{4}} \ge {\textstyle{\frac{19}{10}}} > 0$} \\ &> 0. \end{align*} Thus the function $h(\beta)$ is increasing over the interval $\frac{7}{8} \le \beta \le \frac{2N}{2N+1}$. Since $h(\beta) = \ln g(\beta)$, the function $g(\beta)$ is increasing as well --- and it suffices to confirm that $g\left(\frac{2N}{2N+1}\right) \le 1$ in order to establish the claim. Now recall that $g\left(\frac{2N}{2N+1}\right) = g_1\left(\frac{2N}{2N+1}\right) + g_2\left(\frac{2N}{2N+1}\right)$, for the functions~$g_1(\beta)$ and~$g_2(\beta)$ as defined at lines~\eqref{eq:definition_of_g1} and~\eqref{eq:definition_of_g2} respectively. Applying these definitions one can see that \begin{align*} g_1\left({\textstyle{\frac{2N}{2N+1}}}\right) &= q^{-1} (q-1)^{\frac{2N}{2N+1}} \frac{\left({\textstyle{\frac{1}{2N+1}}}\right)^{-\frac{1}{2N+1}}} {\left({\textstyle{\frac{2N}{2N+1}}}\right)^{\frac{4N}{2N+1}}} \\ &= \left( 1 - {\textstyle{\frac{1}{q}}} \right) \frac{(q-1)^{-\frac{1}{2N+1}} \cdot \left( {\textstyle{\frac{1}{2N+1}}} \right)^{-\frac{1}{2N+1}}} {\left( {\textstyle{\frac{2N}{2N+1}}} \right)^{\frac{4N}{2N+1}}} \\ &\le \frac{16N+8}{16N+9} \cdot \frac{\left((16N+8) \cdot \left( {\textstyle{\frac{1}{2N+1}}}\right)\right)^{-\frac{1}{2N+1}}} {\left( {\textstyle{\frac{2N}{2N+1}}} \right)^{\frac{4N}{2N+1}}} \tag{since $q \ge 16N+9$} \\ &= \frac{16N+8}{16N+9} \cdot \frac{8^{-\frac{1}{2N+1}}} {\left( {\textstyle{\frac{2N}{2N+1}}} \right)^{\frac{4N}{2N+1}}} \\ &\le \frac{16N+8}{16N+9} \cdot \frac{8^{-\frac{1}{2N+1}}} {e^{-\frac{2}{2N+1}}} \tag{since ${\textstyle{\left( \frac{2N}{2N+1} \right)}}^{\frac{4N}{2N+1}} \ge e^{-\frac{2}{2N+1}}$} \\ &= \frac{16N+8}{16N+9} \cdot \left( \frac{e^2}{8} \right)^{\frac{1}{2N+1}} \\ &\le \frac{16N+8}{16N+9} \tag{since $\textstyle{\frac{e^2}{8}} < 1$} \end{align*} and \begin{align*} g_2\left({\textstyle{\frac{2N}{2N+1}}}\right) &= \left(1 - {\textstyle{\frac{1}{q}}}\right) \frac{\left( {\textstyle{\frac{2N}{2N+1}}} \right)^{\frac{2}{2N+1}} } { \left( \frac{q^{\frac{N+1}{N}}}{\frac{1}{2N+1}} \right)^{\frac{2N}{2N+1}} \left( {\textstyle{\frac{1}{2N+1}}} \right) } \\ &= \left(1 - {\textstyle{\frac{1}{q}}}\right) \frac{\left( {\textstyle{\frac{2N}{2N+1}}} \right)^{\frac{2}{2N+1}} } { \left( q^{\frac{N+1}{N}} \right)^{\frac{2N}{2N+1}} \left( {\textstyle{\frac{1}{2N+1}}} \right)^{\frac{1}{2N+1}} } \\ &= \left(1 - {\textstyle{\frac{1}{q}}}\right) \frac{\left( {\textstyle{\frac{(2N)^2}{2N+1}}} \right)^{\frac{1}{2N+1}} q^{-\frac{1}{2N+1}} } {q} \\ &\le \left(1 - {\textstyle{\frac{1}{q}}}\right) \frac{\left( {\textstyle{\frac{(2N)^2}{(2N+1)^2}}} \right)^{\frac{1}{2N+1}} }{q} \tag{since $q \ge 16N+9 \ge 2N+1$} \\ &\le \left(1 - {\textstyle{\frac{1}{q}}}\right) \cdot {\textstyle{\frac{1}{q}}} \\ &\le \frac{1}{q} \\ &\le \frac{1}{16N+9} \tag{since $q \ge 16N+9$} \end{align*} as needed to establish that $g\left(\frac{2N}{2N+1}\right) \le 1$ and complete the proof. \end{proof} The following is now a straightforward consequence of Lemmas~\ref{lem:first_asymptotic_interval}--\ref{lem:third_asymptotic_interval}. \begin{corollary} \label{cor:asymptotic_primes} If $N \ge 18$, $c_4$ and~$q$ are as shown at line~\eqref{eq:asymptotic_constraints_on_constants}, and $\beta_0 = \frac{2}{c_4} = \frac{2N}{2N+1}$, the the inequality at line~\eqref{eq:big_goal} is satisfied. \end{corollary} \subsection{Splitting the Sum to Get a Better Bound for $\boldsymbol{\rho_0}$} \label{ssec:splitting_sum} Let $\epsilon$ be a positive constant. Suppose now that $\Delta$ is a positive integer whose depends on~$\epsilon$ and the field size~$q$. It follows from the above (for appropriate choices of~$q$, $\beta_0$ and~$c_4$) that \begin{align*} \rho_0 &\le \sum_{1 \le j < k \beta_0} \beta^{k \beta} \\ &= \sum_{1 \le j < k \beta_0} \left(\frac{j}{k}\right)^j & \tag{since $\beta = \frac{j}{k}$} \\ &= \zeta+ \theta, \end{align*} where \begin{equation} \label{eq:defn_of_zeta} \zeta = \sum_{1 \le j \le \Delta} \left( \frac{j}{k} \right)^j \end{equation} and \begin{equation} \label{eq:defn_of_theta} \theta = \sum_{\Delta < j < k \beta_0} \left( \frac{j}{k} \right)^j. \end{equation} It follows from the above that \begin{align*} \theta &\le \sum_{\Delta < j \le k \beta_0} \left( \frac{j}{k} \right) ^j\\ &\le \sum_{\Delta < j \le k \beta_0} \beta_0^j \\ &\le \beta_0^{\Delta + 1} \sum_{j \ge 0} \beta_0^j \\ &= \frac{\beta_0^{\Delta+1}}{1 - \beta_0} \\ &\le {\textstyle{\frac{1}{10}}} \epsilon \end{align*} provided that \[ \Delta \ge \frac{\ln (\epsilon^{-1}) + \ln 10 - \ln (1 - \beta_0)}{\ln (1/\beta_0)} - 1. \] Choices of~$\Delta$ that satisfy this inequality for various field sizes (and corresponding choices of~$\beta_0$) are shown in Figure~\ref{fig:choices_of_Delta} on page~\pageref{fig:choices_of_Delta}. \begin{figure}[t!] \begin{center} \begin{tabular}{ccc|ccc|ccc} $q$ & $\beta_0$ & $\Delta$ & $q$ & $\beta_0$ & $\Delta$ & $q$ & $\beta_0$ & $\Delta$ \\[2pt] \hline $2$ \rule{0pt}{12pt} & $\frac{6}{43}$ & $\lceil \frac{51}{100} \ln (\epsilon^{-1}) + \frac{1}{4} \rceil$ & $9$ & $\frac{6}{11}$ & $\lceil \frac{33}{20} \ln (\epsilon^{-1}) + \frac{41}{10} \rceil$ & $47$--$59$ & $\frac{6}{7}$ & $\lceil \frac{13}{2} \ln (\epsilon^{-1}) + \frac{133}{5} \rceil$ \\ $3$ \rule{0pt}{12pt} & $\frac{1}{4}$ & $\lceil \frac{73}{100} \ln (\epsilon^{-1}) + \frac{9}{10} \rceil$ & $11$ & $\frac{4}{7}$ & $\lceil \frac{179}{100} \ln (\epsilon^{-1}) + \frac{47}{10} \rceil$ & $61$--$71$ & $\frac{8}{9}$ & $\lceil \frac{17}{2} \ln (\epsilon^{-1}) + \frac{149}{4} \rceil$ \\ $4$ \rule{0pt}{12pt} & $\frac{8}{25}$ & $\lceil \frac{22}{25} \ln (\epsilon^{-1}) + \frac{7}{5} \rceil$ & $13$ & $\frac{3}{5}$ & $\lceil \frac{49}{25} \ln (\epsilon^{-1}) + \frac{107}{20} \rceil$ & $73$--$83$ & $\frac{10}{11}$ & $\lceil \frac{21}{2} \ln (\epsilon^{-1}) + \frac{242}{5} \rceil$ \\ $5$ \rule{0pt}{12pt} & $\frac{3}{8}$ & $\lceil \frac{51}{50} \ln (\epsilon^{-1}) + \frac{19}{10} \rceil$ & $16$--$19$ & $\frac{2}{3}$ & $\lceil \frac{99}{40} \ln (\epsilon^{-1}) + \frac{37}{5} \rceil$ & $\ge 89$ & $\frac{12}{13}$ & $\lceil \frac{25}{2} \ln (\epsilon^{-1}) + \frac{599}{10} \rceil$ \\ $7$ \rule{0pt}{12pt} & $\frac{6}{13} $ & $\lceil \frac{13}{10} \ln (\epsilon^{-1}) + \frac{14}{5} \rceil$ & $23$--$29$ & $\frac{3}{4}$ & $\lceil \frac{87}{25} \ln (\epsilon^{-1}) + \frac{119}{10} \rceil$ & \\ $8$ \rule{0pt}{12pt} & $\frac{1}{2}$ & $\lceil \frac{29}{20} \ln (\epsilon^{-1}) + \frac{17}{5} \rceil$ & $31$--$43$ & $\frac{4}{5}$ & $\lceil \frac{9}{2} \ln (\epsilon^{-1}) + \frac{50}{3} \rceil$ & \\ \end{tabular} \end{center} \caption{Choices of $\Delta$ for Various Field Sizes} \label{fig:choices_of_Delta} \end{figure} When $N \ge 18$ and $\beta_0 = \frac{2N}{2N+1}$ (as in Subsection~\ref{ssec:asymptotics_1}) \begin{align*} \frac{\ln (\epsilon^{-1}) + \ln 10 - \ln(1 - \beta_0)}{\ln(1/\beta_0} - 1 &= \frac{\ln (\epsilon^{-1} + \ln 10 - \ln \left( {\textstyle{\frac{1}{2N+1}}}\right)}{\ln \left(1 + {\textstyle{\frac{1}{2N}}}\right)} - 1 \\ &\le \frac{\ln (\epsilon^{-1}) + \ln 10 + \ln (2N+1)}{{\textstyle{\frac{1}{2N+1}}}} - 1 \tag{since $\ln \left(1 + \frac{1}{2N}\right) \ge \frac{1}{2N+1}$} \\ &= (2N+1) \ln (\epsilon^{-1}) + (2N+1) \ln (2N+1) + (2N+1) \ln 10 - 1. \end{align*} It therefore suffices to ensure that \[ \Delta \ge \lceil (2N+1) \ln (\epsilon^{-1}) + (2N+1) \ln (2N+1) + (2N+1) \ln 10 - 1 \rceil \] in this case. It also follows from the above that \begin{align*} \zeta &\le \sum_{1 \le j \le \Delta} \left( \frac{j}{k} \right)^j \\ &\le \sum_{1 \le j \le \Delta} \left( \frac{\Delta}{k} \right)^j \\ &< \sum_{j \ge 1} \left( \frac{\Delta}{k} \right)^j \\ &= \frac{\Delta/k}{1-\Delta/k} \\ &\le \textstyle{\frac{4}{5}} \epsilon \end{align*} provided that $k \ge \lceil (\frac{5}{4} \epsilon^{-1} + 1) \Delta \rceil$, and this is the case if $k \ge \lceil (\frac{5}{4} \epsilon^{-1} + 1) (\widehat{\Delta} + 1) \rceil$, where $\Delta = \lceil \widehat{\Delta} \rceil$. Suitable choices of~$k$ for small field sizes are as shown in Figure~\ref{fig:choices_of_k} on page~\pageref{fig:choices_of_k}. \begin{figure}[t!] \begin{center} \begin{tabular}{c|c|c} $q$ & Lower Bound for $k$ & Bound when $\epsilon = \frac{1}{10}$ \\[2pt] \hline $2$ \rule{0pt}{12pt} & $\lceil \frac{51}{80} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{25}{16} \epsilon^{-1} + \frac{51}{100} \ln (\epsilon^{-1}) + \frac{5}{4} \rceil$ & $33$ \\ $3$ \rule{0pt}{12pt} & $\lceil \frac{73}{80} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{19}{8} \epsilon^{-1} + \frac{73}{100} \ln (\epsilon^{-1}) + \frac{19}{10} \rceil$ & $49$ \\ $4$ \rule{0pt}{12pt} & $\lceil \frac{11}{10} \epsilon^{-1} \ln (\epsilon^{-1}) + 3 \epsilon^{-1} + \frac{22}{25} \ln (\epsilon^{-1}) + \frac{12}{5} \rceil$ & $60$ \\ $5$ \rule{0pt}{12pt} & $\lceil \frac{51}{40} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{29}{8} \epsilon^{-1} + \frac{51}{50} \ln (\epsilon^{-1}) + \frac{29}{10} \rceil$ & $71$ \\ $7$ \rule{0pt}{12pt} & $\lceil \frac{13}{8} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{19}{4} \epsilon^{-1} + \frac{13}{10} \ln (\epsilon^{-1}) + \frac{19}{5} \rceil$ & $92$ \\ $8$ \rule{0pt}{12pt} & $\lceil \frac{29}{16} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{11}{2} \epsilon^{-1} + \frac{29}{20} \ln (\epsilon^{-1}) + \frac{22}{5} \rceil$ & $105$ \\ $9$ \rule{0pt}{12pt} & $\lceil \frac{33}{16} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{51}{8} \epsilon^{-1} + \frac{33}{20} \ln (\epsilon^{-1}) + \frac{51}{10} \rceil$ & $121$ \\ $11$ \rule{0pt}{12pt} & $\lceil \frac{179}{80} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{57}{8} \epsilon^{-1} + \frac{179}{100} \ln (\epsilon^{-1}) + \frac{57}{10} \rceil$ & $133$ \\ $13$ \rule{0pt}{12pt} & $\lceil \frac{49}{20} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{127}{16} \epsilon^{-1} + \frac{49}{25} \ln (\epsilon^{-1}) + \frac{127}{20} \rceil$& $147$ \\ $16$--$19$ \rule{0pt}{12pt} & $\lceil \frac{99}{32} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{21}{2} \epsilon^{-1} + \frac{99}{40} \ln (\epsilon^{-1}) + \frac{42}{5} \rceil$& $191$ \\ $23$--$29$ \rule{0pt}{12pt} & $\lceil \frac{87}{20} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{129}{8} \epsilon^{-1} + \frac{87}{25} \ln (\epsilon^{-1}) + \frac{129}{10} \rceil$ & $283$ \\ $31$--$43$ \rule{0pt}{12pt} & $\lceil \frac{45}{8} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{265}{12} \epsilon^{-1} + \frac{9}{2} \ln (\epsilon^{-1}) + \frac{53}{3} \rceil$ & $379$ \\ $47$--$59$ \rule{0pt}{12pt} & $\lceil \frac{65}{8} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{69}{2} \epsilon^{-1} + \frac{13}{2} \ln (\epsilon^{-1}) + \frac{138}{5} \rceil$ & $575$ \\ $61$--$71$ \rule{0pt}{12pt} & $\lceil \frac{85}{8} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{765}{16} \epsilon^{-1} + \frac{17}{2} \ln (\epsilon^{-1}) + \frac{153}{4} \rceil $ & $781$ \\ $73$--$83$ \rule{0pt}{12pt} & $\lceil \frac{105}{8} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{247}{4} \epsilon^{-1} + \frac{21}{2} \ln (\epsilon^{-1}) + \frac{247}{5} \rceil$ & $994$ \\ $\ge 89$ \rule{0pt}{12pt} & $\lceil \frac{125}{8} \epsilon^{-1} \ln (\epsilon^{-1}) + \frac{609}{8} \epsilon^{-1} + \frac{25}{2} \ln (\epsilon^{-1}) + \frac{609}{10} \rceil$& $1211$ \end{tabular} \end{center} \caption{Choices of~$k$ for Various Field Sizes} \label{fig:choices_of_k} \end{figure} When $N \ge18$, $c_4 = 2 + \frac{1}{N}$, and $\beta_0 =\frac{2}{c_4} = \frac{2N}{2N+1}$, it suffices that \begin{multline} \label{eq:asymptotic_bound_for_k} k \ge \left\lceil \left({\textstyle{\frac{5}{4}}} \epsilon^{-1} + 1\right) ((2N+1) \ln (\epsilon^{-1}) + (2N+1) \ln (2N+1) + (2N+1) \ln 10 \right\rceil \\ \in \Theta(\epsilon^{-1} N(\ln (\epsilon^{-1}) + \ln N)). \end{multline} In particular, when $\epsilon = \frac{1}{10}$, it suffices to ensure that \begin{equation} \label{eq:asymptotic_bound_for_k_with_specific_failure_probability} k \ge \left\lceil (2N+1) \ln (2N+1) = {\textstyle{\frac{167}{5}}} \ln (2N+1) \right\rceil. \end{equation} It now follows that $\rho_0 \le \frac{1}{10} \epsilon + \frac{4}{5} \epsilon = \frac{9}{10} \epsilon$ provided that this constraint on~$k$ can be satisfied. \subsection{Completion of the Analysis for the Case $\boldsymbol{q \le n^2}$} Suppose next that one wishes to ensure that $\rho_1 \le \frac{1}{20}\epsilon$, so that $\rho \le \frac{19}{20} \epsilon$. If the constraint on~$k$, described above, can be satisfied, then it follows by the inequality at line~\eqref{eq:bound_for_rho1} that it suffices to choose~$c_2$ such that $2q^{-c_2} \le \frac{\epsilon}{20}$, that is, \[ c_2 \ge \log_q (40 \epsilon^{-1}) = \frac{\ln (\epsilon^{-1}) + \ln 40}{\ln q}. \] It remains to choose $c_2 + \ell$ rows of the $(n + \ell) \times n$ matrix~$B$ uniformly and independently from $\matgrp{\ensuremath{\textup{\textsf{F}}}}{1}{n}$. Suppose that that the set of rows of the original matrix~$A$, and the $k$ ``sparse'' rows selected as described above, are linearly independent. Then it follows by Lemma~\ref{lem:dense_selection_of_vectors} that (once the remaining rows of~$B$ are chosen uniformly and independently) the rank of~$B$ is less than~$n$ with probability at most~$q^{-\ell}$. Furthermore, if $q \ge 3$ then the top $n$~rows of~$B$ are linearly independent --- so that we can set $\ell = 0$ --- with probability at least $1 - \frac{1}{q-1}$. Thus, if we wish to ensure that this probability is at most $\frac{1}{20}\epsilon$ then it suffices to ensure that either \[ \ell \ge \log_q (20 \epsilon^{-1}) = \textstyle{\frac{\ln (\epsilon^{-1}) + \ln 20}{\ln q}} \quad \text{or} \quad q \ge 20 \epsilon^{-1} + 1 \text{ (and $\ell = 0$).} \] It now follows that if $c_4$, $c_2$ and~$\ell$ have all been chosen as described above then the probability that $B$ has rank less than~$n$ is at most \[ \rho + \frac{\epsilon}{20} (1-\rho) \le \rho + \frac{\epsilon}{20} \le \frac{19 \epsilon}{20} + \frac{\epsilon}{20} = \epsilon. \] Choices of~$c_2$ and~$\ell$ satisfying the above constraints, along with the constraints at line~\eqref{eq:constraints_on_c2_and_c4}, are shown in Figure~\ref{fig:choices_of_c2_and_ell}, for the case that $\epsilon = \frac{1}{10}$. \begin{figure}[t!] \begin{center} \begin{tabular}{ccc|ccc|ccc} $q$ & $c_2$ & $\ell$ & $q$ & $c_2$ & $\ell$ & $q$ & $c_2$ & $\ell$ \\ \hline $2$ \rule{0pt}{12pt} & $9$ & $8$ & $5$ & $4$ & $4$ & $16$--$19$ & $3$ & $2$ \\ $3$ \rule{0pt}{12pt} & $6$ & $5$ & $7$ & $4$ & $3$ & $23$--$89$ & $2$ & $2$ \\ $4$ \rule{0pt}{12pt} & $5$ & $4$ & $8$--$13$ & $3$ & $3$ & & \end{tabular} \end{center} \caption{Choices of~$c_2$ and~$\ell$ for Various Field Sizes when $\epsilon = \frac{1}{10}$} \label{fig:choices_of_c2_and_ell} \end{figure} It now suffices to set $\sigma = c_3 \left( \frac{q-1}{q} \right)= 3 c_4 \left( \frac{q-1}{q} \right)$, for $c_4$ as given in Figure~\ref{fig:choices_of_c4} and $\tau = c_2 + \ell$, for $c_2$ and~$\ell$ as given in Figure~\ref{fig:choices_of_c2_and_ell}, in order to establish the claim in Theorem~\ref{thm:constants_for_conditioner} when $k = n - m - c_2$ is greater than or equal to the lower bound given in Subsection~\ref{ssec:splitting_sum} and $q \le n^2$, for $c_2$ as given above. Setting $\upsilon$ to be the sum of~$\ell$ and the lower bound for~$k$, described in Subsection~\ref{ssec:splitting_sum}, suffices to establish these claims when $n - m - c_2$ is less than the lower bound for~$k$ and $q \le n^2$ as well, for $c_2$ as above. The values shown in Figure~\ref{fig:summary_table} have been obtained using these equations. Similarly, Theorem~\ref{thm:second_constants_for_conditioner} can be obtained by setting $c_4 = 2 + \frac{1}{N}$ for $N \ge 18$, $c_3 = 3 c_4 = 6 + \frac{3}{N}$, $\sigma = c_3 (1 - \frac{1}{q})$, $c_2 = \left\lceil \frac{\ln (40 \epsilon^{-1})}{\ln q} \right\rceil$, $\ell = \left\lceil \frac{\ln 20 \epsilon^{-1}}{ln q} \right\rceil$ if $\ell \le 20 \epsilon^{-1} + 1$, setting $\ell = 0$ otherwise, setting $\tau = c_2 + \ell$, and setting $\upsilon$ to the sum of~$\ell$ and the lower bound for~$k$ described in Subsection~\ref{ssec:splitting_sum}, above. \subsection{Analysis for the Case $\boldsymbol{q > n^2}$} A slight variant of the argument from Wiedemann~\cite{wied86} can be applied when $q > n^2$. Since dense linear algebra is certainly adequate for computations on small matrices it will be assumed that $n \ge 7$, so that $q \ge 49$. Suppose, once again, that $A$ is an $m \times n$ matrix over a field~$\ensuremath{\textup{\textsf{F}}}$ that is either infinite or has size $q > n^2$. Let $\widehat{q}$ be the largest power of a prime that is less than or equal to~$n^2$. It suffices to apply the above construction, using the choices of $c_4$, $c_3$, $\beta_0$, $c_2$ and~$\ell$ appropriate for a field with size~$\widehat{q}$ (so that $\ell = 0$) --- except that, after choosing entries of rows that might be nonzero, the remaining entries of the matrix~$B$, to be filled in, should be chosen uniformly and independently from a finite subset~$S$ of~$\ensuremath{\textup{\textsf{F}}}$ with size at least $n^2$, rather than from the finite field with size~$\widehat{q}$. In order to see that this process is reliable, consider yet another matrix --- namely, a matrix~$\widehat{B}$ obtained by placing a distinct indeterminate into each row entry that is assigned a value from~$S$, above, instead of~$0$. Let us denote the indeterminate placed into the $i^{\text{th}}$ new row, in column~$j$, by $z_{i, j}$. Since $\ell = 0$ this results in an $n \times n$ matrix whose entries are elements of~$\ensuremath{\textup{\textsf{F}}}$ and indeterminates. Let $\widehat{f}$ be the determinant of this matrix --- a multivariate polynomial with total degree at most $n-m$, since only $n-m$ rows include indeterminates, and each entry of such a row has total degree at most one. Since the rows of the $m \times n$ matrix~$A$ are linearly independent, there exists a sequence of integers $i_1, i_2, \dots, i_m$ such that \[ 1 \le i_1 < i_2 < \dots < i_m \le n \] and such that the $m \times m$ submatrix, including columns $i_1, i_2, \dots, i_m$, is nonsingular. Permuting rows of~$A$ as needed, we may assume without loss of generality that the entry of the $j^{\text{th}}$ row of~$A$ in column~$i_j$ is nonzero, for $1 \le j \le m$. Consequently, if the $j^{\text{th}}$ row of~$A$ was replaced by a row whose $i_j^{\text{th}}$ entry is~$1$ and whose other entries are~$0$, this would result in an $n \times m$ matrix~$\widehat{A}$ whose rows are linearly independent as well. Indeed, the $m \times m$ submatrix including the entries in columns $i_1, i_2,\dots, i_m$ would have determinant~$1$. Similarly, the rows of this matrix are linearly independent when the entries are viewed as elements of the finite field~$\ensuremath{\textup{\textsf{F}}}_{\widehat{q}}$ with size~$\widehat{q}$, instead of as elements of~$\ensuremath{\textup{\textsf{F}}}$. Let us call this matrix (an $m \times n$ matrix with entries in~$\ensuremath{\textup{\textsf{F}}}_{\widehat{q}}$) $\widetilde{A}$. Note that the process, described above, to produce new rows to obtain~$B$ from~$A$, is independent of the entries in the rows of~$A$ --- it only depends on the number~$m$ of rows and~$n$ of columns of~$A$. With that noted, let us consider yet another $n \times n$ matrix, namely a matrix with entries in~$F_{\widehat{q}}$ whose first $m$ rows are the rows of~$\widetilde{A}$ and whose remaining rows are produced by initially deciding to set the same entries of rows to~$0$ as for the new rows of~$B$, and whose remaining entries are chosen uniformly and independently from~$\ensuremath{\textup{\textsf{F}}}_{\widehat{q}}$. It follows by the analysis for the case $q \le n^2$ that this matrix is nonsingular with some probability $\sigma \ge \frac{8}{9}$. Let us suppose that this is the case. Then there must exist a set of column indices $i_{m+1}, i_{m+2}, \dots, i_n$ such that \[ 1 \le i_{m_1} < i_{m+2}, \dots, i_n \le n, \quad \{ i_1, i_2, \dots, i_m \} \cup \{i_{m+1}, i_{m+2}, \dots, i_n \} = \{1, 2, \dots, n \}, \] and the entries of~$\widetilde{A}$ in row~$j$ and column~$i_j$ are all nonzero, for $1 \le j \le n$. Consequently if $m+1 \le j \le n$ then the entry of~$\widehat{B}$ in row~$j$ and column~$i_j$ is an indeterminate, $z_{j-m, i_j}$, rather than zero. It now follows that the above polynomial~$\widehat{f}$ is not identically zero (in this case): For if one sets the value of each indeterminate $z_{j-m, i_j}$ to be~$1$ and one sets the value of all other indeterminates to be~$0$, then the value of this polynomial is the product of $\pm 1$ and the determinant of the submatrix of~$A$ including the entries in columns $i_1, i_2, \dots, i_m$ --- which is nonsingular, as noted above. It now follows by an application of the Schwartz-Zippel lemma~\cite{schwar80, zipp79} that the above matrix~$B$ is singular, in this particular case, with probability at most $\frac{n-m}{|S|} \le \frac{n}{|S|} \le \frac{1}{n}$. It follows that the overall probability that $B$ is singular is at most \[ (1 - \sigma) + \textstyle{\frac{\sigma}{n}} \le \textstyle{\frac{1}{9} + \frac{8}{9n}}, \] as needed to complete the proofs of Theorems~\ref{thm:constants_for_conditioner} and~\ref{thm:second_constants_for_conditioner}. \bibliographystyle{plain}
2c44af72a8794995c64512ad5fd172dd7402e050
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{A numerical example} \label{Sec-6} For the sake of the discussion in this section, we adopt the view of the Financial Stability Board and the International Monetary Fund that the systemic risk can be caused by impairment of all or parts of the financial system, and more formally, we call the risk factor $j\in\{1,\ldots,l+m\}$ `systemic', if $c_{i,j}=1$ for at least two distinct r.c.'s $i\in\{1,\ldots,n\}$. Also, we call the risk factor $j\in\{1,\ldots,l+m\}$ `idiosyncratic', if $c_{i,j}=1$ for only one risk component $i\in\{1,\ldots,n\}$. Consider obligors in a default risk portfolio, each of which is exposed to exactly three distinct categories of fatal risk factors, e.g., systemic with fully-commonotonic occurrence times of the r.f.'s (category A), systemic with conditionally independent occurrence times of the r.f.'s (category B) and idiosyncratic with independent occurrence times of the r.f.'s (category C). We assume that the risk factors from distinct risk categories are independent and that the hitting times (or occurrences) of defaults of the r.c.'s are exponentially-distributed with random parameters distributed gamma. In fact, the future lifetime r.v. of the $i$-th r.c., $i=1,\ldots,n$ has exponential distribution with the random parameter $\sigma_i^{-1}\sum_{j=1}^{l+m}c_{i,j}\Lambda_j$, where $\Lambda_j$ are distributed gamma with unit rate parameters. Then Definition \ref{MRF-stoch-Def} readily implies that the joint default times of the aforementioned r.c.'s has d.d.f. (\ref{ddf-gp}). To illustrate the effect of the dependence structure on the joint default probability we further set the dimension to $n=2$ and specialize the set-up above along the lines in Section 16.8 of Engelmann and Rauhmeier (2011) as well as employing the 2014's Annual Global Corporate Default Study and Rating Transitions of Standard \& Poor's (Standard \& Poor's (2015)). More specifically, we assume the existence of six r.f.'s and set each $\mu:=\mathbf{E}[\Lambda_j]\equiv 1/1.8$, then fix the time horizon to $15$ years and choose the corresponding default probability, $p$ say, to be equal to $0.3198$ (on par with the `B' credit rating of highly speculative entities). This yields the multivariate probability structure of Definition \ref{MRF-stoch-Def} with identically distributed margins having the parameters $\sigma_i\equiv \sigma=122.39$ and $\xi_{c,i}\equiv 3.33$, for $i=1$ and $2$. Then we explore three different exposures of the obligors to the systemic and idiosyncratic r.f.'s. The distinct exposures are stipulated by appropriate choices of the $c$ parameters gathered by matrices $A_c^{(k)},\ k=1,\ 2,\ 3$. We compare the aforementioned three exposures with the reference case in which no systemic risk presents, that is the joint d.d.f. of default times is a bivariate Pareto with independent margins. We note in passing that the expressions for the d.d.f.'s below readily follow from Theorem \ref{def-gp}, whereas the values of the Pearson correlation coefficient are in non-trivial cases obtained with the help of Theorem \ref{joint_m}. \begin{itemize} \item [Case (1).] Only the systemic (category A) and idiosyncratic (category C) risks present. The exposure is represented schematically with the use of the following matrix, in which the rows and the columns represent r.c.'s and r.f.'s, respectively \[ A_c^{(1)}=\left( \begin{array}{cccc||cccc} 1 & 1 & 1 & 1 & 1 & 1 &0 & 0\\ 1 & 1 & 1 & 1 & 0 & 0& 1 & 1\\ \end{array} \right). \] The joint d.d.f. of the risk components is given by \[ \overline{F}^{(1)}(x_1,x_2)=\left( 1+\frac{\max(x_1,x_2)}{\sigma} \right)^{-4\mu} \left( 1+\frac{x_1}{\sigma} \right)^{-2\mu}\left( 1+\frac{x_2}{\sigma} \right)^{-2\mu} , \] where $x_1,x_2$ are all in $\mathbf{R}_{+}$. This is obviously the d.d.f. of the bivariate Pareto distribution of Asimit et al. (2010). In this r.p., the Pearson correlation coefficient between the r.c.'s is $0.36$. \end{itemize} \noindent \begin{itemize} \item [Case (2).] There are four conditionally independent r.f.'s (category B) and two uncorrelated idiosyncratic r.f.'s (category C). The exposure is gathered by the following block matrix \[ A_c^{(2)}=\left( \begin{array}{c|cccc|cccc} ~ & 1 & 1 & 1 & 1 & 1 & 1 &0 & 0\\ ~ & 1 & 1 & 1 & 1 & 0 & 0& 1 & 1\\ \end{array} \right). \] The joint d.d.f. of the risk components is given by \begin{eqnarray*} &&\overline{F}^{(2)}(x_1,x_2)\\ &&=\left( 1+\frac{x_1+x_2}{\sigma} \right)^{-4\mu} \left( 1+\frac{x_1}{\sigma} \right)^{-2\mu} \left( 1+\frac{x_2}{\sigma} \right)^{-2\mu}, \end{eqnarray*} where $x_1,x_2$ are all in $\mathbf{R}_{+}$. This case corresponds to the bivariate Pareto model of Su \& Furman (2016). In this r.p., the Pearson correlation coefficient between the r.c.'s is $0.14$. \item [Case (3).] The r.p. admits the most general form that is proposed in the current paper. Namely r.f.'s from all three categories (A, B and C) present. The exposure block matrix is given by \[ A_c^{(3)}=\left( \begin{array}{cc|cc|ccccc} 1 & 1 & 1 & 1 & 1 & 1 &0 & 0\\ 1 & 1 & 1 & 1 & 0 & 0& 1 & 1\\ \end{array} \right). \] The joint d.d.f. of the risk components is \[ \overline{F}^{(3)}(x_1,x_2)=\left( 1+\frac{\max(x_1,x_2)}{\sigma}\right)^{-2\mu} \left( 1+\frac{x_1+x_2}{\sigma}\right)^{-2\mu} \left( 1+\frac{x_1}{\sigma} \right)^{-2\mu}, \left( 1+\frac{x_2}{\sigma} \right)^{-2\mu}, \] where $x_1,x_2$ are all in $\mathbf{R}_{+}$. In this r.p., the Pearson correlation coefficient between the r.c.'s is equal to $0.23$. \end{itemize} \subsection{Expected times of the first default} The left panel of Figure \ref{fig:CTEmin} depicts the values of $CTE_q[X_-]$ for $q\in[0,\ 1)$, $X_-\in\mathcal{X}$ and portfolios (1) to (3) as well as the reference portfolio, denoted by $(\perp)$. As the risk components are identically distributed, it is not difficult to see that the following ordering holds \begin{equation} \label{stord1} \overline{F}^{(1)}_-\geq_{st}\overline{F}^{(3)}_-\geq_{st}\overline{F}^{(2)}_-\geq_{st} \overline{F}^{(\perp)}_-, \end{equation} where `$\geq_{st}$' denotes first order stochastic dominance (FSD). Furthermore, since the CTE risk measure is known to preserve the FSD ordering, we also have that \[ CTE^{(1)}_q[X_-]\geq CTE^{(3)}_q[X_-]\geq CTE^{(2)}_q[X_-]\geq CTE^{(\perp)}_q[X_-] \] for all $q\in[0,\ 1)$ and $X_-\in\mathcal{X}$. This conforms to Figure \ref{fig:CTEmin} (left panel), which hints that the r.p.'s with more significantly correlated r.c.'s enjoy higher, and thus more favourable, occurrence times of the first default. \begin{figure}[h!] \centering \includegraphics[width=7cm, height=7cm]{cte_min.eps} \includegraphics[width=7cm, height=7cm]{cte_max.eps}\\ \caption{Conditional expected times of first (left panel) and last (right panel) default for portfolios (1) - (3) and the reference portfolio $(\perp)$ for `B' rating r.p.'s with the probability of default $p=0.3198$ and $\mu=1/1.8$. Proposition \ref{CTEparMin} is employed to compute the values of $CTE_q$ for $q\in[0,\ 1)$.} \label{fig:CTEmin} \end{figure} The downside of high correlations is elucidated in Figure \ref{fig:minCase}, in which we leave the probability of default $p$ to be equal to $0.3198$ (`B' rating), but vary the $\mu$ parameter that stipulates the effect of the risk factors. In this respect, we observe that the r.p.'s with stronger correlations between r.c.'s are more sensitive to the changes in the $\mu$ parameter, and therefore such r.p.'s must be monitored and stress-tested more frequently. \begin{figure}[h!] \centering \includegraphics[width=7cm,height=7cm]{min_c1.eps} \includegraphics[width=7cm,height=7cm]{min_c2.eps}\\ \includegraphics[width=7cm,height=7cm]{min_c3.eps} \includegraphics[width=7cm,height=7cm]{min_ind.eps} \caption{Conditional expected times of first default for portfolios (1) (top left panel), (2) (top right panel), (3) (bottom left panel) and reference $(\perp)$ (bottom right panel) with the parameter $\mu$ varying from $1/1.8$ to $1/2.6$ and the default probability $p=0.3198$. Proposition \ref{CTEparMin} is employed to compute the values of $CTE_q$ for $q\in[0,\ 1)$.} \label{fig:minCase} \end{figure} \subsection{Expected times of the last default} Figure \ref{fig:CTEmin} (right panel) depicts the values of $CTE_q[X_+]$ for $q\in[0,\ 1),$ $X_+\in\mathcal{X}$ and portfolios (1) to (3) as well as the reference portfolio $(\perp)$. We have that \[ \overline{F}^{(\perp)}_+\geq_{st}\overline{F}^{(2)}_+\geq_{st} \overline{F}^{(3)}_+ \geq_{st} \overline{F}^{(1)}_+, \] and hence \[ CTE^{(\perp)}_q[X_+]\geq CTE^{(2)}_q[X_+]\geq CTE^{(3)}_q[X_+] \geq CTE^{(1)}_q[X_+] \] for all $q\in[0,\ 1)$ and $X_+\in\mathcal{X}$. This conforms with the right panel of Figure \ref{fig:CTEmin}. Unlike in the case of the first default, we observe that if the time of the last default is of interest and the distributions of the r.c.'s are fixed, then assuming stronger correlations between r.c.'s yields a more conservative assessment of the expected time of the last default. \subsection{Solvency bonus indices} Figure \ref{fig:econCTE} depicts the solvency bonus indices for portfolios (1) to (3). As expected, stronger dependencies between default times are associated with higher values of $\beta$. \begin{figure}[h!] \centering \includegraphics[width=7cm,height=7cm]{cond_cte.eps} \includegraphics[width=7cm,height=7cm]{con_c1.eps}\\ \includegraphics[width=7cm,height=7cm]{con_c2.eps} \includegraphics[width=7cm,height=7cm]{con_c3.eps} \caption{Comparison of solvency bonus indices for portfolios (1) to (3) (top left panel), as well as the solvency bonus indices solely for portfolio (1) (top right panel), (2) (bottom left panel) and (3) (bottom right panel) with the parameter $\mu$ varying from $1/1.8$ to $1/2.6$ and the default probability $p=0.3198$. Proposition \ref{survival-function} is used to compute the values of the solvency bonus indices.} \label{fig:econCTE} \end{figure} \section{Conclusions} \label{Sec-7} The latest Solvency II directives require the insurers to recognize the interdependencies within and among different liability classes. As such, the new paradigm arguably brings to an end the indisputable role of the assumption of independence that has shaped both research and applications in actuarial science and general quantitative risk management in the 20th century. The choice of an appropriate probability dependence model is however not an easy call. Indeed, while there exists only one way to describe stochastic independence, the forms of stochastic dependence are infinite. In this paper we have introduced a new class of Multiple Risk Factor dependence structures. On the one hand, these structures emerge as an extension of the popular CreditRisk$^+$ approach, and as such they formally describe default risk portfolios exposed to an arbitrary number of fatal risk factors with conditionally exponential hitting times that can be independent, positively dependent and even fully comonotonic. On the other hand, the MRF structures can be viewed as a quite general family of multivariate probability distributions with Pareto of the 2nd kind univariate margins, and in this role, they can model risk portfolios of (insurance) losses with heavy tailed and positively dependent risk components. It often happens in mathematical sciences that generalizing an object highlights the essence of the matter and helps to understand it better. By generalizing the classical multivariate Pareto distribution of Arnold (1983, 2015), among others, in this paper we have extended the range of the attainable Pearson correlations to the entire $[0,\ 1]$ interval and complemented some formal dependence analysis. As factor models have been known to produce under-correlated default times, the MRF structures introduced and studied herein may provide a possible route to `inject the required amounts of extra correlation', and they may thereby become of interest to banks, credit unions and insurance companies. In addition, as the new realms of excessively prudent risk management make particular effort to model non-hedgeable and heavy tailed risks, the family of MRF multivariate distributions may be of interest well beyond the context of credit risk. \section*{Acknowledgements} We are grateful to Prof. Dr. Paul Embrechts and all participants of the ETHs Series of Talks in Financial and Insurance Mathematics for feedback and insights. \section*{Disclosure statement} No potential conflict of interest was reported by the authors. \section*{Funding} Our research has been supported by the Natural Sciences and Engineering Research Council (NSERC) of Canada. Jianxi Su also acknowledges the financial support of the Government of Ontario and MITACS Canada via, respectively, the Ontario Graduate Scholarship program and the Elevate Postdoctoral fellowship. \section*{References} \hangindent=\parindent\noindent {Adalsteinsson, G. (2014).} \textit{The Liquidity Risk Management Guide: From Policy to Pitfalls.} Wiley, Chichester. \hangindent=\parindent\noindent {Arnold, B. C. (1983).} \textit{Pareto Distributions.} International Cooperative Publishing House, Fairland. \hangindent=\parindent\noindent {Arnold, B. C. (2015).} \textit{Pareto Distributions}, 2nd ed. CRC Press, Boca Raton. \hangindent=\parindent\noindent Asimit, A. V., Furman, E. \& Vernic, R. (2010). On a multivariate Pareto distribution. \textit{Insurance: Mathematics and Economics} \textbf{46}(2), 308--316. \hangindent=\parindent\noindent Asimit, A. V., Furman, E. \& Vernic, R. (2016). Statistical inference for a new class of multivariate Pareto distributions. \textit{Communications in Statistics - Simulation and Computation} \textbf{45}(2), 456--471. \hangindent=\parindent\noindent Azizpour, S. \& Giesecke, K. (2008). {Self-exciting corporate defaults : Contagion vs. frailty.} Technical report, Stanford University, Stanford. \hangindent=\parindent\noindent Benson, D. A., Schumer, R. \& Meerschaert, M. M. (2007). Recurrence of extreme events with power-law interarrival times. \textit{Geophysical Research Letters} \textbf{34}, 1--5. \hangindent=\parindent\noindent Bowers, N. L., Gerber, H. U., Hickman, J. C., Jones, D. A. \& Nesbitt, C. J. (1997). \textit{Actuarial Mathematics}, 2nd ed. Society of Actuaries, Schaumburg. \hangindent=\parindent\noindent Bunke, H. C. (1969). \textit{A Primer on American Economic History.} Random House, New York. \hangindent=\parindent\noindent Cebri\'{a}n, A. C., Denuit, M. \& Lambert, P. (2003). Generalized Pareto fit to the Society of Actuaries' large claims database. \textit{North American Actuarial Journal} \textbf{7}(3), 18--36. \hangindent=\parindent\noindent Chavez-Demoulin, V., Embrechts, P. \& Hofert, M. (2015). An extreme value approach for modeling operational risk losses depending on covariates. \textit{Journal of Risk and Insurance}, in press. \hangindent=\parindent\noindent Chiragiev, A. \& Landsman, Z. (2009). Multivariate flexible Pareto model: Dependency structure, properties and characterizations. \textit{Statistics and Probability Letters} \textbf{79}(16), 1733--1743. \hangindent=\parindent\noindent Das, S. R., Duffie, D., Kapadia, N. \& Saita, L. (2007). Common failings: How corporate defaults are correlated. \textit{Journal of Finance} \textbf{62}(1), 93--117. \hangindent=\parindent\noindent Denuit, M., Dhaene, J., Goovaerts, M. \& Kaas, R. (2005). \textit{Actuarial Theory for Dependent Risk: Measures, Orders and Models.} Wiley, Chichester. \hangindent=\parindent\noindent Duffie, D. \& Singleton, K. J. (1999). Modeling term structures of defaultable bonds. \textit{Review of Financial Studies} \textbf{12}(4), 687--720. \hangindent=\parindent\noindent Engelmann, B. \& Rauhmeier, R. (2011). \textit{The Basel II Risk Parameters: Estimation, Validation, Stress Testing - with Applications to Loan Risk Management.} Springer, Berlin. \hangindent=\parindent\noindent Feller, W. (1966). \textit{An Introduction to Probability Theory and Its Applications.} Wiley, New York. \hangindent=\parindent\noindent Frey, R. \& McNeil, A. J. (2003). Dependent defaults in models of portfolio credit risk. \textit{Journal of Risk} \textbf{6}(1), 59--92. \hangindent=\parindent\noindent Furman, E. \& Landsman, Z. (2005). Risk capital decomposition for a multivariate dependent gamma portfolio. \textit{Insurance: Mathematics and Economics} \textbf{37}(3), 635--649. \hangindent=\parindent\noindent Furman, E. \& Zitikis, R. (2008a). Weighted premium calculation principles. \textit{Insurance: Mathematics and Economics} \textbf{42}(1), 459--465. \hangindent=\parindent\noindent Furman, E. \& Zitikis, R. (2008b). Weighted risk capital allocations. \textit{Insurance: Mathematics and Economics} \textbf{43}(2), 263--269. \hangindent=\parindent\noindent Gabaix, X., Gopikrishnan, P., Plerou, V. \& Stanley, H. E. (2003). A theory of power-law distributions in financial market fluctuations. \textit{Nature} \textbf{423}, 267--270. \hangindent=\parindent\noindent Geweke, J. \& Amisano, G. (2011). Hierarchical Markov normal mixture models with applications to financial asset returns. \textit{Journal of Applied Econometrics} \textbf{26}(1), 1--29. \hangindent=\parindent\noindent Giesecke, K. (2003). A simple exponential model for dependent defaults. \textit{The Journal of Fixed Income} \textbf{13}(3), 74--83. \hangindent=\parindent\noindent Gordy, M. B. (2000). A comparative anatomy of credit risk models. \textit{Journal of Banking and Finance} \textbf{24}(1), 119--149. \hangindent=\parindent\noindent Gradshteyn, I. S. \& Ryzhik, I. M. (2014). \textit{Table of Integrals, Series, and Products}, 8th ed. Academic Press, New York. \hangindent=\parindent\noindent Joe, H. (1997). \textit{Multivariate Models and Dependence Concepts.} CRC Press, Boca Raton. \hangindent=\parindent\noindent Koedijk, K. G., Schafgans, M. M. A. \& de Vries, C. G. (1990). The tail index of exchange rate returns. \textit{Journal of International Economics} \textbf{29}(1-2), 93--108. \hangindent=\parindent\noindent Lehmann, E. L. (1966). Some concepts of dependence. \textit{The Annals of Mathematical Statistics} \textbf{37}(5), 1137--1153. \hangindent=\parindent\noindent Longin, F. M. (1996). The asymptotic distribution of extreme stock market returns. \textit{Journal of Business} \textbf{69}(3), 383--408. \hangindent=\parindent\noindent Marshall, A. W. \& Olkin, I. (1967). A generalized bivariate exponential distribution. \textit{Journal of Applied Probability} \textbf{4}(2), 291--302. \hangindent=\parindent\noindent McNeil, A. J., Frey, R. \& Embrechts, P. (2005). \textit{Quantitative Risk Management: Concepts, Techniques, and Tools.} Princeton University Press, Princeton. \hangindent=\parindent\noindent Moschopoulos, P. G. (1985). The distribution of the sum of independent gamma random variables. \textit{Annals of the Institute of Statistical Mathematics} \textbf{37}(1), 541--544. \hangindent=\parindent\noindent Nelsen, R. B. (2006). \textit{An Introduction to Copulas}, 2nd ed. Springer, New York. \hangindent=\parindent\noindent Soprano, A., Crielaard, B., Piacenza, F. \& Ruspantini, D. (2010). \textit{Measuring Operational and Reputational Risk: A Practitioner's Approach.} Wiley, Chichester. \hangindent=\parindent\noindent Standard \& Poor's (2015). Default, transition and recovery: 2014 annual global corporate default study and rating transitions. Technical report, Standard \& Poor's, New York. \hangindent=\parindent\noindent Su, J. \& Furman, E. (2016). A form of multivariate Pareto distribution with applications to financial risk measurement. \textit{ASTIN Bulletin: The Journal of the International Actuarial Association}, in press. \hangindent=\parindent\noindent Sweeting, P. (2011). \textit{Financial Enterprise Risk Management.} Cambridge University Press, Cambridge. \hangindent=\parindent\noindent Vernic, R. (2011). Tail conditional expectation for the multivariate Pareto distribution of the second kind: Another approach. \textit{Methodology and Computing in Applied Probability} \textbf{13}(1), 121--137. \hangindent=\parindent\noindent Wang, S. (1996). Premium calculation by transforming the layer premium density. \textit{ASTIN Bulletin: The Journal of the International Actuarial Association} \textbf{26}(1), 71--92.
b1d4923264d4447fa4f7879042a9dcf7f6f72f27
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section*{Introduction} Let $\mathcal{P}\subset\mathbb{R}^d$ be a $d$-dimensional lattice polytope, that is, a convex polytope all of whose vertices belong to $\mathbb{Z}^d$. Let $\operatorname{Vol}(\mathcal{P})$ denote the \emph{normalized volume} of $\mathcal{P}$, which is $d!$ times the Euclidean volume (Lebesgue measure) of $\mathcal{P}$. For $k\in \mathbb{Z}_{>0}$, the \emph{lattice point enumerator} $i(\mathcal{P},k)$ counts the number of lattice points in $k\mathcal{P}=\{k\alpha:\alpha\in \mathcal{P}\}$, the $k$th dilation of $\mathcal{P}$. That is, \[ i(\mathcal{P},k)=\#(k\mathcal{P}\cap \mathbb{Z}^d), \ \ k\in\mathbb{Z}_{>0}. \] Provided that $\mathcal{P}$ is a lattice polytope, it is known that $i(\mathcal{P},k)$ is a polynomial in the variable $k$ of degree $d$ (\cite{Ehrhart}). The \emph{Ehrhart Series} for $\mathcal{P}$, $\Ehr{\mathcal{P}}{z}$, is the rational generating function \[ \Ehr{\mathcal{P}}z= 1+\sum_{k\geq 1}i(\mathcal{P},k)z^k=\frac{\delta(\mathcal{P},z)}{(1-z)^{d+1}} \] where $\delta(\mathcal{P},z)=1+\delta_1z+\delta_2z^2+\cdots+\delta_dz^d$ is the \emph{$\delta$-polynomial} of $\mathcal{P}$ (cf. \cite[Chapter 9]{ACCP}). The $\delta$-polynomial is endowed with the following properties: \begin{itemize} \item $\delta_0=1$, $\delta_1=i(\mathcal{P},1)-(d+1)$, and $\delta_d=\#(\mathcal{P}\setminus \partial\mathcal{P}\cap \mathbb{Z}^d)$; \item $\delta_i\geq 0$ for all $0\leq i\leq d$ (\cite{StanleyNonNeg}); \item If $\delta_d\neq 0$, then $\delta_1\leq \delta_i$ for each $0\leq i\leq d-1$ (\cite{HibiBounds}). \end{itemize} For proofs of the first three properties of the coefficients, the reader should consult \cite[Chapter 9]{ACCP} or \cite[Chapter 3]{BeckRobins}. The Ehrhart series and $\delta$-polynomials for polytopes have been studied extensively. For a detailed background on these topics, please refer to \cite{BeckRobins,Ehrhart,ACCP,EC1}. Given two polytopes $\mathcal{P}_1$ and $\mathcal{P}_2$ in $\mathbb{R}^d$, we say that $\mathcal{P}_1$ and $\mathcal{P}_2$ are \emph{unimodularly equivalent} if there exists a unimodular matrix $U\in\mathbb{Z}^{d\times d}$ (i.e. $\det(U)=\pm 1$) and an integral vector ${\boldsymbol v} \in \mathbb{Z}^d$, such that $\mathcal{P}_2=f_U(\mathcal{P}_1)+{\boldsymbol v}$, where $f_U$ is the linear transformation defined by $U$, i.e., $f_U({\bf v}) = {\bf v} U$ for all ${\bf v} \in \mathbb{R}^d$. . We write $\mathcal{P}_1\cong \mathcal{P}_2$ in the case of unimodular equivalence. It is clear that if $\mathcal{P}_1\cong\mathcal{P}_2$, then $\delta(\mathcal{P}_1,z)=\delta(\mathcal{P}_2,z)$. We say that a lattice polytope $\mathcal{P}$ is \emph{reflexive} if the origin is the unique interior lattice point of $\mathcal{P}$ and its dual polytope \[ \mathcal{P}^\vee=\left\{y\in\mathbb{R}^d \ : \langle x,y \rangle\leq 1 \mbox{ for all } x\in\mathcal{P} \right\} \] is a lattice polytope. Moreover, it follows from \cite{HibiDualPolytopes} that the following statements are equivalent: \begin{itemize} \item $\mathcal{P}$ is unimodularly equivalent to some reflexive polytope; \item $\delta(\mathcal{P},z)$ is of degree $d$ and is symmetric, that is $\delta_i=\delta_{d-i}$ for $0\leq i \leq \lfloor \frac{d}{2} \rfloor$. \end{itemize} A polytope $\mathcal{P}$ is called \emph{self dual} if $\mathcal{P}$ is unimodularly equivalent to its dual polytope $\mathcal{P}^\vee$. This is an extremely rare property in reflexive polytopes, especially for reflexive simplices. There are two families known self dual reflexive simplices. The first such family is given in \cite{BNill} and the second family is given in \cite{TsuchiyaDeltaVector}. A construction for self dual reflexive polytopes is given in \cite{TsuchiyaDeltaVector}, though these polytopes are not simplicial and hence not simplices. In this paper, we provide a new family of self dual reflexive simplices $Q_n$ with small volume. We now define a family of reflexive simplices which are self dual. For $n \geq 2$, let $Q_n$ denote the $n-1$ dimensional simplex with $\mathcal{V}$-representation \[ Q_n:=\mathrm{conv} \begin{bmatrix} 1 & 1-n & 0 & 0 & \cdots & 0\\ 1 & 1 & 2-n & 0 & \cdots & 0\\ 1 & 1 & 1 & 3-n & \cdots & 0\\ \vdots & \vdots & \vdots & & \ddots & \vdots\\ 1 & 1 & 1 & \cdots & 1 & -1\\ \end{bmatrix} \] where we use the convention that the $x_{n-1}$ coordinate is given by the first row and the $x_1$ is given by the last row. We will adopt for simplicity the notation $C_{i}$ where $i\in\{0,\cdots,n-1\}$ for each column (vertex) such that $Q_n=\mathrm{conv}[C_{0} \ C_{1} \ \cdots \ C_{n-1}]$. We have the following theorem. \begin{theorem}\label{selfdual} For $n\geq 2$, we have $Q_n\cong Q_n^\vee$. \end{theorem} It behooves us to introduce an $\mathcal{H}$-representation for the simplex to compute its dual polytope. We now give such a representation. \begin{proposition}\label{hrepQn} For $n \geq 2$, $Q_n$ has the $\mathcal{H}$-representation \[ Q_n=\left. \begin{cases} x\in \mathbb{R}^{n-1} \ : & kx_k-\sum_{i=1}^{k-1}x_i\leq 1 \ , \ \ 1\leq k \leq n-1\\ & -\sum_{i=1}^{n-1}x_i\leq 1\\ \end{cases} \right\}. \] \end{proposition} \begin{proof} It is sufficient to show that the vertices of $Q_n$ each satisfy precisely $n-1$ of the halfspace inequalities with equality and satisfies the other inequalty strictly. Let $f_k({\boldsymbol x})=kx_k-\sum_{i=1}^{k-1}x_i$, and $f_{n}({\boldsymbol x})=-\sum_{i=1}^{n-1}x_i$. For a vertex $C_j$, we have that $f_k(C_j)= 1$ for all $k\neq n-j$. This follows, because if $k<n-j$, we have $f_k(C_j)=(k)(1)-\sum_{i=1}^{k-1}1=1$, if $k>n-j$ with $k\neq n$, we have $f_k(C_j)=(n-j)-\sum_{i=1}^{n-j-1}1=1$, and if $k=n>n-j$, we have $f_n(C_j)= -(j-n)-\sum_{i=1}^{n-j-1}1=1$. In the case of $k=n-j$, $f_{n-j}(C_j)=-(n-j)^2-(n-1-j)<1$ if $j\neq 0$ and $j\neq n-1$. For $j=0$ we have $f_n(C_0)=1-n<1$ and for $j=n-1$, we have $f_1(C_{n-1})=-1<1$. Thus, we have the correct $\mathcal{H}$-representation. \end{proof} By \cite[Corollary 35.3]{ACCP}, and Proposition \ref{hrepQn}, it is clear that $Q_n^\vee =-Q_n$. Therefore, we have shown Theorem \ref{selfdual}. \begin{remark} We should note that $\operatorname{Vol}(Q_n)=n!$. For $n\geq 4$, it is immediate that these polytopes are different than previously known self dual reflexive simplices given in \cite{BNill,TsuchiyaDeltaVector}. \end{remark} Moreover, the self dual reflexive simplex of $Q_n$ has an interesting $\delta$-polynomial and a special triangulation \begin{theorem}\label{ehrhartseries} Let $n \geq 2$.\\ \textnormal{(i)} We have $\delta(Q_n,z)=A_n(z),$ where $A_n(z)$ is the Eulerian polynomial.\\ \textnormal{(ii)} $Q_n$ has a regular, flag, unimodular triangulation. \end{theorem} \begin{proof} For a lattice polytope $\mathcal{P} \subset \mathbb{R}^d$, we set $\Pyr{\mathcal{P}}=\mathrm{conv}(\mathcal{P}\times\{0\}, (0,\cdots,0,1)) \subset \mathbb{R}^{d+1}$. Then it is well-known that $\delta(\Pyr{\mathcal{P}},z)=\delta(\mathcal{P},z)$ (cf. \cite[Section 2.4]{BeckRobins}) and $\mathcal{P}$ has a regular, flag, unimodular triangulation if and only if $\Pyr{\mathcal{P}}$ has a regular, flag, unimodular triangulation (cf. \cite[Section 4.2]{Triangulations}). Let $R_n$ denote the $n$ dimensional simplex with $\mathcal{V}$-representation \[ R_n:=\mathrm{conv} \begin{bmatrix} 0 & n & n & n & \cdots & n\\ 0 & 0 & n-1 & n-1 & \cdots & n-1\\ 0 & 0 & 0 & n-2 & \cdots & n-2\\ \vdots & \vdots & \vdots & & \ddots & \vdots\\ 0 & 0 & 0 & \cdots & 0 & 1\\ \end{bmatrix} \subset \mathbb{R}^n. \] This polytope $R_n$ is called a \emph{lecture hall polytope}. Notice that $\Pyr{Q_n}$ is unimodularly equivalent to $R_n$. Let $\widetilde{R_n}$ be the polytope defined from $R_n$ by removing the $(n+1)$th column and $n$th row, let $\mathcal{U}_{n}$ denote the $(n-1)\times(n-1)$ upper triangular matrix defined by $(\mathcal{U}_{n})_{ij}=1$ if $i\leq j$ and $(\mathcal{U}_{n})_{ij}=0$ otherwise, let $\mathbf{1}$ denote the all ones vector, and let $\mathbf{0}$ denote the all zeros vector. Then we have \[ Q_n\cong -f_{\mathcal{U}_{n}}\left( Q_n-\mathbf{1}\right)=\widetilde{R_n}. \] Hence it follows that \[ \Pyr{Q_n}\cong \operatorname{Pyr}\left({\widetilde{R_n}}\right)\cong R_n. \] It is known that for $n \geq 2$, $\delta(R_n,z)=A_n(z)$ (\cite{SavageSchuster}) and $R_n$ has a regular, flag, unimodular triangulation (\cite{BBKSZ}). Therefore, the assertion follows. \end{proof}
f7ab4dee5a708f272b611f41f4c564506e294d73
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Wall-bounded turbulent shear flows are of general interest in many engineering applications. Three-dimensional (3D) turbulent channel flow bounded by two parallel walls is one of the canonical flow considered for direct numerical simulation (DNS). Starting with the seminal work of Kim {\it et al}.~\cite{KMM}, many DNSs have been performed for increasingly higher Reynolds number, taking advantage of the growing power of supercomputers (see, e.g., a review article \cite{Smits}). Currently the DNS with the highest friction based Reynolds number, $Re_\tau$, of $5200$ has been carried out by Lee and Moser~\cite{Lee}. Turbulent flows are typically characterized by the excitation of a multitude of spatial and temporal scales, which involves a large number of degrees of freedom interacting nonlinearly. Self-organization of the flow into coherent vortices is observed, even at large Reynolds number~\cite{BR}, where one observes that these vortices are superimposed to a random background flow~\cite{She}. Moreover, turbulence exhibits significant spatial and temporal intermittency, especially in the dissipative range. This implies that the strongest contributions become sparser and sparser while going to small scale in space and time. Wavelets being well-localized functions in space and in scale, they yield efficient multi-scale decompositions, which have been applied to analyze, model and compute turbulent flows since 1988~\cite{FR,Mene91,Farge92,SV10}. Decomposing turbulent flows into a wavelet basis yields a sparse representation, namely the most energetic contributions are concentrated in few wavelet coefficients having strong intensity, while the large majority of the remaining wavelet coefficients have negligibly small intensity. The presence of coherent structures superimposed to a random background flow motivated the development of the coherent vorticity extraction (CVE) method. The idea of CVE, proposed by Farge {\it et al}.~\cite{FaPeSc2001,FaScKe99}, defines coherent structures as what remains after denoising the flow vorticity. Vorticity is better localized in space than velocity, and thus more intermittent, its wavelet decomposition is sparser and only few coefficients are necessary to represent the coherent structures. The main reason is that, in contrast to the velocity, vorticity preserves Galilean invariance and has stronger topological properties owned to Helmholtz' and Kelvin's theorems. Numerous applications of CVE can be found for periodic domains in the literature starting with homogeneous isotropic turbulence~\cite{FaScKe99,FaPeSc2001,FS01,FSPWR03pf,Vasi,OYSFK07}, temporally developing mixing layers~\cite{SFPR05jfm} and homogeneous shear flow with and without rotation~\cite{Jacobitz}. For wall-bounded flows, the situation becomes more complex, because no-slip boundary conditions have to be taken into account. Indeed, no-slip boundary conditions generate vorticity due to the viscous flow interactions with the walls. For turbulent boundary layers, Khujadze {\it et al}.~\cite{Khu} obtained an efficient algorithm to extract coherent vorticity, constructing a locally refined grid using wavelets with mirror boundary conditions. However, this construction does not yield a multiresolution analysis where the basis functions are no more isotropic. Fr\"{o}hlich \& Uhlmann ~\cite{FU} constructed wavelets based on second kind Chebyshev polynomials and applied them to channel flow data. Scale-wise statistics in the wall-normal directions have thus been performed. However, no fast wavelet transform (FWT) is available for these Chebyshev wavelet bases. Two-dimensional (2D) wavelets have also been applied to wall-parallel planes in channel flows, in order to examine turbulent statistics, in particular statistics of energy transfer~\cite{DM1,JR}. The aim of the present work is to examine the role of coherent and incoherent flow contributions in 3D turbulent channel flow. We propose a novel construction of 3D isotropic orthogonal wavelets using boundary wavelets in the wall-normal direction and periodic wavelets in the wall-parallel directions. To this end, Cohen-Daubechies-Jawerth-Vial (CDJV) boundary wavelets~\cite{CDJV,CDV} having three vanishing moments, and the periodized Coiflet 30 wavelets~\cite{Daub} having ten vanishing moments are employed. These wavelets are orthogonal, the FWT can be used while taking into account boundary conditions, and the basis yields a multiresolution analysis. Hence, the basis functions are isotropic since they have only one scale in all three spatial directions. DNS computation of the channel flow has been performed, and the data are analyzed at different time instants, using the above boundary adapted 3D isotropic wavelets. The flow vorticity is decomposed into an orthogonal wavelet series, and we apply a thresholding to split the coefficients into two sets, the coherent and incoherent ones. The coherent vorticity, reconstructed from the few strongest wavelet coefficients, well preserves the turbulent statistics of the total flow, while the incoherent vorticity, reconstructed from the remaining large majority of the coefficients that are very weak, corresponds to a noise-like background flow. The corresponding coherent and incoherent velocity fields are reconstructed from the coherent and incoherent vorticity fields, respectively, using the Biot-Savart relation satisfying the no-slip conditions at the walls. Thus, we can efficiently examine the role of coherent vorticity in turbulent channel flow. Other conventional methods, such as the $Q$-criterion and the $\lambda_2$ method~\cite{Jeong0,Jeong} could be used to identify coherent vortices in physical space, as regions for which $Q$ or $\lambda_2$ is above a given threshold. Here, $Q$ is the second-invariant of the 3D velocity gradient tensor, and $\lambda_2$ is the second largest eigenvalue of $S_{ij}S_{jk}+A_{ij}A_{jk} $, where $S_{ij}$ and $A_{ij}$ are respectively the symmetric and antisymmetric tensor of the velocity gradient tensor. It should be noticed that these quantities do not preserve the scale information about the vortices, as the smoothness of the flow field is not preserved due to the clipping of vorticity in physical space. In contrast, the proposed wavelet filtering does preserve the smoothness of the coherent vorticity field and the multiscale properties of the coherent structures. The paper is organized as follows: Section \ref{sec2} presents the DNS computation and the data we analyze, including the methodology. The construction of isotropic wavelets is described, and the CVE method is summarized. Numerical results are shown in Sec. \ref{sec3}. Conclusions and perspectives are given in Sec. \ref{sec4}. \section{\label{sec2}DNS and methodology} \subsection{\label{sec2a}Direct Numerical Simulation} \begin{figure}[tb] \begin{center} \includegraphics[width=80mm]{./FIGURES/coordinate_gray-eps-converted-to.pdf} \caption{Flow configuration for the turbulent channel flow.} \label{coordinate} \end{center} \end{figure} We consider 3D incompressible fluid flow in a channel bounded by two parallel walls subjected to a streamwise mean pressure gradient, which is a canonical flow configuration. It is illustrated in Fig. \ref{coordinate} together with the Cartesian coordinate system $\bm{x}=(x_1,x_2,x_3)$, where the walls are at $x_2=\pm h$, $x_2$ being the wall-normal direction and $h$ the half width of the channel. The domain size in the streamwise $x_1$-direction is $2 \pi h$, and the size in the spanwise $x_3$-direction is $\pi h$. Periodic boundary conditions are respectively imposed in $x_1$- and $x_3$-directions, while in the $x_2$-direction no-slip boundary conditions are satisfied at the walls. The fluid flow motion obeys the Navier-Stokes equations with the incompressibility condition, \begin{eqnarray} & & \partial_t v_i + \partial_j (v_j v_i) = -\partial_i p + G \delta_{i 1} + \nu \partial_j \partial_j v_i, \label{NSeq} \\ & & \partial_j v_j = 0, \label{divu} \end{eqnarray} where $v_i$ $(i=1,2,3)$ is the $i$-th velocity component, $p$ is the pressure fluctuation, $G$ is the intensity of the mean pressure gradient in the $x_1$-direction, $\delta_{ij}$ is the Kronecker delta, $\nu$ is the kinematic viscosity, $t$ is time, and $\partial_t=\partial/\partial t$ and $\partial_i = \partial/\partial x_i$. Einstein's summation convention is used for repeated indices. We performed DNS of turbulent channel flow at $Re_\tau$ of $395$ using $N_1 N_2 N_3$ grid points, where $Re_\tau \equiv u_\tau h/\nu$, $u_\tau$ is the friction velocity defined by $[ \nu d U_1(x_2)/d x_2 ]^{1/2}$ at $x_2=-h$. The velocity field $v_j$ is decomposed as $v_j=U_j(x_2) + u_j$ with $U_j$ being the mean velocity defined as $U_j= \langle v_i \rangle$, and $u_j$ are the velocity fluctuations. Here $\langle \cdot \rangle$ denotes the $x_2$-dependent spatial average of $ {\cdot} $ over the $x_1$-$x_3$ plane, and $N_i$ is the number of the grid points in the $x_i$-direction, $N_1=N_3=256$ and $N_2=192$. The toroidal and poloidal representation of Eqs.~(\ref{NSeq}) and (\ref{divu}) is employed, in order to satisfy the incompressibility constraint as done by Kim {\it et al}.~\cite{KMM}. We used the Fourier pseudo-spectral method in the $x_1$-$x_3$ planes, and the Chebyshev-tau method in the $x_2$-direction. The Chebyshev collocation points are given by $x_2=h \cos \{\pi(2 j+1)/(2 N_2)\}$ $(j=0, 1, \cdots, N_2-1)$ (see, e.g., Appendix B in Ref.~\cite{Canuto}). The aliasing errors are removed by the $3/2$ rule in the $x_1$-$x_3$ planes, and by the $2/3$ rule in the $x_2$-direction. Time advancement is carried out using first-order implicit Euler method for the viscous terms, and a third-order Runge-Kutta method for the nonlinear terms and the mean pressure gradient term $G$ whose value is determined so that the total flow rate is kept constant. The DNS code has been developed in Ref.~\cite{MorishitaETC13}. Statistical quantities shown in this paper are obtained by time averaging over 40 DNS snapshots with intervals of $0.5$ washout time, defined by $ 2\pi h/U_1$. The averaging starts after the total Reynolds stress, $- \langle u_1 u_2 \rangle + \nu d U_1/d x_2$ has become quasi-stationary. \subsection{\label{sec2b}Wavelets} \begin{figure*}[tb] \begin{center} \includegraphics[height=150mm]{./FIGURES/algor-eps-converted-to.pdf} \caption{Flowchart of the CVE procedure. } \label{CVEalgor} \end{center} \end{figure*} In this subsection, we briefly summarize one-dimensional (1D) orthogonal periodized wavelets and 1D orthogonal boundary wavelets. Then, we propose a 3D orthogonal isotropic wavelet transform constructed by tensor product of these 1D wavelets. The CVE based on orthogonal wavelets to extract coherent vorticity out of turbulent channel flow is described in Sec. \ref{sec2c}. In Fig. \ref{CVEalgor}, we present the flowchart of the CVE method used here. \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/coif30_sca-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/coif30_wav-eps-converted-to.pdf} \caption{Coiflet 30 wavelet on the periodic domain: scaling function $\phi^P_{8,i}(x)$ (left) and corresponding wavelet $\psi^P_{8,i}(x)$ (right) both at scale $j=8$.} \label{coiflet30} \end{figure*} \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/cdjv_sca-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/cdjv_wav-eps-converted-to.pdf} \caption{CDJV wavelet on the interval: three scaling functions $\phi^B_{8,i}(x)$ (left) and three wavelets $\psi^B_{8,i}(x)$ (right) at scale $j=8$ and position $i=0$ (red), $63$ (green) and $127$ (blue) are shown. } \label{cdjv} \end{figure*} We first consider 1-periodic wavelets $\psi^P(x)$ and their corresponding scaling function $\phi^P(x)$, and orthogonal boundary adaptive wavelets $\psi^B(x)$ and their scaling function $\phi^B(x)$, with the boundaries at $x=(0,1)$. Wavelets at scale $j$ are obtained by dilation, so that $\psi^\gamma_{j,0}(x)=2^{-j/2}\psi^\gamma(2^{-j}x)$ and $\phi^\gamma_{j,0}(x)=2^{-j/2}\phi^\gamma(2^{-j}x)$, where $\gamma=P, B$. The periodized orthogonal wavelets are also self-similar with respect to translation. Then the scaling function $\phi^\gamma$ and wavelet function $\psi^\gamma$ at scale $2^{-j}$ $(j\ge 0)$ and position $2^{-j}i$ $(i=0,1,\cdots,2^{j}-1)$, $\phi^P_{j,i}(x)$ and $\psi^P_{j,i}(x)$, are defined as $\phi^P_{j,i}(x)=2^{-j/2}\phi^P(2^{-j}x-2^{-j}i)$ and $\psi^P_{j,i}(x)=2^{-j/2}\psi^P(2^{-j}x - 2^{-j}i)$. In contrast, $\psi^B_{j,0}$ and $\phi^B_{j,0}$ are no more translation invariant due to the boundary conditions, which modify the wavelets as position $i$ changes. Readers interested in the details of boundary adapted wavelets may refer to the textbook~\cite{Mallat}, as the construction of $\psi^B_{j,i}$ and $\phi^B_{j,i}$ is rather technical. All wavelets used here are orthonormal, i.e., $ \int_0^1 \psi_{j,i}^\gamma (x) \psi_{j',i'}^\gamma (x) d x =\delta_{ii'}\delta_{jj'} $, $ \int_0^1 \phi_{j,i}^\gamma (x) \phi_{j,i'}^\gamma (x) d x =\delta_{ii'}$, and $ \int_0^1 \psi_{j,i}^\gamma (x) \phi_{j,i'}^\gamma (x) d x =0 $. In this paper, we use Coiflet 30 wavelets~\cite{Daub} in the $x_1$- and $x_3$-directions, and the CDJV wavelets having 3 vanishing moments~\cite{CDJV,CDV} in the $x_2$-direction, both wavelets being compactly supported. The Coiflet 30 wavelets are quasi-symmetric and have 10 vanishing moments. The largest scale $2^{-j_0}$ of the CDJV wavelets satisfies $2^{j_0-1} \ge 3$~\cite{CDV}. The illustrations of these wavelet functions are shown in Figs. \ref{coiflet30} and \ref{cdjv}. The 3D orthogonal wavelets $\Psi^\mu$ $(\mu=1,2,\cdots,7)$ are obtained by tensor product such that \begin{equation} \begin{split} & \Psi^\mu_{j,{\bm i}} (x_1,x_2,x_3) =\\ & \left\{ \begin{array}{lll} \psi^P_{j,i_1}(x_1) \phi^B_{j',i_2}(x_2) \phi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=1, \\ \phi^P_{j,i_1}(x_1) \psi^B_{j',i_2}(x_2) \phi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=2, \\ \phi^P_{j,i_1}(x_1) \phi^B_{j',i_2}(x_2) \psi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=3, \\ \psi^P_{j,i_1}(x_1) \phi^B_{j',i_2}(x_2) \psi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=4, \\ \psi^P_{j,i_1}(x_1) \psi^B_{j',i_2}(x_2) \phi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=5, \\ \phi^P_{j,i_1}(x_1) \psi^B_{j',i_2}(x_2) \psi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=6, \\ \psi^P_{j,i_1}(x_1) \psi^B_{j',i_2}(x_2) \psi^P_{j,i_3}(x_3) & {\mathrm{for}} & \mu=7, \end{array} \right. \end{split} \end{equation} where ${\bm i}=(i_1,i_2,i_3)$, $j'=j_0+j$ and $j=0,\cdots,J-1$. The corresponding scaling function is defined as $\Phi(x_1,x_2,x_3)=\phi^P(x_1) \phi^B(x_2) \phi^P(x_3)$. Now, let us consider a 3D vector field ${\bm w}({\bm x})=(w_1, w_2, w_3)$ in the computational domain $D$, where $D=\{x_1,x_2,x_3|0\le x_1 \le 2\pi h, -h \le x_2 \le h, 0\le x_1 \le \pi h \}$. Before applying this wavelet decomposition, we interpolate ${\bm w}({\bm x})$ on an equidistant grid in the $x_2$-direction from the DNS data non-uniformly sampled on $N_2$ Chebyshev grid points in the wall-normal direction. We thus get ${\bm w}({\bm x})$ uniformly sampled on $N_2'$ equidistant grid points at $x_{2,n}= h\{-1 + 2 n /(N_2'-1)\}$ $(n=0,\cdots,N_2'-1)$ using the Chebyshev interpolation~\cite{Clenshaw}. We choose $N_2'$ to be equal to $2048$ so that the flow field near the walls is kept well-resolved. We have $2/N_2' \sim 8\pi^2/N_2^2$, which shows that the grid width after the interpolation is comparable to the minimum grid width of the Chebyshev grid. In the $x_1$- and $x_3$-directions, we keep ${\bm w}({\bm x})$ uniformly sampled on $N_1$ and $N_3(=N_1)$ equidistant grid points, respectively. The field ${\bm w}({\bm x})$, now sampled on $N_1 \times N_2'\times N_3 $ equidistant grid points, can then be decomposed into an isotropic orthogonal wavelet series as follows; \begin{equation} {\bm w}({\bm x}) = {\bar{\bm w}} + \sum_{\mu=1}^7 \sum_{j=0}^{J-1} \sum_{i_1,i_2,i_3=0}^{2^{j}-1} {\widetilde {\bm w}}^\mu_{j,{\bm i}} \Psi^\mu_{j,{\bm i}}({\bm x}) , \end{equation} with wavelet coefficients computed with wavelets $\Psi^\mu_{j,{\bm i}}$ \begin{equation} \begin{split} & {\widetilde {\bm w}}^\mu_{j,{\bm i}} = \\ & \frac{1}{{\mathcal{V}}} \int_0^{L_1} d x_1 \int_{-h}^h d x_2 \int_0^{L_3} d x_3 \, {\bm w}({\bm x}) \Psi^\mu_{j,{\bm i}} \left( \frac{x_1}{2 \pi h}, \frac{x_2+h}{2h},\frac{x_3}{\pi h} \right), \end{split} \end{equation} \and the mean value computed with scaling function $ \Phi$ \begin{equation} \begin{split} & \bar{{\bm w}} = \\ & \frac{1}{{\mathcal{V}}} \int_0^{2\pi h} d x_1 \int_{-h}^h d x_2 \int_0^{\pi h} d x_3 \, {\bm w}({\bm x}) \Phi \left( \frac{x_1}{2 \pi h}, \frac{x_2+h}{2h} ,\frac{x_3}{\pi h} \right), \end{split} \end{equation} where $J=\log_2 N_1$ and ${\mathcal{V}}=4 \pi^2 h^3$. \subsection{\label{sec2c}Coherent vorticity extraction} \begin{figure}[tb] \begin{center} \includegraphics[width=80mm]{./FIGURES/cr_ens-eps-converted-to.pdf} \caption{Compression curve for the CVE: \% of retained enstrophy per unit volume vs. \% of retained wavelet coefficients. The circle corresponds to the threshold $T$ used for CVE in the following. } \label{cr_vs_ens} \end{center} \end{figure} We extract coherent vorticity out of turbulent channel flow data using the CVE method based on the wavelet decomposition of vorticity ${\bm \omega}(=\nabla \times {\bm v})$. In the following, we summarize our method. Then, since coherent structures do not have a universal definition yet, we define coherent structures as what remains after denoising. As first guess we consider the simplest type of noise, namely additive, Gaussian and white, i.e., uncorrelated. Readers interested in details of this ansatz may refer to the original articles, e.g., Refs. ~\cite{FaPeSc2001,FaScKe99,OYSFK07}. The CVE method is based on nonlinear thresholding the orthogonal wavelet coefficients of vorticity. To this end the vorticity ${\bm \omega}$, interpolated on a sufficiently fine equidistant grid, is decomposed into an orthogonal wavelet series using the FWT. Applying thresholding to the wavelet coefficients, we split the flow into coherent and incoherent contributions. The corresponding coherent and incoherent vorticity fields are then obtained by inverse wavelet transform. In previous work, we used Donoho's threshold \cite{Donoho} to determine the value of the threshold and estimate the variance of the incoherent vorticity using an iterative scheme. Azzalini {\it et al}. \cite{Azz} investigated the convergence of the iterative scheme and for isotropic turbulence Okamoto {\it et al.} \cite{OYSFK07} found that, depending on the Reynolds number, 8.7 \% and 6.0 \% are obtained for $Re_\lambda=167$ and $Re_\lambda=732$, respectively. In Ref. \cite{FaPeSc2001}, Farge {\it et al.} used one iteration only, which was sufficient to get good compression while preserving the statistics of the total flow. For the turbulent channel flow studied here we tried Donoho's threshold and found that very few wavelet coefficients keep almost the whole enstrophy of the flow, which is illustrated in the compression curve, shown in Fig. \ref{cr_vs_ens}. The flow visualization in Fig. \ref{vis_abs} shows tube-like coherent vortex structures of different intensity, which are very strong close to the wall and much weaker in the center of the channel. In the current work, we propose instead an {\it ad hoc} criterion for the threshold defined by $T=\langle |{\widetilde {\bm \omega}}^\mu_{j,{\bm i}}| \rangle_w +\alpha \langle \left(|{\widetilde {\bm \omega}}^\mu_{j,{\bm i}}| - \langle |{\widetilde {\bm \omega}}^\mu_{j,{\bm i}}| \rangle \right)^2 \rangle_w^{1/2}$, where $\langle |{\widetilde {\bm \omega}}^\mu_{j,{\bm i}}| \rangle_w = \sum_{\mu=1}^7 \sum_{j=0}^{J-1} \sum_{i_1,i_2,i_3=0}^{2^j-1} |{\widetilde {\bm \omega}}^\mu_{j,{\bm i}}|/(N_1N_2'N_3) $. Our aim is to retain only those wavelet coefficients which are responsible for the nonlinear dynamics of the flow, even if the fully developed turbulent regime has not been yet reached. We set $\alpha =0.75$ in the threshold value $T$ such that both velocity and vorticity statistics (as a function of $x_2$), together with the nonlinear dynamics and structures, are well preserved by the coherent flow. Using the inverse FWT, the coherent vorticity ${\bm \omega}_{c}$ is reconstructed from the wavelet coefficients whose intensity is larger than the threshold value $T$. The incoherent vorticity ${\bm \omega}_{i}$ is then obtained using ${\bm \omega}_{i} = {\bm \omega} - {\bm \omega}_{c}$. To get ${\bm \omega}_{c}$ and ${\bm \omega}_{i}$ sampled on the Chebyshev grid points, which are useful and efficient for data analysis presented in Sec. \ref{sec3}, we perform a cubic spline interpolation in the $x_2$-direction. Owing to the orthogonality of the wavelet decomposition, ${\bm \omega}_c $ is orthogonal to ${\bm \omega}_i $ and thus $Z_t=Z_c+Z_i$, where $Z_t$, $Z_c$ and $Z_i$ are respectively the total, coherent and incoherent enstrophy per unit volume, defined as $Z_\alpha =\int \!\! \int \!\! \int_D |{\bm \omega}_\alpha|^2 d {\bm x} /(2 {\mathcal{V}} )$ ($\alpha=t, c, i$). The coherent velocity ${\bm v}_c$ and the incoherent velocity ${\bm v}_i$ are computed from ${\bm \omega}_c$ and ${\bm \omega}_i$ by solving Biot-Savart's relation, $\nabla^2 {\bm v}=-\nabla \times {\bm \omega}$, respectively. It is noted that ${\bm v}_i$ and ${\bm v}_c $ are weakly non orthogonal, i.e., the cross term $ \int {\bm v}_i \, {\bm v}_c \, d{\bm x}$ is below $0.4 \%$ of the total energy. \section{\label{sec3}Numerical results} Now we analyze 40 snapshots of DNS data for the turbulent channel flow with intervals of 0.5 washout times, and we ensemble-average over those 40 snapshots to guarantee well-converged statistical results. We examine contributions of coherent and incoherent flows obtained with the previously described CVE method. Quantities with the superscript $^+$ are expressed in wall units, i.e., they are non-dimensionalized by $u_\tau$ and $\nu$. We define the distance from the wall $y$ as $y=x_2+1$. \subsection{Visualization} \begin{figure*}[tb] \begin{tabular}{c|c} \begin{minipage}{0.5\textwidth} \begin{center} \includegraphics[width=70mm]{./FIGURES/wa_tot1-eps-converted-to.pdf}\\ \includegraphics[width=70mm]{./FIGURES/wa_coh1-eps-converted-to.pdf}\\ \includegraphics[width=70mm]{./FIGURES/wa_inc1-eps-converted-to.pdf} \end{center} \end{minipage} & \begin{minipage}{0.5\textwidth} \begin{center} \includegraphics[width=60mm]{./FIGURES/wa_tot2-eps-converted-to.pdf}\\ \includegraphics[width=60mm]{./FIGURES/wa_coh2-eps-converted-to.pdf}\\ \includegraphics[width=60mm]{./FIGURES/wa_inc2-eps-converted-to.pdf} \end{center} \end{minipage} \end{tabular} \caption{Visualization of total vorticity $\bm{\omega}$ (green), coherent vorticity $\bm{\omega}_c$ (red) and incoherent vorticity $\bm{\omega}_i$ (blue). The left column presents isosurfaces $|\bm{\omega}^{+}|=|\bm{\omega}_c^{+}|=0.3$ and $|\bm{\omega}_i^{+}|=0.1$. The right column shows their zooms in the wall region where $0 \le x_1 \le 0.71\pi h$, $ -0.12 h \le x_2 \le h$, and $0.5\pi h \le x_3 \le \pi h$.} \label{vis_abs} \end{figure*} \begin{figure*}[tb] \begin{tabular}{c|c} \begin{minipage}{0.5\textwidth} \begin{center} \includegraphics[width=70mm]{./FIGURES/ww_tot1-eps-converted-to.pdf}\\ \includegraphics[width=70mm]{./FIGURES/ww_coh1-eps-converted-to.pdf}\\ \includegraphics[width=70mm]{./FIGURES/ww_inc1-eps-converted-to.pdf} \end{center} \end{minipage} & \begin{minipage}{0.5\textwidth} \begin{center} \includegraphics[width=45mm]{./FIGURES/ww_tot2-eps-converted-to.pdf}\\ \includegraphics[width=45mm]{./FIGURES/ww_coh2-eps-converted-to.pdf}\\ \includegraphics[width=45mm]{./FIGURES/ww_inc2-eps-converted-to.pdf} \end{center} \end{minipage} \end{tabular} \caption{Visualization of total vorticity $\bm{\omega}$ (green), coherent vorticity $\bm{\omega}_c$ (red) and incoherent vorticity $\bm{\omega}_i$ (blue). Isosurfaces $|\bm{\omega}^{+}|=|\bm{\omega}_c^{+}|=\langle|\bm{\omega}^{+}|\rangle+3\langle(|\bm{\omega}^{+}|-\langle|\bm{\omega}^{+}|\rangle)^2\rangle^{1/2}$ and $|\bm{\omega}_i^{+}|=\langle|\bm{\omega}_i^{+}|\rangle+3\langle(|\bm{\omega}_i^{+}|-\langle|\bm{\omega}_i^{+}|\rangle)^2\rangle^{1/2}$ are shown. The right column presents corresponding zooms in the core where $ 0 \le x_1 \le 0.79\pi h$, $-0.78 h \le x_2 \le 0.78 h$ and $ 0 \le x_3 \le 0.5\pi h$.} \label{vis_sd} \end{figure*} Visualizations of isosurface values of the modulus of vorticity for the total, coherent and incoherent flows given at the same time instant are shown in Figs. \ref{vis_abs} and \ref{vis_sd}. Corresponding zooms are also presented to see flow structures more clearly. Figure \ref{vis_abs} shows that the most intense vorticity structures are near the walls. Since the incoherent vorticity is much weaker than the total and coherent vorticities in Fig. \ref{vis_abs}, the isosurface value for the incoherent vorticity ${\bm \omega}_i$ is reduced by a factor 3 compared to the coherent and total vorticities. On the other hand, Fig. \ref{vis_sd} visualizes vorticity structures in the core region, using $y^+$-dependent isosurface values, $|\bm{\omega}^{+}|=|\bm{\omega}_c^{+}|=\langle|\bm{\omega}^{+}|\rangle+3\langle(|\bm{\omega}^{+}|-\langle|\bm{\omega}^{+}|\rangle)^2\rangle^{1/2}$ and $|\bm{\omega}_i^{+}|=\langle|\bm{\omega}_i^{+}|\rangle+3\langle(|\bm{\omega}_i^{+}|-\langle|\bm{\omega}_i^{+}|\rangle)^2\rangle^{1/2}$, recalling $\langle \cdot \rangle$ denotes the $y^+$-dependent spatial average of $ {\cdot} $ over each wall-parallel plane. We observe that the total flow exhibits intense vortex tubes near the walls, as in previous DNS (e.g., Ref. ~\cite{Blackburn}), but we also see them in the core region, however they are less intense. Looking at the coherent flow, we find that these tubes are well preserved by ${\bm \omega}_c$, which is reconstructed from only $ 0.55\% $ of the $256^2 \times 2048(\simeq 13 \times 10^7)$ wavelet coefficients, i.e., $5.9\%$ of the original $256^2 \times 192(\simeq 1.2 \times 10^7)$ grid points. The coherent flow retains almost all of the total energy and enstrophy, $99.9 \%$ of the total energy and $99.7 \%$ of the total enstrophy. In contrast, the incoherent vorticity ${\bm \omega}_i$ looks less organized without exhibiting vortex tubes near the walls and in the core region. Although the incoherent flow is represented by the remaining majority of wavelet coefficients, it retains a negligible amount of energy, namely $2.3\times10^{-3} \%$ of the total energy, and only $0.5 \%$ of total enstrophy. \subsection{Mean velocity and vorticity statistics} \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/meanvel_log-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/meanvor_log-eps-converted-to.pdf} \caption{Streamwise mean velocity $U_1^+(y^+)$ (left) and spanwise mean vorticity $\Omega_3^+(y^+)$ for total, coherent and incoherent flows, together with the DNS results of Ref. ~\cite{MKM} in the lin-log representation.} \label{meanvor} \end{figure*} We analyze the statistics of the mean velocity and vorticity profiles of the coherent and incoherent flows, and compare them with the total flow. The results are averaged over 40 snapshots. Figure \ref{meanvor} shows the $y^+$-dependence of the streamwise mean velocity $U_1^+(y^+)$ and of the spanwise mean vorticity $\Omega_3^+(y^+)$, averaged over $x_1$-$x_3$ planes, for the total, coherent and incoherent flows. It is observed that the coherent flow perfectly preserves $U_1^+(y^+)$ and $\Omega_3^+(y^+)$, while both incoherent contributions are very weak. It can be noted that $U_2^+(y^+)$ vanishes identically and that $U_3^+(y^+)$ almost vanishes for the total, coherent and incoherent flows. This implies that $\Omega_1^+(y^+)$ is almost zero, and $\Omega_2^+(y^+)$ is identically zero. The comparison of $U_1^+$ with the DNS data at $Re_\tau=395$ in Moser {\it et al}.~\cite{MKM} confirms the validity of the present DNS. \subsection{Statistics of velocity and vorticity fluctuations} \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/RMS_vel-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/RMS_vor-eps-converted-to.pdf} \caption{RMS of $u_j^+$ (left) and RMS of $\omega_j^+$ (right) for total, coherent and incoherent flows. } \label{RMS_middle} \end{figure*} The root-mean-square (RMS) of the velocity fluctuations $u_j^+$ $(j=1,2,3)$ as a function of $y^+$ are shown in Fig. \ref{RMS_middle} (left). Again, we find an excellent agreement between the total and the coherent flow for all values of $y^+$, while the incoherent contribution is negligibly small. For RMS of the vorticity fluctuations $\omega_j^+$ in Fig. \ref{RMS_middle} (right), the coherent contributions well preserves the total RMS of $\omega_j^+$. The vorticity RMS of the incoherent flow is much smaller than that of the total flow. \subsection{Probability density functions of velocity and vorticity} \begin{figure*}[tb] \begin{tabular}{c|c} \begin{minipage}{0.5\textwidth} \centering \includegraphics[width=80mm]{./FIGURES/PDF_u1_w-eps-converted-to.pdf}\\ \includegraphics[width=80mm]{./FIGURES/PDF_u1_l-eps-converted-to.pdf}\\ \includegraphics[width=80mm]{./FIGURES/PDF_u1_c-eps-converted-to.pdf} \end{minipage} & \begin{minipage}{0.5\textwidth} \centering \includegraphics[width=80mm]{./FIGURES/PDF_w3_w-eps-converted-to.pdf}\\ \includegraphics[width=80mm]{./FIGURES/PDF_w3_l-eps-converted-to.pdf}\\ \includegraphics[width=80mm]{./FIGURES/PDF_w3_c-eps-converted-to.pdf} \end{minipage} \end{tabular} \caption{PDFs of $u_1^+$; (top left) $y^{+}=4.6$ viscous sublayer, (top right) $y^{+}=96.8$ around the log region, and (bottom) $y^{+}=378.8$ around the center of the channel. } \label{PDF} \end{figure*} Figure \ref{PDF} (left) shows the probability density functions (PDFs), estimated using histograms with $200$ bins, of the streamwise velocity fluctuations $u_1^+$ for the total, coherent and incoherent flows at three representative positions $y^+$: in the viscous sublayer, the log region, and near the center of the channel. In all cases, we observe that the PDFs for the total and coherent velocity fluctuations perfectly superimpose, which indicates that high order statistics are well preserved by the coherent flow. We also find that the velocity PDFs remain skewed in the different regions and agree well with the data of Ref.~\cite{MKM}, using appropriate renormalization. In contrast, the PDFs of the incoherent velocity fluctuations are symmetric, and have strongly reduced variances. For the incoherent velocity, we analyzed $y^+$-dependent flatness, and found values around 4 in the viscous sublayer and in the log region, which decrease to 3.6 near the center of the channel. For the $y^+$-dependent skewness, fluctuations around zero are observed with an amplitude below 0.05. The PDFs of the incoherent velocity well superimpose the logistic distribution with zero mean and the variances $\sigma^2(y^+)$ of the incoherent velocity, though their flatness is 1.2, which is much smaller than the PDFs of the incoherent velocity. The distributions $P(u_1^+)$ are given by $\exp (-\pi u_1^+ /s)/\{ s(1+\exp (-u_1^+ /s) \}$, where $s=3^{1/2} \sigma(y^+)/\pi$. In Fig. \ref{PDF} (right), we illustrate the PDFs of the vorticity fluctuations $\omega_3^{+}$ at three representative positions $y^+$: in the viscous sublayer, the log region, and near the center of the channel. The coherent vorticity fluctuations well represent the total vorticity PDFs which are skewed in all cases, while the corresponding incoherent PDFs are symmetric. The variances of these incoherent PDFs are respectively significantly weaker than the variances of the total and coherent PDFs. \subsection{Energy spectra} To get insight into the scale distribution of turbulent kinetic energy we analyze the 1D energy spectra of the streamwise velocity $u_1^+$ in the streamwise direction $E^+(k_1 h,y^+)$, which is defined as $E^+(k_1 h,y^+) = \sum' |{\hat u}_i(k_1 h,y^{+},k_3 h)|^2 /2$, where ${\hat u}_i(k_1 h,y,k_3 h)$ is the Fourier transform of ${\bm u}^+({\bm x})$ in the $x_1$-$x_3$ planes, $\sum'$ denotes the summation in terms of all $k_3$. The results are shown in Fig.~\ref{spectra}, again for the total, coherent an incoherent flows at three representative positions; in the viscous sublayer, the log layer and near the center of the channel. The dimensionless wavenumber in the $x_1$-direction is denoted by $k_1 h$. Figure \ref{spectra} shows that the spectral distribution of turbulent kinetic energy is well preserved by the coherent flow, from the viscous sublayer to the center of the channel. In contrast, the incoherent energy exhibits an almost flat spectrum, which corresponds to equipartition of incoherent energy, i.e., decorrelation of the incoherent flow in physical space. At large wavenumbers close to the cut-off scale, imposed by the resolution of the DNS, we find that the incoherent energy dominates the total energy in the viscous sublayer and the log layer, while it dominates the coherent energy around the center of the channel. However, this is not surprising since the wavelet decomposition is orthogonal for vorticity but not for velocity, due to the fact that the Biot-Savart operator is not diagonal in wavelet space. Note that ${\hat {\bm \omega}}_c^+ (k_1,y^+,k_3)$ and ${\hat {\bm \omega}}_i^+ (k_1,y^+,k_3)$ are not orthogonal for any fixed $(k_1,k_3)$ at every $y^+$. But even though the cross-term $\langle {\bm \omega}_c^{+} {\bm \cdot} {\bm \omega}_i^{+} \rangle(x_2) \neq 0$, its averaged contribution vanishes, $\int_{-h}^h d x_2 \langle {\bm \omega}_c^+ {\bm \cdot} {\bm \omega}_i^+ \rangle(x_2) =0 $. \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/Spec_kx_w_u1-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/Spec_kx_l_u1-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/Spec_kx_c_u1-eps-converted-to.pdf} \caption{Dimensionless energy spectra of $u_1^+$ in the $x_1$-direction at three representative $y^+$: (top left) $y^+=4.6$ viscous sublayer, (top right) $y^+=96.8$ around the log region, and (bottom) $y^+=378.8$ around the center of the channel. } \label{spectra} \end{figure*} \begin{figure}[tb] \begin{center} \includegraphics[scale=0.8]{./FIGURES/cr_scale-eps-converted-to.pdf} \caption{Scale-dependent compression rate.} \label{cr_scale} \end{center} \end{figure} The compression is most efficient for small scales, i.e., small $j$ and large wavenumbers (Fig. \ref{cr_scale}). This implies that the scale-by-scale incoherent enstrophy is comparable or larger than the scale-by-scale coherent enstrophy. \subsection{Nonlinear dynamics} \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/E_PTPi-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/E_eV-eps-converted-to.pdf} \caption{Production term $P$, turbulent diffusion term $T$ and pressure diffusion term $\Pi$ vs. $y^+$ (left). Energy dissipation $\epsilon$ and viscous diffusion $V$ vs. $y^+$ (right). Coherent and incoherent contributions are presented together with the total one.} \label{Energy_budget_m} \end{figure*} \begin{figure*}[tb] \includegraphics[width=80mm]{./FIGURES/E_Pe-eps-converted-to.pdf} \includegraphics[width=80mm]{./FIGURES/Reystress-eps-converted-to.pdf} \caption{The ratios of total, coherent and incoherent productions $P$ to the dissipation of turbulent kinetic energy for total flow, $\epsilon$, vs. $y^+$(left). The Reynolds stress $-\langle u_1^+ u_2^+ \rangle$ vs. $y^{+}$(right).} \label{Povere} \end{figure*} To get further insight into the nonlinear dynamics, we consider the energy budget given in the equation for $\langle u_j^+ u_j^+ \rangle/2$ per unit mass~\cite{MasourKimMoin}: \begin{equation} \begin{split} & \frac{1}{2} \left( \partial_t + U_j^+ \partial_j \right) \left< u_j^{+} u_j^{+} \right> = \\ & P({\bm v}^+) + T({\bm u}^+) + \Pi({\bm u}^+,p^+) - \epsilon({\bm u}^+) + V({\bm u}^+), \label{En_budget} \end{split} \end{equation} where $P({\bm v}^+) = - \left< u_j^{+} u_l^{+} \right> \partial_l U_j^{+}$, $T({\bm u}^+) = - \partial_l \left< u_j^{+} u_j^{+} u_l^{+} \right>/2$, $\Pi({\bm u}^+,p^+) = - \partial_l \left< p^{+} u_l^{+} \right>$, $\epsilon({\bm u}^+) = \nu \left< \partial_l u_j^{+} \partial_l u_j^{+} \right>$, $V({\bm u}^+) = \nu \partial_l \partial_l \left< u_j^{+} u_j^{+} \right>$. In Fig. \ref{Energy_budget_m} (left), we see that three nonlinear coherent contributions, corresponding to production $P({\bm v}_c^+)$, turbulent diffusion $T({\bm u}_c^+)$ and pressure diffusion $\Pi({\bm u}_c^+,p^+_c)$, are in good agreement with the corresponding total ones. Hence, the coherent flow almost perfectly preserves the nonlinear dynamics. Thus, we anticipate that the departure of the coherent flow from the total flow is negligibly small. Indeed, the incoherent contribution to the different terms, defined by $P({\bm v}^+)-P({\bm v}_c^+)$, $T({\bm u}^+)-T({\bm u}_c^+)$ and $\Pi({\bm u}^+,p^+)- \Pi({\bm u}_c^+,p^+_c)$, is almost zero. The two viscous contributions, $\epsilon({\bm u}^+)$ and $V({\bm u}^+)$, are also well retained by the coherent flow, $\epsilon({\bm u}_c^+)$ and $V({\bm u}_c^+)$, as confirmed in Fig. \ref{Energy_budget_m} (right). In the viscous sublayer, the incoherent flow has small contribution on $\epsilon({\bm u}^+)$ and $V({\bm u}^+)$. The incoherent contribution to the viscous terms, respectively measured by $\epsilon({\bm u}^+)-\epsilon({\bm u}_c^+)$ and $V({\bm u}^+)-V({\bm u}_c^+)$, becomes even smaller and more negligible as $y^+$ increases, a behavior which is expected. The ratio between the production and the dissipation yields insight into the equilibrium of the turbulent flow in the log region, as discussed in Ref.~\cite{MKM}. Figure \ref{Povere} (left) shows this balance. Considering $P({\bm v}_c^+) / \epsilon({\bm u}^+)$, the coherent contribution perfectly superimposes with the ratio of the total flow, $P({\bm v}^+) / \epsilon({\bm u}^+)$. The corresponding quantity for the incoherent flow, $\{P({\bm v}^+)-P({\bm v}_c^+) \}/ \epsilon({\bm u}^+)$, is negligible, as expected from Fig. \ref{Energy_budget_m} (left). The Reynolds stress defined by $-\langle u_1^+ u_2^+ \rangle$ measures the fluctuation of turbulent momentum. The analysis of the Reynolds stress provides detailed information on the contribution to the turbulence production from various events occurring in the flows. Figure \ref{Povere} (right) shows that the coherent Reynolds stress well represents the Reynolds stress for the total flow, while its incoherent contribution is negligibly small. The interaction between the coherent flows is predominant over the stress. In contrast, the remaining interactions play a non-significant role in the stress, not only between the incoherent flows but also between the coherent and the incoherent flows. \section{\label{sec4}Conclusions and perspectives} DNS data of turbulent channel flow at moderate Reynolds number have been analyzed using the coherent vorticity extraction method. Boundary adapted isotropic wavelets have been developed and implemented into a fast wavelet transform. By thresholding the wavelet coefficients, the flow has been decomposed into coherent and incoherent contributions. We found that few percent of wavelet coefficients, i.e., 6 \%, are sufficient to represent the coherent structures of the flow. The low order statistics, mean velocity, mean vorticity, RMS velocity and RMS vorticity of the coherent part agree very well with those of the total flow. A spectral decomposition of turbulent kinetic energy confirms that the coherent flow matches the spectrum all along the inertial range. In contrast, the incoherent flow exhibits energy equipartition which suggests that filtering it out corresponds to modeling turbulent dissipation. In order to obtain reliable statistical results, averaging over 40 flow snapshots has been performed. To get insight into the flow dynamics we analyzed the energy budget and we found that the coherent flow almost perfectly retains the nonlinear dynamics. The production/dissipation ratio of the coherent flow superimposes well the one of the total flow in the log layer, while the interactions between incoherent-incoherent and coherent-incoherent contributions are negligibly small. Although the coherent and incoherent vorticity fields are not perfectly divergence free. The divergence issue is not crucial as discussed in appendix \ref{app_div}. The present construction requires that the DNS data be interpolated onto an equidistant grid. This limits the applicability of the current CVE algorithm as higher resolution DNS data may not be handled due to the implied memory requirements. One way to overcome this is the use of Chebyshev wavelets, see e.g.~\cite{FU,Plonka}. In the appendix \ref{Chewave}, we tested this approach and we have shown that similar results in terms of statistics and compression rate are indeed obtained. The CVE results are encouraging for developing coherent vorticity simulation (CVS) of wall bounded turbulent flows. We anticipate that for higher Reynolds number the compression rate will further improve, similar to what was found for isotropic turbulence~\cite{OYSFK07}. CVS is based on a deterministic computation of the coherent flow evolution using an adaptive orthogonal wavelet basis ~\cite{FS01}. The influence of the incoherent background flow is neglected to model turbulent dissipation. Applications of CVS to turbulent mixing layers and isotropic turbulence can be found in Refs.~\cite{SFPR05jfm} and ~\cite{SIAMMMS}, respectively. Some challenges for future work are that the wavelet bases are not orthogonal in 2D planes for fixed $ y^+$. This implies that 2D statistics cannot be done, especially at small scales. In this case it would be better to apply 2D wavelets in each plane, as done in previous work \cite{DM1}. \begin{acknowledgments} We thank Professor Javier Jim\'enez for fruitful discussions about computing the mean pressure and the divergence issue. The computations were carried out on the FX100 systems at the Information Technology Center of Nagoya University. This work was partly supported by JSPS KAKENHI Grant Numbers (S) 24224003 and (C) 25390149. K.S. thankfully acknowledges financial support from ANR, contract SiCoMHD (ANR-Blanc 2011-045). M.F. and K.S. acknowledge support by the French Research Federation for Fusion Studies within the framework of the European Fusion Development Agreement (EFDA). \end{acknowledgments}
45a72c3603ddbbf061f91e39da13df6ea83a0c0c
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} We study $r$-nets, a powerful tool in computational and metric geometry, with several applications in approximation algorithms. An $r$-net for a metric space $(X,\norm{\cdot}), \, |X|=n$ and for numerical parameter $r$ is a subset $R\subseteq X$ such that the closed $r/2$-balls centered at the points of $R$ are disjoint, and the closed $r$-balls around the same points cover all of $X$. We define approximate $r$-nets analogously. Formally, \begin{dfn} Given a pointset $X \subseteq {\mathbb R}^d$, a distance parameter $r \in {\mathbb R}$ and an approximation parameter $\epsilon >0$, a $(1+\epsilon)r$-net of $X$ is a subset $R \subseteq X$ s.t. the following properties hold: \begin{enumerate} \item (packing) For every $p, q \in R$, $p \neq q $, we have that $\norm{p-q}_2 \geq r$. \item (covering) For every $p \in X$, there exists a $q \in R$ s.t. $\norm{p-q}_2 \leq (1+\epsilon)r$. \end{enumerate} \end{dfn} \paragraph{Previous Work.} Finding $r$-nets can be addressed naively by considering all points of $X$ unmarked and, while there remains an unmarked point $p$, the algorithm adds it to $R$ and marks all other points within distance $r$ from $p$. The performance of this algorithm can be improved by using grids and hashing \cite{HP04}. However, their complexity remains too large when dealing with big data in high dimension. The naive algorithm is quadratic in $n$ and the grid approach is in $O(d^{d/2} n)$, hence it is relevant only for constant dimension $d$~\cite{HR15}. In \cite{HPM05}, they show that an approximate net hierarchy for an arbitrary finite metric can be computed in $O(2^{ddim}n \log n)$, where $ddim$ is the doubling dimension. This is satisfactory when doubling dimension is constant, but requires a vast amount of resources when it is high. When the dimension is high, there is need for algorithms with time complexity polynomial in $d$ and subquadratic in $n$. One approach, which computes $(1+\epsilon)r$-nets in high dimension is that of \cite{EHS15}, which uses the Locality Sensitive Hashing (LSH) method of \cite{AI08}. The resulting time complexity is $\tilde{O}(dn^{2-\Theta(\epsilon)})$, where $\epsilon>0$ is quite small and $\tilde{O}$ hides polylogarithmic factors. In general, high dimensional analogues of classical geometric problems have been mainly addressed by LSH. For instance, the approximate closest pair problem can be trivially solved by performing $n$ approximate nearest neighbor (ANN) queries. For sufficiently small $\epsilon$, this costs $\tilde{O}(dn^{2-\Theta(\epsilon)})$ time, due to the complexity factor of an LSH query. Several other problems have been reduced to ANN queries \cite{GIV01}. Recently, Valiant \cite{Val12}, \cite{Val15} presented an algorithm for the approximate closest pair problem in time $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$. This is a different approach in the sense that while LSH exploits dimension reduction through random projections, the algorithm of \cite{Val15} is inspired by high dimensional phenomena. One main step of the algorithm is that of projecting the pointset up to a higher dimension. \paragraph{Our Contribution.} We present a new randomized algorithm that computes approximate $r$-nets in time subquadratic in $n$ and polynomial in the dimension, and improves upon the complexity of the best known algorithm. Our method does not employ LSH and, with probability $1-o(1)$, it returns $R\subset X$, which is a $(1+\epsilon)r$-net of $X$. We reduce the problem of an approximate $r$-net for arbitrary vectors (points) under Euclidean distance to the same problem for vectors on the unit sphere. Then, depending on the magnitude of distance $r$, an algorithm handling ``small" distances or an algorithm handling ``large" distances is called. These algorithms reduce the Euclidean problem of $r$-nets on unit vectors to finding an $r$-net for unit vectors under inner product (Section~\ref{SGeneral}). This step requires that the multiplicative $1+\epsilon$ approximation of the distance corresponds to an additive $c\epsilon$ approximation of the inner product, for suitable constant $c>0$. Next, we convert the vectors having unit norm into vectors with entries $\{-1, +1\}$ (Section \ref{SInner}). This transformation is necessary in order to apply the Chebyshev embedding of \cite{Val15}, an embedding that damps the magnitude of the inner product of ``far" vectors, while preserving the magnitude of the inner product of ``close" vectors. For the final step of the algorithm, we first apply a procedure that allows us to efficiently compute $(1+\epsilon)$-nets in the case where the number of ``small" distances is large. Then, we apply a modified version of the {\tt Vector Aggregation} algorithm of \cite{Val15}, that exploits fast matrix multiplication, so as to achieve the desired running time. In short, we extend Valiant's framework \cite{Val15} and we compute $r$-nets in time $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$, thus improving on the exponent of the LSH-based construction \cite{EHS15}, when $\epsilon$ is small enough. This improvement by $\sqrt{\epsilon}$ in the exponent is the same as the complexity improvement obtained in \cite{Val15} over the LSH-based algorithm for the approximate closest pair problem. Our study is motivated by the observation that computing efficiently an $r$-net leads to efficient solutions for several geometric problems, specifically in approximation algorithms. In particular, our extension of $r$-nets in high dimensional Euclidean space can be plugged in the framework of~\cite{HR15}. The new framework has many applications, notably the $k$th nearest neighbor distance problem, which we solve in $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$. \paragraph{Paper Organization.} Section \ref{SInner} presents an algorithm for computing an approximate net with respect to the inner product for a set of unit vectors. Section~\ref{SGeneral} translates the problem of finding $r$-nets under Euclidean distance to the same problem under inner product. In Section \ref{Sapps}, we discuss applications of our construction and possible future work. Omitted proofs are included in the Appendices. We use $\norm{\cdot}$ to denote the Euclidean norm $\norm{\cdot}_2$ throughout the paper. \section{Points on a sphere under inner product}\label{SInner} In this section, we design an algorithm for constructing an approximate $\rho$-net of vectors on the sphere under inner product. To that end, we reduce the problem to constructing an approximate net under absolute inner product for vectors that lie on the vertices of a unit hypercube. Since our ultimate goal is a solution to computing $r$-nets with respect to Euclidean distance, we allow additive error in the approximation, which under certain assumptions, translates to multiplicative error in Euclidean distance. In the following, we define rigorously the notion of approximate $\rho$-nets under inner product. \begin{dfn} \label{DfnNetInn} For any $X\subset {\mathbb S}^{d-1}$, an approximate $\rho$-net for $(X,\langle \cdot,\cdot\rangle)$ , with additive approximation parameter $\epsilon>0$, is a subset $C\subseteq X$ which satisfies the following properties: \begin{itemize} \item for any two $p \neq q \in C$, $\langle p, q \rangle < \rho$, and \item for any $x\in X$, there exists $p \in C$ s.t. $\langle x, p \rangle \geq \rho-\epsilon$. \end{itemize} \end{dfn} One relevant notion is that of $\epsilon$-kernels \cite{AHV05}. In $\epsilon$-kernels, one is interested in finding a subset of the input pointset, which approximates its directional width. Such constructions have been extensively studied when the dimension is low, due to their relatively small size. \subsection{Crude approximate nets} In this subsection we develop our basic tool, which is based on the Vector Aggregation Algorithm by \cite{Val15}. This tool aims to compute approximate $\rho$-nets with multiplicative error, as opposed to what we have set as our final goal for this section, namely to bound additive error. Moreover, in the context of this subsection, two vectors are close to each other when the magnitude of their inner product is large, and two vectors are far from each other when the magnitude of their inner product is small. Let $|\langle \cdot,\cdot \rangle|$ denote the magnitude of the inner product of two vectors. \begin{dfn} \label{DfnMagnInnNet} For any $X=[x_1,\ldots, x_n], X' =[x_1',\ldots,x_n'] \subset {\mathbb R}^{d \times n}$, a crude approximate $\rho$-net for $(X,X',|\langle \cdot,\cdot \rangle|)$, with multiplicative approximation factor $c>1$, is a subset $C\subseteq [n]$ which satisfies the following properties: \begin{itemize} \item for any two $i \neq j \in C$, $|\langle x_i, x_j' \rangle| < c \rho$, and \item for any $i\in [n]$, there exists $j \in C$ s.t.\ $|\langle x_i, x_j' \rangle| \geq \rho$. \end{itemize} \end{dfn} {\tt Vector Aggregation} follows the exposition of \cite{Val15}. The main difference is that, instead of the ``compressed'' matrix $Z^T Z$, we use the form $X^T Z$, where $Z$ derives from vector aggregation. Both forms encode the information in the Gram matrix $X^T X$. The matrix $X^TZ$ is better suited for our purposes, since each row corresponds to an input vector instead of an aggregated subset; this extra information may be useful in further problems. \begin{framed} {\tt Vector Aggregation}\\ Input: $X =[x_1,\ldots,x_n] \in {\mathbb R}^{d \times n}$, $X' =[x_1',\ldots,x_n'] \in {\mathbb R}^{d \times n}$, $\alpha \in (0,1)$, $\tau>0$. Output: $n\times n^{1-\alpha}$ matrix $W$ and random partition $S_1 , \ldots , S_{n^{1-\alpha}}$ of $\{x_1,\ldots,x_n\}$. \begin{itemize} \item Randomly partition $[n]$ into $n^{1-\alpha}$ disjoint subsets, each of size $n^{\alpha}$ , denoting the sets $S_1 , \ldots , S_{n^{1-\alpha}}$. \item For each $i = 1, 2, \ldots , 78 \log n$: \begin{itemize} \item Select $n$ coefficients $q_1 ,\ldots , q_n \in \{-1, +1\}$ at random. \item Form the $d\times n^{1-\alpha}$ matrix $Z^i$ with entries $z_{j,k}^i=\sum_{l\in S_k} q_l \cdot x_{j,l}'$ \item $W^i=X^T Z^i$ \end{itemize} \item Define the $n \times n^{1-\alpha}$ matrix $W$ with $w_{i,j} =quartile(|w_{i,j}^1|,\ldots|w_{i,j}^{78 \log n}|)$. \item Output $W$ and $S_1 , \ldots , S_{n^{1-\alpha}}$. \end{itemize} \end{framed} \begin{thm}\label{ThmVecAgg} Let $X \in {\mathbb R}^{d \times n}$, $X' \in {\mathbb R}^{d \times n}$, $\alpha \in (0,1)$, $\tau>0$ the input of {\tt Vector Aggregation}. Then, the algorithm returns a matrix $W$ of size $n\times n^{1-\alpha}$ and a random partition $S_1 , \ldots , S_{n^{1-\alpha}}$, which with probability $1-O(1/n^3)$ satisfies the following: \begin{itemize} \item For all $j\in [n]$ and $k\in [n^{1-\alpha}]$, if $\forall u \in S_k$, $ |\langle x_j, u \rangle|\leq \tau$ then $|w_{j,k}| < 3 \cdot n^{\alpha} \tau$. \item For all $j\in [n]$ and $k\in [n^{1-\alpha}]$ if $\exists u\in S_k$, $|\langle x_j, u \rangle|\geq 3n^{\alpha}\tau$ then $|w_{j,k}| \geq 3 \cdot n^{\alpha} \tau$. \end{itemize} Moreover, the algorithm runs in time $\tilde{O}(dn+n^{2-\alpha}+MatrixMul( n \times d,d \times n^{1-\alpha}))$. \end{thm} For the case of pointsets with many ``small" distances, we rely crucially on the fact that the expected number of near neighbors for a randomly chosen point is large. So, if we iteratively choose random points and delete these and their neighbors, we will end up with a pointset which satisfies the property of having sufficiently few ``small" distances. Then, we apply {\tt Vector Aggregation}. \begin{framed} {\tt Crude ApprxNet}\\ Input: $X =[x_1,\ldots,x_n] \in {\mathbb R}^{d \times n}$, $X' =[x_1',\ldots,x_n'] \in {\mathbb R}^{d \times n}$, $\alpha \in (0,1)$, $\tau>0$. Output: $C'\subseteq [n]$, {$F' \subseteq [n]$}. \begin{itemize} \item $C\gets \emptyset$, $F_1 \gets \emptyset, F_2 \gets \{x_1,\ldots,x_n\}$ \item Repeat $n^{0.5}$ times: \begin{itemize} \item Choose a column $x_i$ uniformly at random. \item $C \gets C \cup \{x_i\}$. \item Delete column $i$ from matrix $X$ and column $i$ from matrix $X'$. \item Delete each column $k$ from matrix $X$, $X'$ s.t. $|\langle x_i, x_k' \rangle| \geq \tau$. \item If there is no column $k$ from matrix $X$ s.t. $|\langle x_i, x_k'\rangle| \geq \tau$, then $F_1 \gets F_1 \cup \{x_i\}$ \end{itemize} \item Run {\tt Vector Aggregation} with input $X$, $X'$, $\alpha$, $\tau$ and output $W$, $S_1,\ldots,S_{n^{1-\alpha}}$. \item For each of the remaining columns $i=1,\ldots$: \begin{itemize} \item For any $|w_{i,j}|\geq 3 n^{\alpha} \tau$: \begin{itemize} \item If more than $n^{1.7}$ times in here, output "ERROR". \item Compute inner products between $x_i$ and vectors in $S_j$. For each vector $x_k' \in S_j$ s.t. $x_k' \neq x_i$ and $|\langle x_i,x_k'\rangle|\geq \tau$, delete row $k$ {and $F_2 \gets F_2 \backslash \{ x_i\}$.} \end{itemize} \item $C \gets C \cup \{x_i\} $ \end{itemize} \item Output indices of $C$ {and $F \gets \{F_1 \cup F_2 \}$}. \end{itemize} \end{framed} \begin{thm}\label{ThmCrudeNet} On input $X =[x_1,\ldots,x_n] \in {\mathbb R}^{d \times n}$, $X' =[x_1',\ldots,x_n'] \in {\mathbb R}^{d \times n}$, $\alpha \in (0,1)$, $\tau>0$, {\tt Crude ApprxNet}, computes a crude $3n^{\alpha}$-approximate $\tau$-net for $X$, $X'$, following the notation of Definition \ref{DfnMagnInnNet}. The algorithm costs time: $$ \tilde{O}(n^{2-\alpha}+ d \cdot n^{1.7+\alpha}+MatrixMul( n \times d,d \times n^{1-\alpha})), $$ and succeeds with probability $1-O(1/n^{0.2})$. Additionally, it outputs a set $F\subseteq R$ with the following property: $\{x_i \mid \forall x_j \neq x_i~ |\langle x_j,x_i \rangle | < \tau \}\subseteq F \subseteq \{x_i \mid \forall x_j \neq x_i~ |\langle x_j,x_i \rangle | < n^a\tau \}$. \end{thm} \begin{proof} We perform $n^{0.5}$ iterations and for each, we compare the inner products between the randomly chosen vector and all other vectors. Hence, the time needed is $O(dn^{1.5})$. In the following, we denote by $X_i$ the number of vectors which have ``large" magnitude of the inner product with the randomly chosen point in the $i$th iteration. Towards proving correctness, suppose first that ${\mathbb E}[X_i]>2n^{0.5}$ for all $i=1, \ldots n^{0.5}$. The expected number of vectors we delete in each iteration of the algorithm is more than $2n^{0.5}+1$. So, after $n^{0.5}$ iterations, the expected total number of deleted vectors will be greater than $n$. This means that if the hypothesis holds for all iterations we will end up with a proper net. Now suppose that there is an iteration $j$ where ${\mathbb E}[X_j] \leq 2n^{0.5}$. After all iterations, the number of ``small" distances are at most $n^{1.5}$ on expectation. By Markov's inequality, when the {\tt Vector Aggregation} algorithm is called, the following is satisfied with probability $1-n^{-0.2}$ : $$ |\{(i,k) \mid |\langle x_i, x_k'\rangle| \geq\tau, i\neq k\}| \leq n^{1.7} . $$ By Theorem \ref{ThmVecAgg} and the above discussion, the number of entries in the matrix $W$ that we need to visit is at most $n^{1.7}$. For each entry, we perform a brute force which costs $d n^\alpha$. Now notice that the first iteration stores centers $c$ and deletes all points $p$ for which $|\langle c,p\rangle| \geq \tau$. Hence, any two centers $c,c'$ satisfy $ |\langle c,p\rangle| < \tau$. In the second iteration, over the columns of $W$, notice that by Theorem \ref{ThmVecAgg}, for any two centers $c,c'$ we have $|\langle c,c'\rangle| <3 n^{\alpha}\tau.$ \end{proof} \subsection{Approximate inner product nets} In this subsection, we show that the problem of computing $\rho$-nets for the inner product of unit vectors reduces to the less natural problem of Definition \ref{DfnMagnInnNet}, which refers to the magnitude of the inner product. The first step consists of mapping the unit vectors to vectors in $\{-1,1\}^{d'}$. The mapping is essentially Charikar's LSH scheme \cite{Cha02}. Then, we apply the Chebyshev embedding of~\cite{Val15} in order to achieve gap amplification, and finally we call algorithm {\tt Crude ApprxNet}, which will now return a proper $\rho$-net with additive error. \begin{thm}[\cite{Val15}] \label{ThmUnif} There exists an algorithm with the following properties. Let $d'=O(\frac{\log n}{\delta^2})$ and $Y \in {\mathbb R}^{d'\times n}$ denote its output on input $X$, $\delta$, where $X$ is a matrix whose columns have unit norm, with probability $1 - o(1/n^2 )$, for all pairs $i, j \in [n]$, $ \Big|{\langle Y_i , Y_j \rangle}/{d'}-\Big(1-2 \cdot {\mathrm{cos}^{-1}(\langle X_i,X_j \rangle)}/{ \pi}\Big)\Big| \leq \delta ,$ where $X_i$, $Y_i$ denote the $i$th column of $X$ and $Y$ respectively. Additionally, the runtime of the algorithm is $O( \frac{d n \log n}{\delta^2})$. \end{thm} The following theorem provides a randomized embedding that damps the magnitude of the inner product of ``far" vectors, while preserving the magnitude of the inner product of ``close" vectors. The statement is almost verbatim that of~\cite[Prop.6]{Val15} except that we additionally establish an asymptotically better probability of success. The proof is the same, but since we claim stronger guarantees on success probability, we include the complete proof in Appendix~\ref{AppThmCheb}. \begin{thm}\label{ThmCheb} Let $Y$, $Y'$ be the matrices output by algorithm ``Chebyshev Embedding" on input $X, X' \in \{-1,1\}^{d\times n} , \tau^{+}\in [-1,1] , \tau^{-} \in [-1,1]$ with $\tau^{-}<\tau^{+}$ , integers $q, d'$. With probability {$ 1 - o(1/n)$} over the randomness in the construction of $Y, Y'$, for all $i, j \in [n]$, $\langle Y_i , Y_j' \rangle$ is within $\sqrt{d'} \log n$ from the value $ T_q \Big(\frac{\langle X_i, X_j'\rangle/d' - \tau^{-}}{\tau^{+}-\tau^{-}}2 -1 \Big) \cdot d' \cdot (\tau^{+}-\tau^{-})^q /{2^{3q-1}}, $ where $T_q$ is the degree-$q$ Chebyshev polynomial of the first kind. The algorithm runs in time $O(d' \cdot n\cdot q)$. \end{thm} \begin{framed} {\tt Inner product ApprxNet}\\ Input: $X =[x_1,\ldots,x_n] $ with each $x_i \in {\mathbb S}^{d-1}$, $\rho \in [-1,1]$, $\epsilon \in (0,1/2]$. Output: Sets $C, F \subseteq [n]$. \begin{itemize} \item If $\rho\leq \epsilon$, then: \begin{itemize} \item $C \gets \emptyset$, $F\gets \emptyset$, $W \gets \{x_1,\ldots,x_n\}$ \item While $W\neq \emptyset$: \begin{itemize} \item Choose arbitrary vector $x\in W$. \item $W \gets W \setminus \{y \in W \mid \langle x,y\rangle \geq \rho-\epsilon \}$ \item $C \gets C \cup \{x\}$ \item If $\forall y \in W$, $\langle x,y \rangle<\rho-\epsilon $ then $F\gets F. \cup \{x\}$ \end{itemize} \item Return indices of $C$, $F$. \end{itemize} \item Apply Theorem \ref{ThmUnif} for input $X$, $\delta=\epsilon/2 \pi$ and output $Y\in \{-1,1\}^{d' \times n}$ for $d'=O(\log n/\delta^2)$. \item Apply Theorem \ref{ThmCheb} for input $Y$, $d''= n^{0.2}$, $q=50^{-1} \log n$, $\tau^-=-1$, $\tau^{+}=1-\frac{2 \cos^{-1}(\rho-\epsilon)}{\pi} +\delta$ and output $Z, Z'$. \item Run algorithm {\tt Crude ApprxNet} with input $\tau=3n^{0.16}$, $\alpha=\sqrt{\epsilon}/500$, $Z,Z'$ and output $C$, $F$. \item Return $C$, $F$. \end{itemize} \end{framed} \begin{thm}\label{AppIPnet} The algorithm {\tt Inner product ApprxNet}, on input $X =[x_1,\ldots,x_n] $ with each $x_i \in {\mathbb S}^{d-1}$, $\rho \in [-1,1]$ and $\epsilon \in (0,1/2]$, computes an approximate $\rho$-net with additive error $\epsilon$, using the notation of Definition \ref{DfnNetInn}. The algorithm runs in time $\tilde{O}(dn+n^{2-\sqrt{\epsilon}/600})$ and succeeds with probability $1-O(1/n^{0.2})$. Additionally, it computes a set $F$ with the following property: $\{x_i \mid \forall x_j \neq x_i~ \langle x_j,x_i \rangle < \rho -\epsilon \}\subseteq F \subseteq \{x_i \mid \forall x_j \neq x_i~ \langle x_j,x_i \rangle < \rho \}$. \end{thm} \section{Approximate nets in high dimensions} \label{SGeneral} In this section, we translate the problem of computing $r$-nets in $({\mathbb R}^d,\|\cdot \|)$ to the problem of computing $\rho$-nets for unit vectors under inner product. One intermediate step is that of computing $r$-nets for unit vectors under Euclidean distance. \subsection{From arbitrary to unit vectors} In this subsection, we show that if one is interested in finding an $r$-net for $({\mathbb R}^d,\|\cdot \|)$, it is sufficient to solve the problem for points on the unit sphere. One analogous statement is used in \cite{Val15}, where they prove that one can apply a randomized mapping from the general Euclidean space to points on a unit sphere, while preserving the ratio of distances for any two pairs of points. The claim derives by the simple observation that an $r$-net in the initial space can be approximated by computing an $\epsilon r/c$-net on the sphere, where $c$ is the maximum norm of any given point envisaged as vector. Our exposition is even simpler since we can directly employ the analogous theorem from \cite{Val15}. \begin{cor}\label{Standardize} There exists an algorithm, {\tt Standardize}, which, on input a $d \times n$ matrix $X$ with entries $x_{i,j} \in {\mathbb R}$, a constant $\epsilon \in (0, 1)$ and a distance parameter $r \in {\mathbb R}$, outputs a $m'\times n $ matrix $Y$, with columns having unit norm and $m'=\log^3n$, and a distance parameter $\rho \in {\mathbb R} $, such that a $\rho$-net of $Y$ is an approximate $r$-net of $X$, with probability $1-o(1/poly(n))$. \end{cor} \begin{comment} \begin{framed} \textbf{Standardize}\\ Input: A $d \times n$ matrix $X$ with entries $x_{i,j} \in {\mathbb R}$, a constant $\epsilon \in (0, 1)$, a distance parameter $r \in {\mathbb R}$. Output: A $m'\times n $ matrix $Y$ with columns having unit norm and $m'=\log^3n$, and a distance parameter $\rho \in {\mathbb R} $, such that our algorithm computes an $r$-net of $X$ given a $\rho$-net of $Y$. \begin{itemize} \item Define two $d$-dimensional vectors $X_{n+1}, X_{n+2}$, s.t.\ $r'=X_{n+1}-X_{n+2}$ and $\|r'\|=r$, and let matrix $X'$ denote the concatenation of $X$, $X_{n+1}$ and $X_{n+2}$ with size $d\times n+2$. \item Perform a Johnson-Lindenstrauss transformation on the columns of $X'$, projecting them to dimension $m'$, so as to yield matrix $X''$. \item Let $c$ denote the magnitude of the largest column of $X''$. Choose a random $m'$-dimensional vector $u$ of magnitude $8c/\epsilon$. \item Let matrix $Y$ be the result of adding $u$ to each column of $X''$ and normalizing all columns so as to have unit norm. \item Define $\rho:=\|Y_{n+1}-Y_{n+2}\|$ to be the new distance parameter. \end{itemize} \end{framed} \end{comment} \subsection{Approximate nets under Euclidean distance} In this subsection, we show that one can translate the problem of computing an $r$-net for points on the unit sphere under Euclidean distance, to finding an $r$-net for unit vectors under inner product as defined in Section \ref{SInner}. Moreover, we identify the subset of the $r$-net which contains the centers that are approximately far from any other point. Formally, \begin{dfn} Given a set of points $X$ and $\epsilon>0$, a set $F\subseteq X$ of $(1+\epsilon)$-approximate $r$-far points is defined by the following property: $\{x\in X \mid \forall x \neq y \in X ~ \|x-y\| > (1+\epsilon)r \}\subseteq F \subseteq \{x\in X \mid \forall x \neq y \in X ~ \|x-y\| > r \}$. \end{dfn} If $r$ is greater than some constant, the problem can be immediately solved by the law of cosines. If $r$ cannot be considered as constant, we distinguish cases $r\geq 1/n^{0.9}$ and $r <1/n^{0.9}$. The first case is solved by a simple modification of an analogous algorithm in \cite[p.13:28]{Val15}. The second case is not straightforward and requires partitioning the pointset in a manner which allows computing $r$-nets for each part separately. Each part has bounded diameter which implies that we need to solve a ``large $r$" subproblem. \begin{thm}\label{ThmLargeRadius} There exists an algorithm, {\tt ApprxNet(Large radius)}, which, for any constant $\epsilon\in (0,1/2]$, $X\subset {\mathbb S}^{d-1}$ s.t. $|X|=n$, outputs a $(1+\epsilon)r$-net and a set of $(1+\epsilon)$-approximate $r$-far points with probability $1-O(1/n^{0.2})$. Additionally, provided $r>1/n^{0.9}$ the runtime of the algorithm is $\tilde{O}(dn^{2-\Theta(\sqrt[]{\epsilon})})$. \end{thm} Let us now present an algorithm which translates the problem of finding an $r$-net for $r<1/n^{0.9}$ to the problem of computing an $r$-net for $r\geq 1/n^{0.9}$. The main idea is that we compute disjoint subsets $S_i$, which are far enough from each other, so that we can compute $r$-nets for each $S_i$ independently. We show that for each $S_i$ we can compute $T_i \subseteq S_i$ which has bounded diameter and $T_i'\subseteq S_i$ such that $T_i$, $T_i'$ are disjoint, each point in $T_i$ is far from each point in $T_i'$, and $|T_i'|\leq 3|S_i|/4$. It is then easy to find $r$-nets for $T_i$ by employing the ApprxNet(Large radius) algorithm. Then, we recurse on $T_i'$ which contains a constant fraction of points from $|S_i|$. Then, we cover points in $S_i \setminus(T_i \cup T_i')$ and points which do not belong to any $S_i$. \begin{framed} {\tt ApprxNet(Small radius)}\\ Input: $X =[x_1,\ldots,x_n]^T $ with each $x_i \in {\mathbb S}^{d-1}$, $r< 1/n^{0.9}$, $\epsilon \in (0,1/2]$. Output: Sets $R, F \subseteq [n]$. \begin{enumerate} \item Project points on a uniform random unit vector and consider projections $p_1,\ldots,p_n$ which wlog correspond to $x_1,\ldots,x_n\in {\mathbb R}^d$. \item Traverse the list as follows \begin{itemize} \item If $|\{j \mid p_j \in [p_i-r,p_i] \}| \leq n^{0.6}$ or $i=n$: \begin{itemize} \item If $|\{j \mid p_j <p_i \}| \leq n^{0.9}$ remove from the list all points $p_j$ s.t. $p_j<p_i-r$ and save set $K=\{x_j \mid p_j\in [p_i-r,p_i] \}$. \item If $|\{j \mid p_j <p_i\}| > n^{0.9}$ save sets $K_i=\{x_j \mid p_j\in [p_i-r,p_i] \} \cup K$, $S_i=\{x_j\mid p_j<p_i-r\}\setminus K$ and remove projections of $S_i$ and $K_i$ from the list. \end{itemize} \end{itemize} \item After traversing the list if we have not saved any $S_i$ go to 5; otherwise for each $S_i$: \begin{itemize} \item For each $u\in S_i$, sample $n^{0.1}$ distances between $u$ and randomly chosen $x_k\in S_i$. Stop if for the selected $u\in S_i$, more than $1/3$ of the sampled points are in distance $\leq r n^{0.6}$. This means that one has found $u$ s.t. $|\{x_k \in S_i, \|u-x_k\|\leq r n^{0.6}\}| \geq |S_i|/4$ with high probability. If no such point was found, output "ERROR". \item Let $0\leq d_1\leq \ldots\leq d_{|S_i|}$ be the distances between $u$ and all other points in $S_i$. Find $c\in[ r n^{0.6},2r n^{0.6}]$ s.t. $|\{j \in [n] \mid d_j \in [c,c+r] \}| <n^{0.4}$, store $W_i=\{x_j \mid d_j \in [c,c+r] \}$, and remove $W_i$ from $S_i$. \item Construct the sets $T_i=\{x_j \in S_i \mid d_j<c\}$ and $T_i'=\{x_j \in S_i \mid d_j > c+r\}$. \begin{itemize} \item For $T_i$, subtract $u$ from all vectors in $T_i$, run {\tt Standardize}, then {\tt ApprxNet (Large radius)}, both with {$\epsilon/4$}. Save points which correspond to output at $R_i$, $F_i$ respectively. \item Recurse on $T_i'$ the whole algorithm, and notice that $|T_i'|\leq 3 |S_i|/4$. Save output at $R_i'$, and $F_i'$ respectively. \end{itemize} \end{itemize} \item Let $R \gets \bigcup_i R_i \cup R_i'$ and $F \gets \bigcup_i F_i \cup F_i'$. Return to the list $p_1,\ldots,p_n$. \begin{enumerate} \item {Remove from $F$ all points which cover at least one point from $\bigcup_i W_i$ or $\bigcup_i K_i$.} \item \label{itm:deleteb} Delete all points $ (\bigcup_i T_i) \setminus (\bigcup_i R_i)$, and $ (\bigcup_i T_i') \setminus (\bigcup_i R_i')$. \item \label{itm:deletec}For each $i$ delete all points in $W_i$ covered by $R_i$, or covered by $R_i'$. \item \label{itm:deleted}For each $i$ delete all points in $K_i$ covered by $R$. \item Finally delete $R$ from the list. Store the remaining points at $F'$. \end{enumerate} \item $R' \gets \emptyset$. Traverse the list as follows: For each $p_i$, check the distances from all $x_j$ s.t. $p_j\in [p_i-r,p_i]$. \begin{itemize} \item If $\exists\, x_j \in R' :$ $\|x_i-x_j\| \leq r$, delete $x_i$ from the list, set $F' \gets F' \backslash \{x_i , x_j\}$ and continue traversing the list. \item If there is no such point $x_j$ then $R \gets R \cup \{x_i\}$ and continue traversing the list. \end{itemize} \item Output indices of $R\gets R \cup R'$ and $F \gets F \cup F'$. \end{enumerate} \end{framed} \begin{thm}\label{ThmSmallr} For any constant $\epsilon>0$, $X\subset {\mathbb S}^{d-1}$ s.t. $|X|=n$, and $r < 1/n^{0.9}$, {\tt ApprxNet(Small radius)} will output a $(1+\epsilon)r$-net and a set of $(1+\epsilon)$-approximate $r$-far points in time $\tilde{O}(dn^{2-\Theta(\sqrt[]{\epsilon})})$, with probability $1-o(1/n^{0.04})$. \end{thm} \begin{proof} Note that points in $S_i$ had projections $p_i$ in sets of contiguous intervals of width $r$; each interval had $\geq n^{0.6}$ points, hence the diameter of the projection of $S_i$ is $\leq n^{0.4}r$. By the Johnson Lindenstrauss Lemma \cite{DG02} we have that for $v \in {\mathbb S}^{d-1}$ chosen uniformly at random: $${\mathrm Pr}\Big[\langle u,v\rangle^{2}\leq \frac{\|u\|^2}{n^{0.4}}\Big]\leq \frac{\sqrt{d} \sqrt{e}}{n^{0.2}}. $$ Hence, $ {\mathbb E}[| \{ x_k,x_j\in S_i \mid \|x_k-x_j\| \geq n^{0.6}r \text{ and } \|p_k-p_j\|\leq n^{0.4} r\}|] \leq |S_i|^2 \cdot \frac{\sqrt{e d}}{n^{0.2}}, $ and the probability $$ {\mathrm Pr}[| \{ x_k,x_j\in S_i \mid \|x_k-x_j\| \geq n^{0.6}r \text{ and } \|p_k-p_j\|\leq n^{0.4} r\}| \geq |S_i|^{1.95}] \leq |S_i|^{0.05} \cdot \frac{\sqrt{e d}}{n^{0.2}}\leq \frac{\sqrt{e d}}{n^{0.15}}. $$ Taking a union bound over all sets $S_i$ yields a probability of failure $o({1}/{n^{0.045}})$. This implies that (for large enough $n$, which implies large enough $|S_i|$) at least $${\binom{|S_i|}{2}} -|S_i|^{1.95}\geq {\frac{|S_i|^2}{4}}$$ distances between points in $S_i$ are indeed small ($\leq n^{0.6}r$). Hence, there exists some point $p_k \in S_i$ which $(n^{0.6}r)$-covers $|S_i|/2$ points. For each possible $p_k$ we sample $n^{0.1}$ distances to other points, and by {Chernoff bounds}, if a point $(n^{0.6}r)$-covers a fraction of more than $1/2$ of the points in $S_i$, then it covers more than $n^{0.1}/3$ sampled points with high probability. Similarly, if a point $(n^{0.6}r)$-covers a fraction of less than $1/4$ of the points in $S_i$, then it covers less than $n^{0.1}/3$ sampled points with high probability. More precisely, for some fixed $u\in S_i$, let $X_j=1$ when for the $j$th randomly chosen point $v\in S_i$, it holds $\| u-v\| \leq n^{0.6}r$ and let $X_j=0$ otherwise. Then, for $Y=\sum_{j=1}^{n^{0.1}} X_j$, it holds: $$ {\mathbb E}[Y]\geq n^{0.1}/2 \implies {\mathrm Pr}[Y\leq n^{0.1}/3 ]\leq \exp(- \Theta(n^{0.1})), $$ $$ {\mathbb E}[Y]\leq n^{0.1}/4 \implies {\mathrm Pr}[Y\geq n^{0.1}/3]\leq \exp(- \Theta(n^{0.1})). $$ Since for any point $x\in T_i$ and any point $y \in T_i'$ we have $\|x-y\|>r$, the packing property of $r$-nets is preserved when we build $r$-nets for $T_i$ and $T_i'$ independently. For each $T_i$, we succeed in building $r$-nets with probability $1-O(1/n^{0.2})$. By a union bound over all sets $T_i$, we have a probability of failure $O(1/n^{0.1})$. Furthermore, points which belong to sets $W_i$ and $K_i$ are possibly covered and need to be checked. For the analysis of the runtime of the algorithm, notice that step \ref{itm:deleteb} costs time $O(d\cdot (\sum_i|T_i|+\sum_i|T_i'|))=O(dn)$. Then, step \ref{itm:deletec} costs time $O(d \cdot \sum_i |W_i|\cdot |T_i|+d \cdot \sum_i |W_i|\cdot |T_i'|)=O(d n^{1.4})$. Finally, notice that we have at most $n^{0.1}$ sets $K_i$. Each $K_i$ contains at most $2n^{0.6}$ points, hence checking each point in $\bigcup_i K_i$ with each point in $R$ costs $O(d n^{1.7})$. Now regarding step 5, consider any interval $[p_i-r,p_i]$ in the initial list, where all points are projected. If $|\{ j \mid p_j \in [p_i-r,p_i]\}\leq 2 n^{0.9}$ then the $i$th iteration in step 5 will obviously cost $O(n^{0.9})$, since previous steps only delete points. If $|\{ j \mid p_j \in [p_i-r,p_i]\}> 2 n^{0.9}$, we claim that $|\{j<i \mid p_j \in [p_i-r,p_i] \text{ and } K_j \text{ is created}\}| \leq 1$. Consider the smallest $j <i$ s.t. $K_j$ is created and $p_j\in [p_i-r,p_i]$. This means that all points $p_k$, for $k\leq j$, are deleted when $p_j$ is visited. Now assume that there exists integer $l \in (j,i)$ s.t. $K_l$ is created. This means that the remaining points in the interval $[p_l-r,p_l]$ are $\leq n^{0.6}$ and all of the remaining points $p_k <p_l$ are more than $n^{0.9}$. This leads to contradiction, since by the deletion in the $j$th iteration, we know that all of the remaining points $p_k <p_l$ lie in the interval $[p_l-r,p_l]$. Now, assume that there exists one $ j<i$ s.t. $p_j \in [p_i-r,p_i]$ and $K_j$ is created. Then, when $p_i$ is visited, there at least $2 n^{0.9}-n^{0.6}>n^{0.9}$ remaining points in the interval $[p_i-r,p_i]$. Hence, there exists $l\geq i$ for which the remaining points in the interval $[p_i-r,p_i]$ are contained in $S_l \cup K_l$. Hence in this case, in step 5, there exist at most $O(n^{0.6})$ points which are not deleted and belong to the interval $[p_i-r,p_i]$. Now assume that there does not exist any $ j<i$ s.t. $p_j \in [p_i-r,p_i]$ and $K_j$ is created. This directly implies that there exists $l\geq i$ for which the remaining points in the interval $[p_i-r,p_i]$ are contained in $S_l \cup K_l$. At last, the total time of the above algorithm is dominated by the calls to the construction of the partial $r$-nets of the sets $T_i$. Thus, the total running time is $O(\sum_{ i} {|T_i|}^{2-\Theta(\sqrt{\epsilon})}+\sum_{ i} {|T_i|'}^{2-\Theta(\sqrt{\epsilon})})= O(\sum_{ i} {|T_i|}^{2-\Theta(\sqrt{\epsilon}})+\sum_{i} {(3|T_i|/4)}^{2-\Theta(\sqrt{\epsilon})})= \tilde{O}(n^{2-\Theta(\sqrt{\epsilon}))}).$ {Finally, taking a union bound over all recursive calls of the algorithm we obtain a probability of failure $o(1/n^{0.04})$.} \end{proof} We now present an algorithm for an $(1+\epsilon)r$-net for points in ${\mathbb R}^d$ under Euclidean distance. \begin{framed} {\tt ApprxNet}\\ Input: Matrix $X=[x_1,\ldots,x_n]$ with each $x_{i} \in {\mathbb R}^d$, parameter $r \in {\mathbb R}$, constant $\epsilon \in (0, 1/2]$. Output: $R \subseteq \{x_1,\ldots,x_n\}$ \begin{itemize} \item Let $Y$, $r'$ be the output of algorithm {\tt Standardize} on input $X$, $r$ with parameter $\epsilon/4$. \item If $r \geq 1/n^{0.9}$ run {\tt ApprxNet(Large radius)} on input $Y$, $\epsilon/4, r'$ and return points which correspond to the set $R$. \item If $r < 1/n^{0.9}$ run {\tt ApprxNet(Small radius)} on input $Y$, $\epsilon/4, r'$ and return points which correspond to the set $R$. \end{itemize} \end{framed} \begin{thm}\label{ApprxNet} Given $n$ points in ${\mathbb R}^d$, a distance parameter $r \in {\mathbb R}$ and an approximation parameter $\epsilon \in (0, 1/2]$, with probability $1-o(1/n^{0.04})$, {\tt ApprxNet} will return a $(1+\epsilon)r-net$, $R$, in $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$ time. \end{thm} \begin{proof} The theorem is a direct implication of Theorems \ref{ThmLargeRadius}, \ref{ThmSmallr}, \ref{Stand}. \end{proof} \begin{thm}\label{DelFar} Given $X\subset{\mathbb R}^d$ such that $|X|=n$, a distance parameter $r \in {\mathbb R}$ and an approximation parameter $\epsilon \in (0, 1/2]$, there exists an algorithm, {\tt DelFar}, that will return, with probability $1-o(1/n^{0.04})$, a set $F'$ with the following properties in $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$ time: \begin{itemize} \item If for a point $p \in X$ it holds that $\forall q\neq p, q \in X$ we have $\|p-q\| > (1+\epsilon)r$, then $p \notin F'$. \item If for a point $p \in X$ it holds that $\exists q\neq p, q \in X$ s.t. $\|p-q\| \leq r$, then $p \in F'$. \end{itemize} \end{thm} \section{Applications and Future work}\label{Sapps} Concerning applications, in \cite{HR15}, they design an approximation scheme, which solves various distance optimization problems. The technique employs a grid-based construction of $r$-nets which is linear in $n$, but exponential in $d$. The main prerequisite of the method is the existence of a linear-time decider (formally defined in Appendix~\ref{Aframework}). The framework is especially interesting when the dimension is constant, since the whole algorithm costs time linear in $n$ which, for some problems, improves upon previously known near-linear algorithms. When the dimension is high, we aim for polynomial dependency on $d$, and subquadratic dependency on $n$. Let us focus on the problem of approximating the {\it $k$th nearest neighbor distance}. \begin{dfn} Let $X\subset {\mathbb R}^d$ be a set of $n$ points, approximation error $\epsilon>0$, and let $d_1\leq \ldots \leq d_n$ be the nearest neighbor distances. The problem of computing an $(1+\epsilon)$-approximation to the {\it $k$th nearest neighbor distance} asks for a pair $x,y\in X$ such that $\|x-y\|\in [(1-\epsilon)d_k,(1+\epsilon)d_k]$. \end{dfn} Now we present an approximate decider for the problem above. This procedure combined with the framework we mentioned earlier, which employs our net construction, results in an efficient solution for this problem in high dimension. \begin{framed} {\tt kth NND Decider}\\ Input: $X \subseteq {\mathbb R}^d$, constant $\epsilon\in (0,1/2]$, integer $k>0$. Output: An interval for the optimal value $f(X, k)$. \begin{itemize} \item Call {\tt DelFar}$(X, \frac{r}{1+\epsilon/4}, \epsilon/4)$ and store its output in $W_1$. \item Call {\tt DelFar}$(X, r, \epsilon/4)$ and store its output in $W_2$. \item Do one of the following: \begin{itemize} \item If $|W_1| > k$, then output $``f(X, k) < r"$. \item If $|W_2| < k$, then output $``f(X, k) > r"$. \item If $|W_1| \leq k$ and $\abs{W_2} \geq k$, then output $``f(X, k) \in [\frac{r}{1+\epsilon/4}, \frac{1+\epsilon/4}r]"$. \end{itemize} \end{itemize} \end{framed} \begin{thm}\label{KND} Given a pointset $X \subseteq {\mathbb R}^d$, one can compute a $(1+\epsilon)$-approximation to the $k$-th nearest neighbor in $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$, with probability $1-o(1)$. \end{thm} To the best of our knowledge, this is the first high dimensional solution for this problem. Setting $k=n$ and applying Theorem \ref{KND} one can compute the {\it farthest nearest neighbor} in $\tilde{O}(dn^{2-\Theta(\sqrt{\epsilon})})$ with high probability. Concerning future work, let us start with the problem of finding a greedy permutation. A permutation $\Pi = <\pi_1, \pi_2,\dots >$ of the vertices of a metric space $(X, \norm{\cdot})$ is a \textit{greedy permutation} if each vertex $\pi_i$ is the farthest in $X$ from the set $\Pi_{i-1} = <{\pi_1,\dots, \pi_{i-1}}>$ of preceding vertices. The computation of $r$-nets is closely related to that of the greedy permutation. The $k$-center clustering problem asks the following: given a set $X \subseteq {\mathbb R}^d$ and an integer $k$, find the smallest radius $r$ such that $X$ is contained within $k$ balls of radius $r$. By \cite{EHS15}, a simple modification of our net construction implies an algorithm for the $(1+\epsilon)$ approximate greedy permutation in time $\tilde{O}(d n^{2-\Theta(\sqrt{\epsilon})} \log \Phi)$ where $\Phi$ denotes the spread of the pointset. Then, approximating the greedy permutation implies a $(2+\epsilon)$ approximation algorithm for $k$-center clustering problem. We expect that one can avoid any dependencies on $\Phi$. \if 0 The Corollaries below follow from Theorem \ref{ApprxNet}, Lemma 3.5\cite{EHS15} and Lemma 2.1\cite{EHS15}. \begin{cor} Let $X$ be a set of $n$ points in ${\mathbb R}^d$, $\epsilon \in (0, 1)$ an error parameter and let $\Phi$ be the spread of the Euclidean metric on $X$. Then, one can compute in $O(dn^{2-\Theta(\sqrt{\epsilon})}\log \Phi)$ expected time a sequence that is a $(1 + \epsilon)$-greedy permutation for the Euclidean metric on $X$, with high probability. \end{cor} \begin{cor} Given a set $X$ of $n$ points in ${\mathbb R}^d$, an integer $k$ and an error parameter $\epsilon \in (0, 1)$, one can compute with high probability a $(2+\epsilon)$-approximation to the optimal $k$-center clustering in $O(dn^{2-\Theta(\sqrt{\epsilon})}\log \Phi)$, where $\Phi$ is the spread of the Euclidean metric on $X$. \end{cor}\fi \subsection*{Acknowledgment.} I.Z.~Emiris acknowledges partial support by the EU H2020 research and innovation programme, under the Marie Sklodowska-Curie grant agreement No 675789: Network ``ARCADES". \newpage \bibliographystyle{alpha}
1e54f8847e4971d0e9efee8d1ef4b0cf27aa096c
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} \emph{Call-by-Push-Value}~\cite{LevyP04} is a class of functional languages generalizing the lambda-calculus in several directions. From the point of view of Linear Logic we understand it as a half-polarized system bearing some similarities with \Eg{}classical Parigot's lambda-mu-calculus, this is why we call it $\CBPV{}$. The main idea of Laurent and Regnier interpretation of call-by-name lambda-mu in Linear Logic~\cite{LaurentRegnier03} (following actually~\cite{Girard91a}) is that all types of the minimal fragment of the propositional calculus (with $\Rightarrow$ as unique connective) are interpreted as \emph{negative} types of Linear Logic which are therefore naturally equipped with structural morphisms: technically speaking, these types are algebras of the $\wn$-monad of Linear Logic. This additional structure of negative types allows to perform logical structural rules on the \emph{right side} of typing judgments even if these formulas are not necessarily of shape $\wn\sigma$, and this is the key towards giving a computational content to classical logical rules, generalizing the fundamental discovery of Griffin on typing $\CALLCC{}$ with Peirce Law~\cite{Griffin90}. From our point of view, the basic idea of $\CBPV{}$ is quite similar, though somehow dual and used in a less systematic way: data types are interpreted as \emph{positive} types of Linear Logic equipped therefore with structural morphisms (as linear duals of negative formulas, they are coalgebras of the $\oc$-comonad) and admit therefore structural rules on the \emph{left side} of typing judgment even if they are not of shape $\EXCL\sigma$. This means that a function defined on a data type can have a \emph{linear function type} even if it uses its argument in a non-linear way: this non-linearity is automatically implemented by means of the structural morphisms the positive data type is equipped with. The basic positive type in Linear Logic is $\EXCL\sigma$ (where $\sigma$ is any type): it is the very idea of Girard's call-by-name translation of the lambda-calculus into Linear Logic to represent the implication type $\sigma\Rightarrow\tau$ by means of the decomposition $\LIMPL{\EXCL\sigma}\tau$. The new idea of $\CBPV{}$ is to generalize this use of the linear implication to any type construction of shape $\LIMPL{\phi}{\tau}$ where $\phi$ is a positive type, without imposing any linearity restriction on the usage of the argument of type $\phi$ used by a function of type $\LIMPL{\phi}{\tau}$. This non-symmetrical restriction in the use of the linear implication motivates our description of $\CBPV$ as a ``half-polarized'' system: in a fully polarized system like Laurent's \emph{Polarized Linear Logic} LLP~\cite{LaurentRegnier03}, one would also require the type $\sigma$ to be negative in $\LIMPL\phi\sigma$ (the last system presented in~\cite{Ehrhard16a} implements this idea) and the resulting formalism would host classical computational primitives such as $\CALLCC$ as well. The price to pay, as illustrated in~\cite{AminiEhrhard15}, is a less direct access to data types: it is impossible to give a function from integers to integers the expected type $\LIMPL\Tnat\Tnat$ (where $\Tnat$ is the type of flat natural numbers satisfying $\Tnat=\PLUS\ONE\Tnat$), the simplest type one can give to such a term is $\LIMPL\Tnat{\wn\Tnat}$ which complicates its denotational interpretation\footnote{One can also consider $\wn$ as the computational monad of linear continuations and use a translation from direct style into monadic style (which, for this monad, is just a version of the familiar CPS translation). This is just a matter of presentation and of syntactic sugar and does not change the denotational interpretation in the kind of concrete models of Linear Logic we have in mind such as the relational model, the coherence space model etc.}. Not being polarized on the right side of implications, $\CBPV$ remains ``intuitionistic'' just as standard functional programming languages whose paradigmatic example is $\PCF$. So what is the benefit of this special status given to positive formulas considered as ``data types''? There are several answers to this question. \begin{itemize} \item First, and most importantly, it gives a \emph{call-by-value access} to data types: when translating $\PCF$ into Linear Logic, the simplest type for a function from integers to integers is $\LIMPL{\EXCL\Tnat}{\Tnat}$. This means that arguments of type $\Tnat$ are used in a call-by-name way: such arguments are evaluated again each time they are used. This can of course be quite inefficient. It is also simply \emph{wrong} if we extend our language with a random integer generator since in that case each evaluation of such an argument can lead to a different value: in $\PCF$ there is no way to keep memory of the value obtained for one evaluation of such a parameter and probabilistic programming is therefore impossible. In $\CBPV$ data types such as $\Tnat$ can be accessed in call-by-value, meaning that they are evaluated once and that the resulting value is kept for further computation: this is typically the behavior of a function of type $\LIMPL\Tnat\Tnat$. This is not compulsory however and an explicit $\oc$ type constructor still allows to define functions of type $\LIMPL{\EXCL\Tnat}{\Tnat}$ in $\CBPV$, with the usual $\PCF$ behavior. \item Positive types being closed under positive Linear Logic connectives (direct sums and tensor product) and under ``least fixpoint'' constructions, it is natural to allow corresponding constructions of positive types in $\CBPV$ as well, leading to a language with rich data type constructions (various kinds of trees, streams etc are freely available) and can be accessed in call-by-value as explained above for integers. From this data types point of view, the $\oc$ Linear Logic connective corresponds to the type of \emph{suspensions} or \emph{thunks} which are boxes (in the usual Linear Logic sense) containing unevaluated pieces of program. \item As already mentioned, since the Linear Logic type constructors $\LIMPL{}{}$ and $\EXCL{}$ are available in $\CBPV$ (with the restriction explained above on the use of $\LIMPL{}{}$ that the left side type must be positive), one can represent in $\CBPV$ both Girard's translations from lambda-calculus into Linear Logic introduced in~\cite{Girard87}: the usual one which is call-by-name and the ``boring one'' which is call-by-value. So our language $\CBPV$ is not intrinsically call-by-value and hosts very simply Girard's representation of call-by-name in Linear Logic as further explained in Section~\ref{sec:PCF-products}. \end{itemize} Concretely, in $\CBPV$, a term of positive type can be a \emph{value}, and then it is discardable and duplicable and, accordingly, its denotational interpretation is a morphism of coalgebras: values are particular terms whose interpretation is easily checked to be such a morphism, which doesn't preclude other terms of positive type to have the same property of course, in particular terms reducing to values! Being a value is a property which can be decided in time at most the size of the term and values are freely duplicable and discardable. The ``$\beta$-rules'' of the calculus (the standard $\beta$-reduction as well as the similar reduction rules associated with tensor product and direct sum) are subject to restrictions on certain subterms of redexes to be values (because they must be duplicable and discardable) and these restrictions make sense thanks to this strong decidability property of being a value. \subsection*{Probabilities in $\CBPV$.} Because of the possibility offered by $\CBPV$ of handling values in a call-by-value manner, this language is particularly suitable for probabilistic functional programming. Contrarily to the common monadic viewpoint on effects, we consider an extension of the language where probabilistic choice is a primitive $\COIN p$ of type $\PLUS\ONE\ONE$ (the type of booleans)\footnote{An not of type $T(\PLUS\ONE\ONE)$ where $T$ would be a computational monad of probabilistic computations.} parameterized by $p\in[0,1]\cap\Rational$ which is the probability of getting $\True$ (and $1-p$ is the probability of getting $\False$). So our probabilistic extension $\pCBPV$ of $\CBPV$ is in direct style, but, more importantly, the denotational semantics we consider is itself in ``direct style'' and does not rely on any auxiliary computational monad of probability distributions~\cite{Saheb-Djahromi80,JonesPlotkin89} (see~\cite{JungTix98} for the difficulties related with the monadic approach to probabilistic computations), random variables~\cite{phdBarker,goubaultvarraca11,Mislove16}, or measures~\cite{StatonYWHK16,HeunenKSY17}. On the contrary, we interpret our language in the model of \emph{probabilistic coherence spaces}~\cite{DanosEhrhard08} that we already used for providing a fully abstract semantics for probabilistic $\PCF$~\cite{EhrhardPaganiTasson14}. A probabilistic coherence space $X$ is given by an at most countable set $\Web X$ (the web of $X$) and a set $\Pcoh X$ of $\Web X$-indexed families of non-negative real numbers, to be considered as some kind of ``generalized probability distributions''. This set of families of real numbers is subject to a closure property implementing a simple intuition of probabilistic observations. Probabilistic coherence spaces are a model of classical Linear Logic and can be seen as $\omega$-continuous domains equipped with an operation of convex linear combination, and the linear morphisms of this model are exactly the Scott continuous functions commuting with these convex linear combinations. Besides, probabilistic coherence spaces can be seen as particular \emph{d-cones}~\cite{TixKP09a} and even \emph{Kegelspitzen}~\cite{KeimelP16}, that is, complete partial orders equipped with a Scott continuous ``convex structure'' allowing to compute probabilistic linear combinations of their elements. Kegelspitzen have been used recently by Rennela to define a denotational model of a probabilistic extension of FPC~\cite{Rennela16}. The main difference with respect to our approach seems to be the fact that non-linear morphisms (corresponding to morphisms of type $\Limpl{\Excl\sigma}{\tau}$ in our setting) are general Scott continuous functions in Rennela's model\footnote{More precisely, in his interpretation of FPC, Rennela uses strict Scott continuous functions, but, along the same lines, it seems clear that Kegelspitzen give rise to a model of PCF where morphisms are general Scott continuous functions.}, whereas they are analytic functions\footnote{Meaning that it is definable by a power series.} in ours, which can be seen as very specific Scott continuous functions. See also~\cite{EhrhardPaganiTasson18} where these functions are seen to be \emph{stable} in a generalized sense. As shown in~\cite{DanosEhrhard08} probabilistic coherence spaces have all the required completeness properties for interpreting recursive type definitions (that we already used in~\cite{EhrhardPaganiTasson11} for interpreting the pure lambda-calculus) and so we are able to associate a probabilistic coherence space with all types of $\pCBPV$. In this model the type $\PLUS\ONE\ONE$ is interpreted as the set of sub-probability distributions on $\{\True,\False\}$ so that we have a straightforward interpretation of $\COIN p$. Similarly the type of flat integers $\Tnat$ is interpreted as a probabilistic coherence space $\Snat$ such that $\Web\Snat=\Nat$ and $\Pcoh\Snat$ is the set of all probability distributions on the natural numbers. Given probabilistic spaces $X$ and $Y$, the space $\Limpl XY$ has $\Web X\times\Web Y$ as web and $\Pcoh{(\Limpl XY)}$ is the set of all $\Web X\times\Web Y$ matrices which, when applied to an element of $\Pcoh X$ gives an element of $\Pcoh Y$. The web of the space $\Excl X$ is the set of all finite multisets of elements of $\Web X$ so that an element of $\Limpl{\Excl X}{Y}$ can be considered as a power series on as many variables as there are elements in $\Web X$ (the composition law associated with the Kleisli category of the $\oc$-comonad is compatible with this interpretation of morphisms as power series). From a syntactic point of view, the only values of $\PLUS\ONE\ONE$ are $\True$ and $\False$, so $\COIN p$ is not a value. Therefore we cannot reduce $\LAPP{\ABST x{\PLUS\ONE\ONE}{M}}{\COIN p}$ to $\Subst M{\COIN p}{x}$ and this is a good thing since then we would face the problem that the boolean values of the various occurrences of $\COIN p$ might be different. We have first to reduce $\COIN p$ to a value, and the reduction rules of our probabilistic $\CBPV$ stipulate that $\COIN p$ reduces to $\True$ with probability $p$ and to $\False$ with probability $1-p$ (in accordance with the kind of operational semantics that we considered in our earlier work on this topic, starting with~\cite{DanosEhrhard08}). So $\LAPP{\ABST x{\PLUS\ONE\ONE}{M}}{\COIN p}$ reduces to $\Subst M\True x$ with probability $p$ and to $\Subst M\False x$ with probability $1-p$, which is perfectly compatible with the intuition that in $\CBPV$ application is a linear operation (and that implication is linear: the type of $\ABST x{\PLUS\ONE\ONE}{M}$ is $\LIMPL{(\PLUS\ONE\ONE)}{\sigma}$ for some type $\sigma$): in this operational semantics as well as in the denotational semantics outlined above, linearity corresponds to commutation with (probabilistic) convex linear combinations. \subsection*{Contents.} The results presented in this paper illustrate the tight connection between the syntactic and the denotational intuitions underpinning our understanding of this calculus. We first introduce in Section~\ref{sec:syntax} the syntax and operational semantics of $\pCBPV$, an abstract programming language very close to Paul Levy's Call by Push Value (CBPV)~\cite{LevyP04}. It however differs from Levy's language mainly by the fact that CBPV computation types products and recursive type definitions have no counterparts in our language. This choice is mainly motivated by the wish of keeping the presentation reasonably short. It is argued in Sections~\ref{subsec:exsyn} and~\ref{sec:PCF-products} that $\pCBPV$ is expressive enough for containing well behaved lazy data types such as the type of streams, and for encoding call-by-name languages with products. In Section~\ref{sec:PCS}, we present the Linear Logic model of probabilistic coherence spaces, introducing mainly the underlying linear category $\PCOH$, where $\pCBPV$ general types are interpreted, and the Eilenberg-Moore category $\EM\PCOH$, where the positive types are interpreted. In order to simplify the Adequacy and Full Abstraction proofs, we restrict actually our attention to a well-pointed subcategory of $\EM\PCOH$ whose objects we call ``dense coalgebras'': this will allow to consider all morphisms as functions. As suggested by one of the referees, there are probably smaller well-pointed subcategories of $\EM\PCOH$ where we might interpret our positive types, and in particular the category of families introduced in~\cite{Abramsky1998} describing call-by-value games. This option will be explored in further work. We prefer here to work with the most general setting as it is also compatible with a probabilistic extension of the last system of~\cite{Ehrhard16a}, which features classical $\CALLCC{}$-like capabilities. We prove then in Section~\ref{sec:adequacy} an Adequacy Theorem whose statement is extremely simple: given a closed term $M$ of type $\ONE$ (which has exactly one value $\ONELEM$), the denotational semantics of $M$, which is an element of $[0,1]$, coincides with its probability to reduce to $\ONELEM$ (such a term can only diverge or reduce to $\ONELEM$). In spite of its simple statement the proof of this result requires some efforts mainly because of the presence of unrestricted\footnote{Meaning that recursive type definitions are not restricted to covariant types.} recursive types in $\pCBPV$. The method used in the proof relies on an idea of Pitts~\cite{Pitts93} and is described in the introduction of Section~\ref{sec:adequacy}. Last we prove Full Abstraction in Section~\ref{sec:FA} adapting the technique used in~\cite{EhrhardPaganiTasson11} to the present $\pCBPV$ setting. The basic idea consists in associating, with any element $a$ of the web of the probabilistic coherence space $\Tsem\sigma$ interpreting the type $\sigma$, a term $\Testt a$ of type\footnote{For technical reasons and readability of the proof, the type we give to $\Testt a$ in Section~\ref{sec:FA} slightly differs from this one.} $\LIMPL{\EXCL\sigma}{\LIMPL{\EXCL\Tnat}{\ONE}}$ such that, given two elements $w$ and $w'$ of $\Pcoh{\Tsem\sigma}$ such that $w_a\not=w'_a$, the elements $\Psem{\Testt a}{}\Prom w$ and $\Psem{\Testt a}{}\Prom{(w')}$ of $\Pcoh{(\LIMPL{\EXCL\Tnat}{\ONE})}$ are different power series depending on a finite number $n$ of parameters (this number $n$ depends actually only on $a$) so that we can find a rational number valued sub-probability distribution for these parameters where these power series take different values in $[0,1]$. Applying this to the case where $w$ and $w'$ are the interpretations of two closed terms $M$ and $M'$ of type $\sigma$, we obtain, by combining $\Testt a$ with the rational sub-probability distribution which can be represented in the syntax using $\COIN p$ for various values of $p$, a $\CBPV$ closed term $C$ of type $\LIMPL{\EXCL\sigma}{\ONE}$ such that the probability of convergence of $\LAPP C{\STOP M}$ and $\LAPP C{\STOP{(M')}}$ are different (by adequacy). This proves that if two (closed) terms are operationally equivalent then they have the same semantics in probabilistic coherence spaces, that is, equational full abstraction. \subsection*{Further developments.} These results are similar to the ones reported in~\cite{EhrhardPaganiTasson15} but are actually different, and there is no clear logical connection between them, because the languages are quite different, and therefore, the observation contexts also. And this even in spite of the fact that $\PCF$ can be faithfully encoded in $\CBPV$. This seems to show that the semantical framework for probabilistic functional programming offered by probabilistic coherence spaces is very robust and deserves further investigations. One major outcome of the present work is a natural extension of probabilistic computation to rich data-types, including types of potentially infinite values (streams etc). Our full abstraction result cannot be extended to inequational full abstraction with respect to the natural order relation on the elements of probabilistic coherence spaces: a natural research direction will be to investigate other (pre)order relations and their possible interactive definitions. Also, it is quite tempting to replace the equality of probabilities in the definition of contextual equivalence by a distance; this clearly requires further developments. \section{Probabilistic Call By Push Value}\label{sec:syntax} We introduce the syntax of $\pCBPV$ of CBPV (where HP stands for ``half polarized''). Types are given by the following BNF syntax. We define by mutual induction two kinds of types: \emph{positive types} and \emph{general types}, given type variables $\zeta$, $\xi$\dots: \begin{align} \text{positive}\quad \phi,\psi,\dots{} &\Bnfeq \ONE \Bnfor \EXCL \sigma \Bnfor \TENS\phi\psi \Bnfor \PLUS\phi\psi \Bnfor \zeta \Bnfor \TREC \zeta\phi \label{eq:cbpv-pos-types} \\ \text{general}\quad \sigma,\tau\dots{} &\Bnfeq \FORG\phi \Bnfor \LIMPL\phi\sigma \label{eq:cbpv-gen-types} \end{align} The type $\ONE$ is the neutral element of $\ITens$ and it might seem natural to have also a type $\ZERO$ as the neutral element of $\IPlus$. We didn't do so because there is no associated constructor in the syntax of terms, and the only closed terms of type $\ZERO$ that one can write are ever-looping terms. Observe also that there are no restriction on the variance of types in the recursive type construction: for instance, in $\TREC\zeta\phi$ is a well-formed positive type if $\phi=\EXCL{(\LIMPL\zeta\zeta)}$, where $\zeta$ has a negative and a positive occurrence. Do well notice that our ``positive types'' are positive in the sense of logical polarities, and \emph{not} of the variance of type variables! Terms are given by the following BNF syntax, given variables $x,y,\dots$: \begin{align*} M,N\dots{} \Bnfeq x & \Bnfor \ONELEM \Bnfor \STOP M \Bnfor \PAIR MN \Bnfor \IN \ell M \Bnfor \IN r M \\ & \Bnfor \ABST x\phi M \Bnfor \LAPP MN \Bnfor \CASE M{x_\ell}{N_\ell}{x_r}{N_r}\\ & \Bnfor \PR \ell M \Bnfor \PR r M \Bnfor \GO M \Bnfor \FIXT x\sigma M \\ & \Bnfor \FOLD M \Bnfor \UNFOLD M \\ & \Bnfor \COIN p, \; p\in [0,1]\cap\mathbb Q \end{align*} \begin{remark} This functional language $\pCBPV$, or rather the sublanguage $\CBPV$ which is $\pCBPV$ stripped from the $\COIN p$ construct, is formally very close to Levy's CBPV. As explained in the Introduction, our intuition however are more related to Linear Logic than to CBPV and its general adjunction-based models. This explains why our syntax slightly departs from Levy's original syntax as described \Eg~in~\cite{LevyP02} and is much closer to the SFPL language of~\cite{MarzRohrStreicher99}: Levy's type constructor $F$ is kept implicit and $U$ is ``$\oc$''. We use LL inspired notations: $\STOP M$ corresponds to $\mathsf{thunk}(M)$ and $\GO M$ to $\mathsf{force}(M)$. Our syntax is also slightly simpler than that of SFPL in that our general types do not feature products and recursive types definitions, we will explain in Section~\ref{sec:PCF-products} that this is not a serious limitation in terms of expressiveness. \end{remark} Figure~\ref{fig:typing_system} provides the typing rules for these terms. A typing context is an expression $\cP=(x_1:\phi_1,\dots,x_k:\phi_k)$ where all types are positive and the $x_i$s are pairwise distinct variables. \begin{figure*} \centering \begin{center} \AxiomC{} \UnaryInfC{$\TSEQ{\cP,x:\phi}{x}{\phi}$} \DisplayProof \quad \AxiomC{$\TSEQ{\cP,x:\phi}{M}{\sigma}$} \UnaryInfC{$\TSEQ{\cP}{\ABST x\phi M}{\LIMPL\phi\sigma}$} \DisplayProof \quad \AxiomC{$\TSEQ{\cP}{M}{\LIMPL\phi\sigma}$} \AxiomC{$\TSEQ{\cP}{N}{\phi}$} \BinaryInfC{$\TSEQ{\cP}{\LAPP MN}{\sigma}$} \DisplayProof \end{center} \begin{center} \AxiomC{$\TSEQ\cP M\sigma$} \UnaryInfC{$\TSEQ\cP{\STOP M}{\EXCL\sigma}$} \DisplayProof \quad \AxiomC{} \UnaryInfC{$\TSEQ{\cP}{\ONELEM}{\ONE}$} \DisplayProof \quad \AxiomC{$\TSEQ\cP{M_\ell}{\phi_\ell}$} \AxiomC{$\TSEQ\cP{M_r}{\phi_r}$} \BinaryInfC{$\TSEQ\cP{\PAIR{M_\ell}{M_r}}{\TENS{\phi_\ell}{\phi_r}}$} \DisplayProof \quad \AxiomC{$\TSEQ\cP M{\phi_i}\quad \small i\in\{\ell,r\}$} \UnaryInfC{$\TSEQ\cP{\IN iM}{\PLUS{\phi_\ell}{\phi_r}}$} \DisplayProof \end{center} \begin{center} \AxiomC{$\TSEQ\cP{M}{\EXCL \sigma}$} \UnaryInfC{$\TSEQ\cP{\GO M}{\sigma}$} \DisplayProof \quad \AxiomC{$\TSEQ\cP M{\TENS{\phi_\ell}{\phi_r}\quad \small i\in\{\ell,r\}}$} \UnaryInfC{$\TSEQ\cP{\PR iM}{\phi_i}$} \DisplayProof \quad \AxiomC{$\TSEQ{\cP,x:\EXCL \sigma}{M}{\sigma}$} \UnaryInfC{$\TSEQ\cP{\FIXT x\sigma M}\sigma$} \DisplayProof \end{center} \begin{center} \AxiomC{$\TSEQ\cP M{\PLUS{\phi_\ell}{\phi_r}}$} \AxiomC{$\TSEQ{\cP,x_\ell:\phi_\ell}{M_\ell}\sigma$} \AxiomC{$\TSEQ{\cP,x_r:\phi_r}{M_r}\sigma$} \TrinaryInfC{$\TSEQ\cP{\CASE M{x_\ell}{M_\ell}{x_r}{M_r}}\sigma$} \DisplayProof \end{center} \begin{center} \AxiomC{} \UnaryInfC{$\TSEQ{\cP}{\COIN p}{\PLUS\ONE\ONE}$} \DisplayProof \end{center} \begin{center} \AxiomC{$\TSEQ{\cP}{M}{\Subst{\psi}{\TREC\zeta\psi}{\zeta}}$} \UnaryInfC{$\TSEQ{\cP}{\FOLD M}{\TREC\zeta\psi}$} \DisplayProof \quad \AxiomC{$\TSEQ{\cP}{M}{\TREC\zeta\psi}$} \UnaryInfC{$\TSEQ{\cP}{\UNFOLD M}{\Subst{\psi}{\TREC\zeta\psi}{\zeta}}$} \DisplayProof \end{center} \caption{Typing system for $\pCBPV$} \label{fig:typing_system} \end{figure*} \subsection{Reduction rules} \emph{Values} are particular $\pCBPV$ terms (they are not a new syntactic category) defined by the following BNF syntax: \begin{align*} V,W\dots{} \Bnfeq x \Bnfor \ONELEM \Bnfor \STOP M \Bnfor \PAIR{V}{W} \Bnfor \IN \ell {V} \Bnfor \IN r {V} \Bnfor \FOLD V\,. \end{align*} Figure~\ref{fig:reduction-rules} defines a deterministic \emph{weak} reduction relation $\Wred$ and a probabilistic reduction $\Rel{\Redone p}$ relation. This reduction is weak in the sense that we never reduce within a "box" $\STOP M$ or under a $\lambda$. The distinguishing feature of this reduction system is the role played by values in the definition of $\Wred$. Consider for instance the case of the term $\PR \ell {\PAIR{M_\ell}{M_r}}$; one might expect this term to reduce directly to $M_\ell$ but this is not the case. One needs first to reduce $M_\ell$ \emph{and} $M_r$ to values before extracting the first component of the pair (the terms $\PR \ell {\PAIR{M_\ell}{M_r}}$ and $M_\ell$ have not the same denotational interpretation in general). Of course replacing $M_i$ with $\STOP{M_i}$ allows a lazy behavior. Similarly, in the $\Wred$ rule for $\mathsf{case}$, the term on which the test is made must be reduced to a value (necessarily of shape $\IN \ell V$ or $\IN r V$ if the expression is well typed) before the reduction is performed. As explained in the Introduction this allows to ``memoize'' the value $V$ for further usage: the value is passed to the relevant branch of the $\mathsf{case}$ through the variable $x_i$. Given two terms $M$, $M'$ and a real number $p\in[0,1]$, $M\Rel{\Redone p} M'$ means that $M$ reduces in one step to $M'$ with probability $p$. We say that $M$ is \emph{weak normal} if there is no reduction $M\Rel{\Redone p}M'$. It is clear that any value is weak normal. When $M$ is closed, $M$ is weak normal iff it is a value or an abstraction. In order to simplify the presentation we \emph{choose} in Figure~\ref{fig:reduction-rules} a reduction strategy. For instance we decide that, for reducing $\PAIR{M_\ell}{M_r}$ to a value, one needs first to reduce $M_\ell$ to a value, and then $M_r$; this choice is of course completely arbitrary. A similar choice is made for reducing terms of shape $\LAPP{M}{N}$, where we require the argument to be reduced first. This choice is less arbitrary as it will simplify a little bit the proof of adequacy in Section~\ref{sec:adequacy} (see for instance the proof of Lemma~\ref{lemma:rel-app-closeness}). We could perfectly define a more general weak reduction relation as in~\cite{Ehrhard16a} for which we could prove a ``diamond'' confluence property but we would then need to deal with a reduction transition system where, at each node (term), several probability distributions of reduced terms are available and so we would not be able to describe reduction as a simple (infinite dimensional) stochastic matrix. We could certainly also define more general reduction rules allowing to reduce redexes anywhere in terms (apart for $\COIN p$ which can be reduced only when in linear position) but this would require the introduction of additional $\sigma$-rules as in~\cite{EhrhardGuerrieri16}. As in that paper, confluence can probably be proven, using ideas coming from~\cite{EhrhardRegnier02,Vaux08} for dealing with reduction in an algebraic lambda-calculus setting. \begin{figure*} \centering \begin{center} \AxiomC{} \UnaryInfC{$\GO{\STOP M}\Rel\Wred M$} \DisplayProof \quad \AxiomC{} \UnaryInfC{$\LAPP{\ABST x\phi M}{V}\Rel\Wred\Subst MVx$} \DisplayProof \quad \AxiomC{}\RightLabel{$i\in\{\ell,r\}$} \UnaryInfC{$\PR i{\PAIR{V_\ell}{V_r}}\Rel\Wred V_i$} \DisplayProof \end{center} \begin{center} \AxiomC{} \UnaryInfC{$\FIXT x\sigma M\Rel\Wred\Subst M{\STOP{(\FIXT x\sigma M)}}x$} \DisplayProof \quad \AxiomC{}\RightLabel{$i\in\{\ell, r\}$} \UnaryInfC{$\CASE{\IN iV}{x_\ell}{M_\ell}{x_r}{M_r}\Rel\Wred\Subst{M_i}{V}{x_i}$} \DisplayProof \end{center} \begin{center} \AxiomC{} \UnaryInfC{$\UNFOLD{\FOLD{V}}\Rel{\Wred}V$} \DisplayProof \end{center} \centering \begin{center} \AxiomC{$M\Rel{\Wred}M'$} \UnaryInfC{$M\Rel{\Redone 1}M'$} \DisplayProof \quad \AxiomC{} \UnaryInfC{$\COIN p\Rel{\Redone p} \IN \ell \ONELEM$} \DisplayProof \quad \AxiomC{} \UnaryInfC{$\COIN p\Rel{\Redone{1-p}} \IN r \ONELEM$} \DisplayProof \end{center} \begin{center} \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\GO M\Rel{\Redone p}\GO{M'}$} \DisplayProof \quad \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\LAPP MV\Rel{\Redone p}\LAPP{M'}V$} \DisplayProof \quad \AxiomC{$N\Rel{\Redone p} N'$} \UnaryInfC{$\LAPP MN\Rel{\Redone p}\LAPP{M}{N'}$} \DisplayProof \quad \AxiomC{$M\Rel{\Redone p} M'$}\RightLabel{$i\in\{\ell, r\}$} \UnaryInfC{$\PR iM\Rel{\Redone p}\PR i{M'}$} \DisplayProof \end{center} \begin{center} \AxiomC{$M_\ell\Rel{\Redone p} M'_\ell$} \UnaryInfC{$\PAIR{M_\ell}{M_r}\Rel{\Redone p}\PAIR{M'_\ell}{M_r}$} \DisplayProof \quad \AxiomC{$M_r\Rel{\Redone p} M'_r$} \UnaryInfC{$\PAIR{V}{M_r}\Rel{\Redone p}\PAIR{V}{M'_r}$} \DisplayProof \quad \AxiomC{$M\Rel{\Redone p} M'$} \RightLabel{$i\in\{\ell, r\}$}\UnaryInfC{$\IN iM\Rel{\Redone p}\IN i{M'}$} \DisplayProof \end{center} \begin{center} \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\CASE{M}{x_\ell}{M_\ell}{x_r}{M_r} \Rel{\Redone p}\CASE{M'}{x_\ell}{M_\ell}{x_r}{M_r}$} \DisplayProof \end{center} \begin{center} \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\FOLD M\Rel{\Redone p}\FOLD{M'}$} \DisplayProof \quad \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\UNFOLD M\Rel{\Redone p}\UNFOLD{M'}$} \DisplayProof \end{center} \caption{Weak and Probabilistic reduction axioms and rules for $\pCBPV$} \label{fig:reduction-rules} \end{figure*} \subsection{Observational equivalence}\label{sec:obs-eq} In order to define observational equivalence, we need to represent the probability of convergence of a term to a normal form. As in~\cite{DanosEhrhard08}, we consider the reduction as a discrete time Markov chain whose states are terms and stationary states are weak normal terms. We then define a stochastic matrix $\Redmats\in[0,1]^{\pCBPV\times\pCBPV}$ (indexed by terms) as \begin{equation*} \Redmats_{M,M'}= \begin{cases} p & \text{if }M\Rel{\Redone p}M'\\ 1 & \text{if $M$ is weak-normal and $M'=M$}\\ 0 & \text{otherwise.} \end{cases} \end{equation*} Saying that $\Redmats$ is stochastic means that the coefficients of $\Redmats$ belong to $\Rseg 01$ and that, for any given term $M$, one has $\sum_{M'}\Redmats_{M,M'}=1$ (actually there are at most two terms $M'$ such that $\Redmats_{M,M'}\not=0$). For all $M,M'\in\pCBPV$, if $M'$ is weak-normal then the sequence $(\Redmats^n_{M,M'})_{n=1}^\infty$ is monotone and included in $[0,1]$, and therefore has a lub that we denote as $\Redmats^\infty_{M,M'}$ which defines a sub-stochastic matrix (taking $\Redmats^\infty_{M,M'}=0$ when $M'$ is not weak-normal). When $M'$ is weak-normal, the number $p=\Redmats^\infty_{M,M'}$ is the probability that $M$ reduces to $M'$ after a finite number of steps. Let us say when two closed terms $M_1$, $M_2$ of type $\sigma$ are \emph{observationally equivalent}: \begin{center} $M_1\Rel\Obseq M_2$, if for all closed term $C$ of type $\LIMPL{\EXCL\sigma}{\ONE}$, $ \Redmats^\infty_{\LAPP C{\STOP{M_1}},\ONELEM} =\Redmats^\infty_{\LAPP C{\STOP{M_1}},\ONELEM}$. \end{center} For simplicity we consider only closed terms $M_1$ and $M_2$. We could also define an observational equivalence on non closed terms, replacing the term $C$ with a context $C[\ ]$ which could bind free variables of the $M_i$'s, this would not change the results of the paper. \subsection{Examples}\label{subsec:exsyn} For the sake of readability, we drop the fold/unfold constructs associated with recursive types definitions; they can easily be inserted at the right places. \subsubsection*{Ever-looping program.} Given any type $\sigma$, we define $\LOOP\sigma=\FIXT x\sigma{\GO x}$ which satisfies $\TSEQ{}{\LOOP\sigma}{\sigma}$. It is clear that $\LOOP\sigma\Rel\Wred\GO{\STOP{(\LOOP\sigma)}}\Rel\Wred\LOOP\sigma$ so that we can consider $\LOOP\sigma$ as the ever-looping program of type $\sigma$. \subsubsection*{Booleans.} We define the type $\BOOL=\PLUS\ONE\ONE$, so that $\TSEQ\cP{\COIN p}\BOOL$. We define the ``true'' constant as $\True =\IN \ell \ONELEM$ and the ``false'' constant as $\False =\IN r \ONELEM$. The corresponding eliminator is defined as follows. Given terms $M$, $N_\ell$ and $N_r$ we set $\IFB M{N_\ell}{N_r}=\CASE M{x_\ell}{N_\ell}{x_r}{N_r}$ where $x_i$ is not free in $N_i$ for $i\in\{\ell,r\}$, so that \begin{center} \AxiomC{$\TSEQ\cP M\BOOL$} \AxiomC{$\TSEQ\cP{N_\ell}\sigma$} \AxiomC{$\TSEQ{\cP}{N_r}\sigma$} \TrinaryInfC{$\TSEQ{\cP}{\IFB M{N_\ell}{N_r}}\sigma$} \DisplayProof \end{center} We have the following weak and probabilistic reduction rules, derived from Figure~\ref{fig:reduction-rules}: \begin{center} \AxiomC{} \UnaryInfC{$\IFB \True{N_\ell}{N_r}\Rel\Wred{N_\ell}$} \DisplayProof\quad \AxiomC{} \UnaryInfC{$\IFB \False{N_\ell}{N_r}\Rel\Wred{N_r}$} \DisplayProof \quad \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\IFB M{N_\ell}{N_r} \Rel{\Redone p}\IFB {M'}{N_\ell}{N_r}$} \DisplayProof \end{center} \subsubsection*{Natural numbers.} We define the type $\NAT$ of unary natural numbers by $\NAT=\PLUS\ONE\NAT$ (by this we mean that $\NAT=\TREC\zeta{(\PLUS\ONE\zeta)}$). We define $\NUM 0=\IN \ell \ONELEM$ and $\NUM{n+1}=\IN r{\NUM n}$ so that we have $\TSEQ\cP{\NUM n}{\NAT}$ for each $n\in\Nat$. Then, given a term $M$, we define the term $\TSUCC M=\IN r{M}$, so that we have \begin{center} \AxiomC{$\TSEQ\cP M\NAT$} \UnaryInfC{$\TSEQ\cP{\TSUCC M}\NAT$} \DisplayProof \end{center} Last, given terms $M$, $N_\ell$ and $N_r$ and a variable $x$, we define an ``\textsf{ifz}'' conditional by $\IFV M{N_\ell}{x}{N_r}=\CASE Mz{N_\ell}x{N_r}$ where $z$ is not free in $N_\ell$, so that \begin{center} \AxiomC{$\TSEQ\cP M\NAT$} \AxiomC{$\TSEQ\cP{N_\ell}\sigma$} \AxiomC{$\TSEQ{\cP,x:\NAT}{N_r}\sigma$} \TrinaryInfC{$\TSEQ{\cP}{\IFV M{N_\ell}{x}{N_r}}\sigma$} \DisplayProof \end{center} We have the following weak and probabilistic reduction rules, derived from Figure~\ref{fig:reduction-rules}: \begin{center} \AxiomC{} \RightLabel{$i\in\{\ell, r\}$} \UnaryInfC{$\IFV{\IN iV}{M_\ell}{x}{M_r}\Rel\Wred\Subst{M_i}{V}{x}$} \DisplayProof\quad \AxiomC{$M\Rel{\Redone p} M'$} \UnaryInfC{$\IFV M{N_\ell}x{N_r} \Rel{\Redone p}\IFV {M'}{N_\ell}x{N_r}$} \DisplayProof \end{center} These conditionals will be used in the examples below. \subsubsection*{Streams.} Let $\phi$ be a positive type and $\STREAM\phi$ be the positive type defined by $\STREAM\phi=\EXCL{(\TENS\phi{\STREAM\phi})}$, that is $\STREAM\phi=\TREC\zeta{\EXCL{(\TENS\phi\zeta)}}$. We can define a term $M$ such that $\TSEQ{}{M}{\LIMPL{\STREAM\phi}{\LIMPL\NAT \phi}}$ which computes the $n$th element of a stream: \begin{align*} M=\FIXTP f{\LIMPL{\STREAM\phi}{\LIMPL\NAT \phi}}{\ABST x{\STREAM\phi}{\ABST y\NAT{}}{}}\IFV y{\PR \ell {(\GO x)}}{z}{\LAPP{\GO f}{\PR r {(\GO x)}}{\,z}} \end{align*} Let $O=\STOP{(\LOOP{\TENS\phi{\STREAM\phi}})}$, a term which represents the ``undefined stream'' (more precisely, it is a stream which is a value, but contains nothing, not to be confused with $\LOOP{\STREAM\phi}$ which has the same type but is not a value). We have $\Tseq{}{O}{\STREAM\phi}$, and observe that the reduction of $\LAPP MO$ converges (to an abstraction) and that $\LAPP M{O\Appsep\NUM 0}$ diverges. Conversely, we can define a term $N$ such that $\TSEQ {}N{\LIMPL{\EXCL{(\LIMPL\NAT\phi)}}{\STREAM\phi}}$ which turns a function into the stream of its successive applications to an integer. \begin{align*} N=\FIXTP F{\LIMPL{\EXCL{(\LIMPL\NAT\phi)}}{\STREAM\phi}} {\ABST f{\EXCL{(\LIMPL\NAT\phi)}}{}} \STOP{\PAIR{\LAPP{\GO f}{\NUM 0}}{\LAPP{\GO F} { \STOP{(\ABST x\NAT{\LAPP{\GO f}{\TSUCC x}})} } }} \end{align*} Observe that the recursive call of $F$ is encapsulated into a box, which makes the construction lazy. As a last example, consider the following term $P$ such that $\Tseq{}{P}{\LIMPL{(\TENS{\STREAM\phi}{\STREAM\phi})}{\LIMPL{(\PLUS\NAT\NAT)}{\phi}}}$ given by \begin{align*} P=\ABST{y}{\TENS{\STREAM\phi}{\STREAM\phi}} {\ABST{c}{\PLUS\NAT\NAT}{\CASE{c}{x}{\LAPP M{x\Appsep\PR \ell y}}{x}{\LAPP M{x\Appsep\PR r y}}}} \end{align*} Take $\phi=\ONE$ and consider the term $Q=\PAIR{\STOP{\PAIR{\ONELEM}{O}}}O$, then we have $\Tseq{}Q{\TENS{\STREAM\One}{\STREAM\One}}$, and observe that $\LAPP P{Q\Appsep{\IN \ell {\NUM 0}}}$ converges to $\ONELEM$ whereas $\LAPP P{Q\Appsep{\IN r {\NUM 0}}}$ diverges. These examples suggest that $\STREAM\phi$ behaves as should behave a type of streams of elements of type $\phi$. \subsubsection*{Lists.} There are various possibilities for defining a type of lists of elements of a positive type $\phi$. The simplest definition is $\lambda_0=\PLUS\ONE{(\TENS\phi{\lambda_0})}$. This corresponds to the ordinary ML type of lists. But we can also define $\lambda_1=\PLUS\ONE{\EXCL{(\TENS\phi{\lambda_1})}}$ and then we have a type of lazy lists (or terminable streams) where the tail of the list is computed only when required. Here is an example of a term $L$ such that $\Tseq{}{L}{\lambda_1}$, with $\phi=\BOOL=\PLUS\ONE\ONE$ which is a list of random length containing random booleans: \begin{align*} L=\FIXT x{\lambda_1}{\IFB{\COIN{1/4}} {\IN \ell \ONELEM}{\IN r {\STOP{(\TENS{\COIN{1/2}}{\GO x})}}}} \end{align*} Then $L$ will reduce with probability $\frac 14$ to the empty list $\IN \ell \ONELEM$, and with probability $\frac 34$ to the list $L'=\IN r{\STOP{(\TENS{\COIN{1/2}}{L})}}$ (up to the equivalence of $\GO{\STOP L}$ with $L$) which is a value. One can access this value by evaluating the term $\CASE{L'}{z}{\_}{y}{\GO y}$ where $\_$ is any term of type $\TENS\BOOL{\lambda_1}$ (we know that this branch will not be evaluated) and this term reduces to $\PAIR{\True}{L}$ or $\PAIR{\False}{L}$ with probability $\frac 12$. In turn each of these terms $\PAIR bL$ will reduce to $\PAIR b{\IN \ell \ONELEM}$ with probability $\frac14$ and to $\PAIR b{L'}$ with probability $\frac 34$. We can iterate this process, defining a term $R$ of type $\LIMPL{\lambda_1}{\lambda_0}$ which evaluates completely a terminable stream to a list: \begin{align*} R=\FIXT{f}{(\LIMPL{\lambda_1}{\lambda_0})} {\ABST x{\lambda_1}{\CASE{x}{z}{\IN \ell \ONELEM}{z}{\LAPP{\ABST y{ \TENS{\Bool}{\lambda_1}}{\PAIR{\PR\ell y}{\LAPP{\GO f}{\PR r y}}}}{\GO z}}}}\,. \end{align*} Then $\LAPP RL$, which is a closed term of type $\lambda_0$, terminates with probability $1$. The expectation of the length of this ``random list'' is $\sum_{n=0}^\infty n(\frac 34)^n=12$. We could also consider $\lambda_2=\PLUS\ONE{(\TENS{\EXCL \sigma}{\lambda_2})}$ which allows to manipulate lists of objects of type $\sigma$ (which can be a general type) without accessing their elements. \subsubsection*{Probabilistic tests.} If $\TSEQ{\cP}{M_i}{\sigma}$ for $i=1,2$, we set $\DICE{p}{M_1}{M_2}=\IFB{\COIN p}{M_1}{M_2}$ and this term satisfies $\Tseq{\cP}{\DICE{p}{M_1}{M_2}}{\sigma}$. If $M_i$ reduces to a value $V_i$ with probability $q_i$, then $\DICE{p}{M_1}{M_2}$ reduces to $V_1$ with probability $p\,q_1$ and to $V_2$ with probability $(1-p)q_2$. Let $n\in\Nat$ and let $\Vect p=(\List p0n)$ be such that $p_i\in[0,1]\cap\Rational$ and $p_0+\cdots+p_n\leq 1$. Then one defines a closed term $\Ran{\Vect p}$, such that $\Tseq{}{\Ran{\Vect p}}{\Tnat}$, which reduces to $\Num i$ with probability $p_i$ for each $i\in\{0,\dots,n\}$. The definition is by induction on $n$. \begin{equation*} \Ran{\Vect p}= \begin{cases} \Num 0 & \hspace{-10em}\text{if }p_0=1\text{ whatever be the value of }n\\ \IFB{\COIN{p_0}}{\Num 0}{\LOOP\Tnat}& \hspace{-10em}\text{if }n=0\\ \IFB{\COIN{p_0}} {\Num 0} {\TSUCC{\Ran{\frac{p_1}{1-p_0},\dots,\frac{p_n}{1-p_0}}}} &\text{otherwise} \end{cases} \end{equation*} \medskip As an example of use of the test to zero conditional, we define, by induction on $k$, a family of terms $\Probe_k$ such that $\Tseq{}{\Probe_k}{\LIMPL\Tnat\ONE}$ and that tests the equality to $k$: \begin{align*} \Probe_0 = \Abst x\Tnat{\IFV{x}{\ONELEM}{z}{\LOOP\ONE}}\qquad \Probe_{k+1} = \Abst x\Tnat{\IFV{x}{\LOOP\ONE}{z}{\LAPP{\Probe_k}z}} \end{align*} For $M$ such that $\Tseq{}{M}{\Tnat}$, the term $\LAPP{\Probe_k}M$ reduces to $\ONELEM$ with a probability which is equal to the probability of $M$ to reduce to $\Num k$. \subsubsection{Notation.}\label{subsec:not-terms} Now, we introduce terms that will be used in the definition of testing terms in the proof of Full Abstraction in Section~\ref{sec:FA}. First, we define $\Pprod_k$ such that $\Tseq{}{\Pprod_k}{\LIMPL{\ONE}{\cdots\LIMPL{}{\LIMPL\ONE{\LIMPL{\phi}\phi}}}}$ (with $k$ occurrences of $\ONE$): \begin{align*} \Pprod_0 = \ABST y{\phi} y\qquad \Pprod_{k+1} = \ABST x\ONE{\Pprod_k}\,. \end{align*} Given for each $i\in\{0,\dots,k\}$, $M_i$ such that $\Tseq{\cP}{M_i}\ONE$ and $\Tseq{\cP}{N}{\phi}$, the term $\LAPP{\LAPP{\LAPP{\Pprod_{k+1}}{M_0}\cdots}{M_k}}N$ reduces to a value $V$ with probability $p_0\,\cdots\,p_k\,q$ where $p_i$ is the probability of $M_i$ to reduce to $\ONELEM$ and $q$ is the probability of $N$ to reduce to $V$. We use the notations: \begin{equation*} M_0\cdot N=\LAPP{\LAPP{\Pprod_1}{M_0}}N\qquad M_0\AND\cdots\AND M_{k-1} =\left\{ \begin{array}{l} \ONELEM\text{ if $k=0$}\\ \LAPP{\LAPP{\Pprod_{k}}{M_0}{\cdots} }M_{k-1}\text{ otherwise}, \end{array}\right. \end{equation*} so that $\TSEQ{\cP}{M_0\cdot N}{\phi}$ and the probability that $M_0\cdot N$ reduces to $V$ is $p_0\, q$ and $\TSEQ{\cP}{M_0\AND\cdots\AND M_{k-1}}{\ONE}$ and $M_0\AND\cdots\AND M_{k-1}$ reduces to $\ONELEM$ with probability $p_0\,\cdots\,p_{k-1}$. \medskip Given a general type $\sigma$ and terms $\List M0{k-1}$ such that, for any $i\in\{0,\dots,k-1\}$, $\Tseq{}{M_i}\sigma$, we define close terms $\Pchoose^\sigma_i(\List M 0{k-1})$ for $i\in\{0,\dots,n-1\}$ such that $\Tseq{}{\Pchoose^\sigma_i(\List M0{k-1})}{\LIMPL\Tnat{{\sigma}}}$ \begin{align*} \Pchoose^\sigma_0(\List M 0{k-1}) &= \Abst z\Tnat{\LOOP\sigma}\\ \Pchoose^\sigma_{i+1}(\List M 0{k-1}) &= \Abstpref z\Tnat \IFV z{M_0}{y}{\LAPP{\Pchoose^\sigma_{i}(\List M 1{k-1})}{y}} \text{ if }i\le k-1 \end{align*} Given a term $P$ such that $\Tseq{\cP}{P}{\Tnat}$ and $p_i$ the probability of $P$ to reduce to $\Num i$ for any $i$, the first steps of the reduction are probabilistic: \begin{align*} \forall i\in\{0,\dots, k\},\ {\LAPP{\Pchoose^\sigma_{k+1}(\List M0k)}{P}}\Rel{\Redonetr {p_i}} {\LAPP{\Pchoose^\phi_{k+1}(\List M0k)}{\Num i}} \end{align*} the next steps of the reduction are deterministic: \begin{align*} \LAPP{\Pchoose^\phi_{k+1}(\List M0k)}{\Num i} \Rel{\Wredtr} M_i \end{align*} \medskip As we will see more precisely in Paragraph~\ref{subsec:sem-type-term}, a term of type $\Tnat$ can be seen as a sub-probability distribution over $\Nat$. Given integers $0\le l\le r$, we define by induction the term $\Pext lr$ of type $\Limpl\Tnat\Tnat$: \begin{eqnarray*} \Pext 00&=&\ABST z\Tnat{\IFV z{\Num 0}x{\LOOP\Tnat}}\\ \forall r>0,\ \Pext 0r&=& \Pchoose^\Tnat_{r+1}{\TUPLE{\Num 0,\dots,\Num r}}\\ \Pext {l+1}{r+1}&=&\ABST{z}{\Tnat}{\IFV z{\LOOP\Tnat}{x}{\TSUCC{\LAPP{\Pext lr}x}}} \end{eqnarray*} such that if $\Tseq{}{P}{\Tnat}$, then $\LAPP{\Pext lr}P$ extracts the sub-probability distribution with support $\subseteq\{l,\dots,r\}$. Indeed, for any $i\in\{l,\dots,r\}$ $\LAPP{\Pext lr}P$ reduces to $\Num i$ with probability $p_i$ where $p_i$ is the probability of $P$ to reduce to $\Num i$. We also introduce, for $\Vect n=(\List n0k)$ a sequence of $k+1$ natural numbers, a term $\Pwin{i}{\Vect n}$ of type $\Limpl\Tnat\Tnat$ which extracts the sub-probability distribution whose support is in the $i^{th}$ window of length $n_i$ for $0\le i< k$: \begin{eqnarray*} \Pwin 0{\Vect n}&=& \Pext 0{n_0-1}\\ \Pwin {i+1}{\Vect n}&=&\Pext{n_0+\cdots+n_{i}}{n_0+\cdots+n_i+ n_{i+1}-1}. \end{eqnarray*} \subsection{On products and recursive definitions of general types}\label{sec:PCF-products} This section has nothing to do with probabilities, so we consider the deterministic language $\CBPV$, which is just $\pCBPV$ without the $\COIN p$ construct. Our $\CBPV$ general types, which are similar to Levy's CBPV computation types~\cite{LevyP04} or to the SFPL general types~\cite{MarzRohrStreicher99}, have $\LIMPL{}{}$ as only type constructor. This may seem weird when one compares our language with CBPV where products and recursive definitions are possible on computation types and are used for encoding CBN functional languages such as $\PCF$ extended with products. For keeping our presentation reasonably short, we will not consider the corresponding extensions of $\CBPV$ in this paper. Instead, we will shortly argue that such a $\PCF$ language with products can be easily encoded in our $\CBPV$. Another reason for not introducing a product $\IWith$ at the level of general types is that this would introduce a useless redundancy in the language since a type like $\LIMPL{\TENS{\EXCL{\sigma_1}}{\EXCL{\sigma_2}}}\tau$ would then be represented equivalently as $\LIMPL{\EXCL{(\sigma_1\IWith\sigma_2)}}\tau$. Concerning recursive type definitions, it is true that adding them at the level of general types would allow to define interesting types such as $\TREC\zeta{(\LIMPL{\EXCL\zeta}{\zeta})}$, yielding a straightforward encoding of the pure lambda-calculus in our language. This goal can nevertheless be reached (admittedly in a slightly less natural way) by using the positive recursive type definition $\phi=\TREC\zeta{\EXCL{(\LIMPL{\zeta}{\zeta})}}$. A pure term $t$ will be translated into a $\CBPV$ term $\Cterm t$ such that $\Tseq{x_1:{\phi_1},\dots,x_n:\phi}{\Cterm t}{\LIMPL\phi\phi}$, where the list $\List x1n$ contains all the free variables of $t$. This translation is defined inductively as follows: $\Cterm x=\GO x$, $\Cterm{(\App st)}=\GO{(\LAPP{\Cterm s}{\STOP{(\Cterm t)}})}$ and $\Cterm{(\Abs xs)}=\Abs x{\STOP{(\Cterm s)}}$. The examples we provide in Section~\ref{subsec:exsyn} also show that our recursive definitions restricted to positive types allow to introduce a lot of useful data types (integers, lists, trees, streams etc). So we do not see any real motivations for adding recursive general type definitions (and their addition would make the proof of adequacy in Section~\ref{sec:adequacy} even more complicated). Coming back to the redundancy of products, consider the following grammar of types \begin{align*} A,B,\dots\Bnfeq \Cnat\Bnfor \Impl AB\Bnfor\Cprod AB \end{align*} and the following language of terms \begin{align*} s,t,u,\dots\Bnfeq x\Bnfor\Num n\Bnfor\Csuc s\Bnfor\Cpred s \Bnfor\Cifz stu \Bnfor \Abst xAs\Bnfor\App st\Bnfor\Cpair st \Bnfor \Cproj \ell s\Bnfor \Cproj r s\Bnfor\Cfix xAs \end{align*} We call this languages $\PCF$ as it is a straightforward extension of the original $\PCF$ of~\cite{Plotkin77}. The typing rules are described in Figure~\ref{fig:typing-PCF}. A typing context is a sequence $\Gamma=(x_1:A_1,\dots,x_n:A_n)$ where the variables are pairwise distinct. We explain now how we interpret this simple programming language in $\CBPV$. \subsubsection*{Translating $\PCF$ types.} With any type $A$, we associate a \emph{finite sequence} of general types $\Ctype A=(\Ctype A_1,\dots,\Ctype A_n)$ whose length $n=\Clen A$ is given by: $\Clen\Cnat=1$, $\Clen{\Impl AB}=\Clen B$ and $\Clen{\Cprod AB}=\Clen A+\Clen B$. Given a sequence $\Vect\sigma=(\List\sigma 1n)$ of general types we define $\EXCL{\Vect\sigma}$ by induction: $\EXCL{()}=\ONE$ and $\EXCL{(\sigma,\Vect\sigma)}=\TENS{\EXCL\sigma}{\EXCL{\Vect\sigma}}$. Given a positive type $\phi$ and a sequence $\Vect\sigma=(\List\sigma 1n)$ of general types, we define $\LIMPL\phi{\Vect\sigma}=(\LIMPL\phi{\sigma_1},\dots,\LIMPL\phi{\sigma_n})$. Using these notations, we can now define $\Ctype A$ as follows: \begin{itemize} \item $\Ctype\Cnat=(\NAT)$ (a one element sequence) where $\NAT$ is the type of integers introduced in Section~\ref{subsec:exsyn}, $\NAT=\TREC\zeta{(\PLUS\ONE\zeta)}$, \item $\Ctype{(\Impl AB)}=\LIMPL{\EXCL{\Ctype A}}{\Ctype B}$, \item $\Ctype{(\Cprod AB)}=\Ccat{\Ctype A}{\Ctype B}$ (list concatenation). \end{itemize} \subsubsection*{Translating $\PCF$ terms.} Let now $s$ be a $\PCF$ term with typing judgment\footnote{So our translation depends on the typing judgment and not only on the term; this is fairly standard and can be avoided by considering typed free variables.} $\Tseq\Gamma sA$. Let $n=\Clen A$, we define a sequence $\Cterm s$ of length $n$ such that $\Tseq{\EXCL\Gamma}{\Cterm s_i}{\Ctype A_i}$ (for $i=1,\dots,n$) as follows (if $\Gamma=(x_1:C_1,\dots,x_k:C_k)$, then $\EXCL\Gamma=(x_1:\EXCL{\Ctype{C_1}},\dots,x_k:\EXCL{\Ctype{C_k}})$). If $s=x_j$, so that $A=C_j$ (for some $j\in\{1,\dots,k\}$), let $(\List\sigma 1n)=\Ctype A$. Since $\EXCL{\Ctype A}=\TENS{\EXCL{\sigma_1}}{(\TENS{\EXCL{\sigma_2}} {\cdots(\TENS{\EXCL{\sigma_{n-1}}}{(\TENS{\EXCL{\sigma_n}}\ONE)})\cdots})}$ we can set $\Cterm x=(\Cterm x_1,\dots,\Cterm x_n)$ where $\Cterm x_i=\GO{\PR\ell {\PR r {\PR r {\cdots\PR r {x}}}}}$ (with $i-1$ occurrences of $\PR r {}$). We set $\Cterm{\Num n}=\IN r {\cdots\IN r{\IN \ell \ONELEM}}$ ($n$ occurrences of $\IN r{}$), $\Cterm{(\Csuc s)}=\IN r{\Cterm s}$, $\Cterm{(\Cpred s)}=\CASE{\Cterm s}{x}{\IN \ell \ONELEM}{x}{x}$. Assume that $s=\Cifz tuv$ with $\Tseq\Gamma t\NAT$, $\Tseq\Gamma uA$ and $\Tseq\Gamma vA$ for some $\PCF$ type $A$. Let $l=\Clen A$. By inductive hypothesis we have $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm t}{\NAT}$, $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm u_i}{\Ctype A_i}$ and $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm v_i}{\Ctype A_i}$ for $i=1,\dots,l$. So we set $\Cterm s=(\CASE{\Cterm t}{z}{\Cterm u_i}{z}{\Cterm v_i})_{i=1}^l$ where $z$ is a variable which does not occur free in $u$ or $v$. Assume now that $\Tseq{\Gamma,x:A}{s}{B}$, we set $\Cterm{(\Abst xAs)}=(\ABST x{\EXCL{\Ctype A}}{\Cterm s_i})_{i=1}^{\Clen B}$. Assume that $\Tseq\Gamma s{\Impl AB}$ and $\Tseq\Gamma tA$. Then, setting $n=\Clen B$, we have $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm s_i}{\LIMPL{\EXCL{\Ctype A}}{\Ctype B_i}}$ for $i=1,\dots,n$ and $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm t_j}{\Ctype A_j}$ for $j=1,\dots,m$ where $m=\Clen A$. Then, setting \[ N=\PAIR{\STOP{(\Cterm{t_1})}}{\PAIR{\cdots}{\PAIR{\STOP{(\Cterm{t_{m-1}})}} {\PAIR{\STOP{(\Cterm{t_{m}})}}\ONELEM}\cdots}} \] we have $\Tseq{\EXCL{\Ctype\Gamma}}{N}{\EXCL{\Ctype A}}$ and we set $\Cterm{(\App st)}=(\LAPP{\Cterm s_i}N)_{i=1}^n$. Assume that $s=\Cpair{s_1}{s_2}$ with $\Tseq\Gamma{s_i}{A_i}$ for $i=1,2$ and $\Tseq\Gamma s{\Cprod{A_1}{A_2}}$. Then we set $\Cterm s=\Ccat{\Cterm{s_1}}{\Cterm{s_2}}$ (list concatenation). Assume that $\Tseq\Gamma s{\Cprod{A_1}{A_2}}$, with $n_i=\Clen{A_i}$ for $i=1,2$. Then we set $\Cterm{(\Cproj \ell s)}=(\Cterm s_1,\dots,\Cterm s_{n_1})$ and $\Cterm{(\Cproj r s)}=(\Cterm s_{n_1+1},\dots,\Cterm s_{n_1+n_2})$. Last assume that $s=\Cfix xAt$ with $\Tseq{\Gamma,x:A}{t}A$ so that, setting $n=\Clen A$, we have $\Tseq{\EXCL{\Ctype\Gamma},x:\EXCL{\Ctype A}}{\Cterm t_i}{\Ctype A_i}$ for $i=1,\dots,n$. Let $\List x1n$ be pairwise distinct fresh variables, and set \[ M_i=\Subst{\Cterm t_i}{\PAIR{{x_1}}{\cdots\PAIR{{x_{n-1}}}{{\PAIR{x_{n}}{\ONELEM}}}\cdots}}{x} \] for $i=1,\dots,n$. We have $\Tseq{\EXCL{\Ctype\Gamma},x_1:\EXCL{\Ctype A_1},\dots,x_n:\EXCL{\Ctype A_n}}{M_i}{\Ctype A_i}$. We are in position of applying the usual trick for encoding mutual recursive definitions using fixpoints operators. For the sake of readability, assume that $n=2$, so we have $\Tseq{\EXCL{\Ctype\Gamma},x_1:\EXCL{\Ctype A_1},x_2:\EXCL{\Ctype A_2}}{M_i}{\Ctype A_i}$ for $i=1,2$. Let $N_1=\FIXT{x_1}{\Ctype{A}_1}{M_1}$ so that $\Tseq{\EXCL{\Ctype\Gamma},x_2:\EXCL{\Ctype A_2}}{N_1}{\Ctype A_1}$. Then we have $\Tseq{\EXCL{\Ctype\Gamma,x_2:\EXCL{\Ctype A_2}}}{\Subst{M_2}{\STOP{N_1}}{x_1}}{\Ctype A_2}$. Therefore we can set $\Cterm s_2=\FIXT{x_2}{\Ctype A_2}{\Subst{M_2}{\STOP{N_1}}{x_1}}$ and we have $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm s_2}{\Ctype A_2}$. Finally we set $\Cterm s_1=\Subst{N_1}{\STOP{(\Cterm s_2)}}{x_2}$ with $\Tseq{\EXCL{\Ctype\Gamma}}{\Cterm s_1}{\Ctype A_1}$. We should check now that this translation is compatible with the operational semantics of our extended $\PCF$ language. A simple way to do so would be to choose a simple model of linear logic (for instance, the relational model) and to prove that the semantics of a $\PCF$ term is equal to the semantics of its translation in $\pCBPV$ stripped from its probabilistic construct $\COIN p$ (interpreting tuples of types using the ``additive'' cartesian product $\IWith$). This is a long and boring exercise. \begin{figure*} \centering \begin{center} \AxiomC{} \UnaryInfC{$\Tseq{\Gamma,x:A}xA$} \DisplayProof \quad \AxiomC{} \UnaryInfC{$\Tseq\Gamma{\Num n}\Cnat$} \DisplayProof \quad \AxiomC{$\Tseq\Gamma s\Cnat$} \UnaryInfC{$\Tseq\Gamma{\Csuc s}\Cnat$} \DisplayProof \quad \AxiomC{$\Tseq\Gamma s\Cnat$} \UnaryInfC{$\Tseq\Gamma{\Cpred s}\Cnat$} \DisplayProof \end{center} \begin{center} \AxiomC{$\Tseq\Gamma s\Cnat$} \AxiomC{$\Tseq\Gamma tA$} \AxiomC{$\Tseq\Gamma uA$} \TrinaryInfC{$\Tseq\Gamma{\Cifz stu}A$} \DisplayProof \quad \AxiomC{$\Tseq{\Gamma,x:A}sB$} \UnaryInfC{$\Tseq\Gamma{\Abst xAs}{\Impl AB}$} \DisplayProof \end{center} \begin{center} \AxiomC{$\Tseq\Gamma s{\Impl AB}$} \AxiomC{$\Tseq\Gamma tA$} \BinaryInfC{$\Tseq\Gamma{\App st}{B}$} \DisplayProof \quad \AxiomC{$\Tseq\Gamma{s_\ell}{A_\ell}$} \AxiomC{$\Tseq\Gamma{s_r}{A_r}$} \BinaryInfC{$\Tseq\Gamma{\Cpair{s_\ell}{s_r}}{\Cprod{A_\ell}{A_r}}$} \DisplayProof \quad \AxiomC{$\Tseq\Gamma s{\Cprod{A_\ell}{A_r}}$} \RightLabel{$i\in\{\ell,r\}$} \UnaryInfC{$\Tseq\Gamma{\Cproj is}{A_i}$} \DisplayProof \end{center} \begin{center} \AxiomC{$\Tseq{\Gamma,x:A}{M}{A}$} \UnaryInfC{$\Tseq\Gamma{\Cfix xAM}{A}$} \DisplayProof \end{center} \caption{Typing rules for a simple call-by-name language with products, PCF} \label{fig:typing-PCF} \end{figure*} \section{Probabilistic Coherent Spaces}\label{sec:PCS} \subsection{Semantics of LL, in a nutshell} \label{sec:LL-semantics-short} \label{sec:LL-based-models} The kind of denotational models we are interested in, in this paper, are those induced by a model of LL, as explained in~\cite{Ehrhard16a}. We remind the basic definitions and notations, referring to that paper for more details. \subsubsection{Models of Linear Logic.}\label{sec:LL-models} A model of LL consists of the following data. A symmetric monoidal closed category $(\cL,\ITens,\One,\Leftu,\Rightu,\Assoc,\Sym)$ where we use simple juxtaposition $g\Compl f$ to denote composition of morphisms $f\in\cL(X,Y)$ and $g\in\cL(Y,Z)$. We use $\Limpl XY$ for the object of linear morphisms from $X$ to $Y$, $\Evlin\in\cL(\Tens{(\Limpl XY)}{X},Y)$ for the evaluation morphism and $\Curlin\in \cL(\Tens ZX,Y)\to\cL(Z,\Limpl XY)$ for the linear curryfication map. For convenience, and because it is the case in the concrete models we consider (such as Scott Semantics~\cite{Ehrhard16a} or Probabilistic Coherent Spaces here), we assume this SMCC to be a $*$-autonomous category with dualizing object $\Bot$. We use $\Orth X$ for the object $\Limpl X\Bot$ of $\cL$ (the dual, or linear negation, of $X$). The category $\cL$ is cartesian with terminal object $\top$, product $\IWith$, projections $\Proj i$. By $*$-autonomy $\cL$ is co-cartesian with initial object $0$, coproduct $\IPlus$ and injections $\Inj i$. By monoidal closeness of $\cL$, the tensor product $\ITens$ distributes over the coproduct $\IPlus$. We are given a comonad $\Excl\_:\cL\to\cL$ with co-unit $\Der X\in\cL(\Excl X,X)$ (\emph{dereliction}) and co-multiplication $\Digg X\in\cL(\Excl X,\Excl{\Excl X})$ (\emph{digging}) together with a strong symmetric monoidal structure (Seely isos $\Seelyz$ and $\Seelyt$) for the functor $\Excl\_$, from the symmetric monoidal category $(\cL,\IWith)$ to the symmetric monoidal category $(\cL,\ITens)$ satisfying an additional coherence condition wrt.~$\Digg{}$. We use $\Int\_$ for the ``De Morgan dual'' of $\Excl\_$: $\Int X=\Orthp{\Exclp{\Orth X}}$ and similarly for morphisms. It is a monad on $\cL$. \subsubsection{The Eilenberg-Moore category.}\label{sec:EM-general} It is then standard to define the category $\EM\cL$ of $\IExcl$-coalgebras. An object of this category is a pair $P=(\Coalgc P,\Coalgm P)$ where $\Coalgc P\in\Obj\cL$ and $\Coalgm P\in\cL(\Coalgc P,\Excl{\Coalgc P})$ is such that $\Der{\Coalgc P}\Compl\Coalgm P=\Id$ and $\Digg{\Coalgc P}\Compl\Coalgm P=\Excl{\Coalgm P}\Compl\Coalgm P$. Then $f\in\EM\cL(P,Q)$ iff $f\in\cL(\Coalgc P,\Coalgc Q)$ such that $\Coalgm Q\Compl f=\Excl f\Compl\Coalgm P$. The functor $\Excl\_$ can be seen as a functor from $\cL$ to $\EM\cL$ mapping $X$ to $(\Excl X,\Digg X)$ and $f\in\cL(X,Y)$ to $\Excl f$. It is right adjoint to the forgetful functor $\Forgca:\EM\cL\to\cL$. Given $f\in\cL(\Coalgc P,X)$, we use $\Prom f\in\EM\cL(P,\Excl X)$ for the morphism associated with $f$ by this adjunction, one has $\Prom f=\Excl f\Compl\Coalgm P$. If $g\in\EM\cL(Q,P)$, we have $\Prom f\Compl g=\Promp{f\Compl g}$. Then $\EM\cL$ is cartesian (with product of shape $\Tens PQ=(\Tens{\Coalgc P}{\Coalgc Q},\Coalgm{\Tens PQ})$ and terminal object $(\One,\Coalgm\One)$, still denoted as $\One$). This category is also co-cartesian with coproduct of shape $\Plus PQ=(\Plus{\Coalgc P}{\Coalgc Q},\Coalgm{\Plus PQ})$ and initial object $(0,\Coalgm0)$ still denoted as $0$. The complete definitions can be found in~\cite{Ehrhard16a}. We use $\Contr P\in\EM\cL(P,\Tens PP)$ (\emph{contraction}) for the diagonal and $\Weak P\in\EM\cL(P,\One)$ (\emph{weakening}) for the unique morphism to the terminal object. We also consider occasionally the \emph{Kleisli category}\footnote{It is the full subcategory of $\EM\cL$ of free coalgebras, see any introductory text on monads and co-monads.} $\Kl\cL$ of the comonad $\oc$: its objects are those of $\cL$ and $\Kl\cL(X,Y)=\cL(\Excl X,Y)$. The identity at $X$ in this category is $\Der X$ and composition of $f\in\Kl\cL(X,Y)$ and $g\in\Kl\cL(Y,Z)$ is defined as \begin{align*} g\Comp f=g\Compl\Excl f\Compl\Digg X\,. \end{align*} This category is cartesian closed but this fact will not play an essential role in this work. \subsubsection{Fixpoints.}\label{parag:sem-fix-points} For any object $X$, we assume to be given $\Sfix_X\in\cL(\Exclp{\Limpl{\Excl X}X},X)$, a morphism such that\footnote{It might seem natural to require the stronger uniformity conditions of \emph{Conway operator}~\cite{PlotkinSimpson00}. This does not seem to be necessary as far as soundness of our semantics is concerned even if the fixpoint operators arising in concrete models satisfy these further properties.} $\Evlin\Compl(\Tens{\Der{\Limpl{\Excl X}{X}}}{\Prom{\Sfix_X}}) \Comp\Contr{\Excl{(\Limpl{\Excl X}{X})}}=\Sfix_X$ which will allow to interpret term fixpoints. In order to interpret fixpoints of types, we assume that the category $\cL$ is equipped with a notion of embedding-retraction pairs, following a standard approach. We use $\Embr\cL$ for the corresponding category. It is equipped with a functor $\Funofemb:\Embr\cL\to\Op\cL\times\cL$ such that $\Funofemb(X)=(X,X)$ and for which we use the notation $(\Ret\phi,\Emb\phi)=\Funofemb(\phi)$ and assume that $\Ret\phi\Compl\Emb\phi=\Id_X$. We assume furthermore that $\Embr\cL$ has all countable directed colimits and that the functor $\Embf=\Proj2\Compl\Funofemb:\Embr\cL\to\cL$ is continuous. We also assume that all the basic operations on objects ($\ITens$, $\IPlus$, $\Orth{(\_)}$ and $\Excl\_$) are continuous functors from $\Embr\cL$ to itself\footnote{This is a rough statement; one has to say for instance that if $\phi_i\in\Embr\cL(X_i,Y_i)$ for $i=1,2$ then $\Ret{(\Tens{\phi_1}{\phi_2})}=\Tens{\Ret{\phi_1}}{\Ret{\phi_2}}$ etc. The details can be found in~\cite{Ehrhard16a}.}. Then it is easy to carry this notion of embedding-retraction pairs to $\EM\cL$, defining a category $\Embr{\EM\cL}$, to show that this category has all countable directed colimits and that the functors $\ITens$ and $\IPlus$ are continuous on this category: $\Embr{\EM\cL}(P,Q)$ is the set of all $\phi\in\Embr\cL(\Coalgc P,\Coalgc Q)$ such that $\Emb\phi\in\EM\cL(P,Q)$. One checks also that $\Excl{}$ defines a continuous functor from $\Embr\cL$ to $\Embr{\EM\cL}$. This allows to interpret recursive types, more details can be found in~\cite{Ehrhard16a}. \subsubsection{Interpreting types.} \label{sec:cbpv-interpretation} Using straightforwardly the object $\One$ and the operations $\ITens$, $\IPlus$, $\Excl{}$ and $\Limpl{}{}$ of the model $\cL$ as well as the completeness and continuity properties explained in Section~\ref{parag:sem-fix-points}, we associate with any positive type $\phi$ and any repetition-free list $\Vect\zeta=(\List\zeta 1n)$ of type variables containing all free variables of $\phi$ a continuous functor $\Tsemca\phi_{\Vect\zeta}:(\Embr{\EM\cL})^n\to\Embr{\EM\cL}$ and with any general type $\sigma$ and any list $\Vect\zeta=(\List\zeta 1n)$ of pairwise distinct type variables containing all free variables of $\sigma$ we associate a continuous functor $\Tsem\sigma_{\Vect\zeta}:(\Embr{\EM\cL})^n\to\Embr{\cL}$. When we write $\Tsem\sigma$ or $\Tsemca\phi$ (without subscript), we assume implicitly that the types $\sigma$ and $\phi$ have no free type variables. Then $\Tsem\sigma$ is an object of $\cL$ and $\Tsemca\phi$ is an object of $\EM\cL$. We have $\Tsem\phi=\Coalgc{\Tsemca\phi}$ that is, considered as a generalized type, the semantics of a positive type $\phi$ is the carrier of the coalgebra $\Tsemca\phi$. Given a typing context $\cP=(x_1:\phi_1,\dots,x_k:\phi_k)$, we define $\Tsem\cP=\Tsemca{\phi_1}\ITens\cdots\ITens\Tsemca{\phi_k}\in\EM\cL$. In the model or probabilistic coherence spaces considered in this paper, we define $\Embr\cL$ in such a way that the only isos are the identity maps. This implies that the types $\TREC\zeta\phi$ and $\Subst\phi{(\TREC\zeta\phi)}\zeta$ are interpreted as \emph{the same object} (or functor). Such definitions of $\Embr\cL$ are possible in many other models (relational, coherence spaces, hypercoherences etc). We postpone the description of term interpretation because this will require constructions specific to our probabilistic semantics, in addition to the generic categorical ingredients introduced so far. \subsection{The model of probabilistic coherence spaces}\label{subsec:model-pcoh} Given a countable set $I$ and $u,u'\in\Realpto I$, we set $\Eval{u}{u'}=\sum_{i\in I}u_iu'_i$. Given $\cF\subseteq\Realpto I$, we set $\Orth\cF=\{u'\in\Realpto I\mid\forall u\in\cF\ \Eval{u}{u'}\leq 1\}$. A \emph{probabilistic coherence space} (PCS) is a pair $X=(\Web X,\Pcoh X)$ where $\Web X$ is a countable set and $\Pcoh X\subseteq\Realpto{\Web X}$ satisfies \begin{itemize} \item $\Biorth{\Pcoh X}=\Pcoh X$ (equivalently, $\Biorth{\Pcoh X}\subseteq\Pcoh X$), \item for each $a\in\Web X$ there exists $u\in\Pcoh X$ such that $u_a>0$, \item for each $a\in\Web X$ there exists $A>0$ such that $\forall u\in\Pcoh X\ u_a\leq A$. \end{itemize} If only the first of these conditions holds, we say that $X$ is a \emph{pre-probabilistic coherence space} (pre-PCS). The purpose of the second and third conditions is to prevent infinite coefficients to appear in the semantics. This property in turn will be essential for guaranteeing the morphisms interpreting proofs to be analytic functions, which will be the key property to prove full abstraction. So these conditions, though aesthetic at first sight, are important for our ultimate goal. \begin{lemma}\label{lemma:pcoh-charact} Let $X$ be a pre-PCS. The following conditions are equivalent: \begin{itemize} \item $X$ is a PCS, \item $\forall a\in\Web X\,\exists u\in\Pcoh X\,\exists u'\in\Orth{\Pcoh X}\ u_a>0\text{ and }u'_a>0$, \item $\forall a\in\Web X\,\exists A>0\,\forall u\in\Pcoh X\,\forall u'\in\Orth{\Pcoh X}\ u_a\leq A\text{ and }u'_a\leq A$. \end{itemize} \end{lemma} The proof is straightforward. We equip $\Pcoh X$ with the most obvious partial order relation: $u\leq v$ if $\forall a\in\Web X\ u_a\leq v_a$ (using the usual order relation on $\Real$). \begin{theorem}\label{th:Pcoh-prop} $\Pcoh X$ is an $\omega$-continuous domain. Given $u,v\in\Pcoh X$ and $\alpha,\beta\in\Realp$ such that $\alpha+\beta\leq 1$, one has $\alpha u+\beta v\in\Pcoh X$. \end{theorem} This is an easy consequence of the hypothesis $\Biorth{\Pcoh X}\subseteq\Pcoh X$. See~\cite{DanosEhrhard08} for details; from this result, we will only use the closure properties: $\Pcoh X$ is closed under sub-probabilistic linear combinations and under lubs of monotonic sequences. Though the $\omega$-continuity property (and the associated way-below relation) does not play any technical role, it is an intuitively satisfactory fact\footnote{The $\omega$-continuity is similar to separability for topological vector spaces.} which means that the ``size'' of our domains remains bounded. \subsubsection{Morphisms of PCSs}\label{subsec:morphisms} Let $X$ and $Y$ be PCSs. Let $t\in(\Realp)^{\Web X\times\Web Y}$ (to be understood as a matrix). Given $u\in\Pcoh X$, we define $\Matapp tu\in\Realpc^{\Web Y}$ by $(\Matapp tu)_b=\sum_{a\in\Web X}t_{a,b}u_a$ (application of the matrix $t$ to the vector $u$)\footnote{This is an unordered sum, which is infinite in general. It makes sense because all its terms are $\geq 0$.}. We say that $t$ is a \emph{(linear) morphism} from $X$ to $Y$ if $\forall u\in\Pcoh X\ \Matapp tu\in\Pcoh Y$, that is \begin{align*} \forall u\in\Pcoh X\,\forall{v'}\in\Orth{\Pcoh Y}\quad\sum_{(a,b)\in\Web X\times\Web Y}t_{a,b}u_av'_b\leq 1\,. \end{align*} The diagonal matrix $\Id\in(\Realp)^{\Web X\times\Web X}$ given by $\Id_{a,b}=1$ if $a=b$ and $\Id_{a,b}=0$ otherwise is a morphism. In that way we have defined a category $\PCOH$ whose objects are the PCSs and whose morphisms have just been defined. Composition of morphisms is defined as matrix multiplication: let $s\in\PCOH(X,Y)$ and $t\in\PCOH(Y,Z)$, we define $\Matapp ts\in(\Realp)^{\Web X\times\Web Z}$ by \begin{align*} (\Matapp ts)_{a,c}=\sum_{b\in\Web Y}s_{a,b}t_{b,c} \end{align*} and a simple computation shows that $\Matapp ts\in\PCOH(X,Z)$. More precisely, we use the fact that, given $u\in\Pcoh X$, one has $\Matapp{(\Matapp ts)}{u}=\Matapp t{(\Matapp su)}$. Associativity of composition holds because matrix multiplication is associative. $\Id_X$ is the identity morphism at $X$. Given $u\in\Pcoh X$, we define $\Norm u_X=\sup\{\Eval u{u'}\mid u'\in\Pcoh{\Orth X}\}$. By definition, we have $\Norm u_X\in[0,1]$. \subsubsection{Multiplicative constructs}\label{sec:multiplicatives} One sets $\Orth X=(\Web X,\Orth{\Pcoh X})$. It results straightforwardly from the definition of PCSs that $\Orth X$ is a PCS. Given $t\in\PCOH(X,Y)$, one has $\Orth t\in\PCOH(\Orth Y,\Orth X)$ if $\Orth t$ is the transpose of $t$, that is $(\Orth t)_{b,a}=t_{a,b}$. One defines $\Tens{X}{Y}$ by $\Web{\Tens{X}{Y}}=\Web{X}\times\Web{Y}$ and \begin{equation*} \Pcohp{\Tens XY}=\Biorth{\{\Tens uv\mid u\in\Pcoh X\text{ and }v\in\Pcoh Y\}} \end{equation*} where $\Tensp uv_{(a,b)}=u_av_b$. Then $\Tens XY$ is a pre-PCS. We have \[ \Pcoh{\Orth{(\Tens{X}{\Orth Y})}}=\Orth{\{\Tens{u}{v'}\mid u\in\Pcoh X\text{ and }v'\in\Pcoh{\Orth Y}\}}=\PCOH(X,Y)\,. \] It follows that $\Limpl XY=\Orth{(\Tens{X}{\Orth Y})}$ is a pre-PCS. Let $(a,b)\in\Web X\times\Web Y$. Since $X$ and $\Orth Y$ are PCSs, there is $A>0$ such that $u_av'_b<A$ for all $u\in\Pcoh X$ and $v'\in\Pcoh{\Orth Y}$. Let $t\in(\Realp)^{\Web{\Limpl XY}}$ be such that $t_{(a',b')}=0$ for $(a',b')\not=(a,b)$ and $t_{(a,b)}=1/A$, we have $t\in\Pcoh{(\Limpl XY)}$. This shows that $\exists t\in\Pcoh{(\Limpl XY)}$ such that $t_{(a,b)}>0$. Similarly we can find $u\in\Pcoh X$ and $v'\in\Pcoh{\Orth Y}$ such that $\epsilon=u_av'_b>0$. It follows that $\forall t\in\Pcoh{(\Limpl XY)}$ one has $t_{(a,b)}\leq 1/\epsilon$. We conclude that $\Limpl XY$ is a PCS, and therefore $\Tens XY$ is also a PCS. \begin{lemma}\label{lemma:PCS-moprh-charact} Let $X$ and $Y$ be PCSs. One has $\Pcoh{(\Limpl XY)}=\PCOH(X,Y)$. That is, given $t\in\Realpto{\Web X\times\Web Y}$, one has $t\in\Pcoh{(\Limpl XY)}$ iff for all $u\in\Pcoh X$, one has $t\Compl u\in\Pcoh Y$. \end{lemma} This results immediately from the definition above of $\Limpl XY$. \begin{lemma}\label{lemma:tens-morph-charact} Let $X_1$, $X_2$ and $Y$ be PCSs. Let $t\in(\Realp)^{\Web{\Limpl{\Tens{X_1}{X_2}}{Y}}}$. One has $t\in\PCOH(\Tens{X_1}{X_2},Y)$ iff for all $u_1\in\Pcoh{X_1}$ and $u_2\in\Pcoh{X_2}$ one has $\Matapp t{(\Tens{u_1}{u_2})}\in\Pcoh Y$. \end{lemma} \proofitem{Proof.} The condition stated by the lemma is clearly necessary. Let us prove that it is sufficient: under this condition, it suffices to prove that \begin{align*} \Orth t\in\PCOH(\Orth Y,\Orth{(\Tens{X_1}{X_2})})\,. \end{align*} Let $v'\in\Pcoh{\Orth Y}$, it suffices to prove that $\Matapp{\Orth t}{v'}\in\Pcoh{\Orth{(\Tens{X_1}{X_2})}}$. So let $u_1\in\Pcoh{X_1}$ and $u_2\in\Pcoh{X_2}$, it suffices to prove that $\Eval{\Matapp{\Orth t}{v'}}{\Tens{u_1}{u_2}}\leq 1$, that is $\Eval{\Matapp t{(\Tens{u_1}{u_2})}}{v'}\leq 1$, which follows from our assumption. \Endproof Let $s_i\in\PCOH(X_i,Y_i)$ for $i=1,2$. Then one defines \begin{align*} \Tens{s_1}{s_2}\in(\Realp)^{\Web{\Limpl{\Tens{X_1}{X_2}}{\Tens{Y_1}{Y_2}}}} \end{align*} by $(\Tens{s_1}{s_2})_{((a_1,a_2),(b_1,b_2))}=(s_1)_{(a_1,b_1)}(s_2)_{(a_2,b_2)}$ and one must check that \[ \Tens{s_1}{s_2}\in\PCOH({\Tens{X_1}{X_2},\Tens{Y_1}{Y_2}})\,. \] This follows directly from Lemma~\ref{lemma:tens-morph-charact}. Let $\One=(\{*\},[0,1])$. There are obvious choices of natural isomorphisms \begin{align*} \Leftu_X&\in\PCOH(\Tens\One X,X)\\ \Rightu_X&\in\PCOH(\Tens X\One,X)\\ \Assoc_{X_1,X_2,X_3}&\in\PCOH(\Tens{\Tensp{X_1}{X_2}}{X_3}, \Tens{X_1}{\Tensp{X_2}{X_3}})\\ \Sym_{X_1,X_2}&\in\PCOH(\Tens{X_1}{X_2},\Tens{X_2}{X_1}) \end{align*} which satisfy the standard coherence properties. This shows that $(\PCOH,\One,\Leftu,\Rightu,\Assoc,\Sym)$ is a symmetric monoidal category. \subsubsection{Internal linear hom.} Given PCSs $X$ and $Y$, let us define $\Evlin\in(\Realp)^{\Web{\Limpl{\Tens{(\Limpl XY)}{X}}{Y}}}$ by \begin{align*} \Evlin_{(((a',b'),a),b)}= \begin{cases} 1 & \text{if }(a,b)=(a',b')\\ 0 & \text{otherwise.} \end{cases} \end{align*} Then it is easy to see that $(\Limpl XY,\Evlin)$ is an internal linear hom object in $\PCOH$, showing that this SMCC is closed. If $t\in\PCOH(\Tens ZX,Y)$, the corresponding linearly curryfied morphism $\Curlin(t)\in\PCOH(Z,\Limpl XY)$ is given by $\Curlin(t)_{(c,(a,b))}=t_{((c,a),b)}$. \subsubsection{$*$-autonomy.} Take $\Bot=\One$, then one checks readily that $(\PCOH,\One,\Leftu,\Rightu,\Assoc,\Sym,\Bot)$ is a $*$-autonomous category. The duality functor $X\mapsto(\Limpl X\Bot)$ can be identified with the strictly involutive contravariant functor $X\mapsto\Orth X$. \subsubsection{Additives}\label{sec:sem-additives} Let $(X_i)_{i\in I}$ be a countable family of PCSs. We define a PCS $\Bwith_{i\in I}X_i$ by $\Web{\Bwith_{i\in I}X_i}=\bigcup_{i\in I}\{i\}\times\Web{X_i}$ and $u\in\Pcohp{\Bwith_{i\in I}X_i}$ if, for all $i\in I$, the family $u(i)\in(\Realp)^{\Web{X_i}}$ defined by $u(i)_a=u_{(i,a)}$ belongs to $\Pcoh{X_i}$. \begin{lemma} Let $u'\in(\Realp)^{\Web{\Bwith_{i\in I}X_i}}$. One has $u'\in\Pcoh{\Orthp{\Bwith_{i\in I}X_i}}$ iff \begin{itemize} \item $\forall i\in I\ u'(i)\in\Pcoh{\Orth{X_i}}$ \item and $\sum_{i\in I}\Norm{u'(i)}_{\Orth{X_i}}\leq 1$. \end{itemize} \end{lemma} The proof is quite easy. It follows that $\Bwith_{i\in I}X_i$ is a PCS. Moreover we can define $\Proj i\in\PCOH(\Bwith_{j\in I}X_j,X_i)$ by \begin{align*} (\Proj i)_{(j,a),a'}= \begin{cases} 1 & \text{if }j=i\text{ and }a=a'\\ 0 & \text{otherwise.} \end{cases} \end{align*} Then $(\Bwith_{i\in I}X_i,(\Proj i)_{i\in I})$ is the cartesian product of the family $(X_i)_{i\in I}$ in the category $\PCOH$. The coproduct $(\Bplus_{i\in I}X_i,(\Inj i)_{i\in I})$ is the dual operation, so that \begin{align*} \Web{\Bplus_{i\in I}X_i}=\Union_{i\in I}\{i\}\times\Web{X_i} \end{align*} and $u\in\Pcoh{(\Bplus_{i\in I}X_i)}$ if $\forall i\in I\ u(i)\in\Pcoh{X_i}$ and $\sum_{i\in I}\Norm{u(i)}_{X_i}\leq 1$. The injections $\Inj j\in\PCOH(X_j,\Bplus_{i\in I}X_i)$ are given by \begin{align*} (\Inj i)_{a',(j,a)}= \begin{cases} 1 & \text{if }j=i\text{ and }a=a'\\ 0 & \text{otherwise.} \end{cases} \end{align*} Given morphisms $s_i\in\PCOH(X_i,Y)$ (for each $i\in I$), then the unique morphism $s\in\PCOH(\Bplus_{i\in I} X_i, Y)$ is given by $s_{(i,a),b}=(s_i)_{a,b}$ and denoted as $\Case_{i\in I}s_i$ (in the binary case, we use $\Case(s_1,s_2)$). \subsubsection{Exponentials} Given a set $I$, a \emph{finite multiset} of elements of $I$ is a function $b:I\to\Nat$ whose \emph{support} $\Supp b=\{a\in I\mid b(a)\not=0\}$ is finite. We use $\Mfin I$ for the set of all finite multisets of elements of $I$. Given a finite family $\List a1n$ of elements of $I$, we use $\Mset{\List a1n}$ for the multiset $b$ such that $b(a)=\Card{\{i\mid a_i=a\}}$. We use additive notations for multiset unions: $\sum_{i=1}^kb_i$ is the multiset $b$ such that $b(a)=\sum_{i=1}^k b_i(a)$. The empty multiset is denoted as $0$ or $\Msetempty$. If $k\in\Nat$, the multiset $kb$ maps $a$ to $k\,b(a)$. Let $X$ be a PCS. Given $u\in\Pcoh X$ and $b\in\Mfin{\Web X}$, we define $u^b=\prod_{a\in\Web X}u_a^{b(a)}\in\Realp$. Then we set $\Prom u=(u^b)_{b\in\Mfin{\Web X}}$ and finally \begin{align*} \Excl X=(\Mfin{\Web X},\Biorth{\{\Prom u\mid u\in\Pcoh X\}}) \end{align*} which is a pre-PCS. We check quickly that $\Excl X$ so defined is a PCS. Let $b=\Mset{\List a1n}\in\Mfin{\Web X}$. Because $X$ is a PCS, and by Theorem~\ref{th:Pcoh-prop}, for each $i=1,\dots,n$ there is $u(i)\in\Pcoh X$ such that $u(i)_{a_i}>0$. Let $(\alpha_i)_{i=1}^n$ be a family of strictly positive real numbers such that $\sum_{i=1}^n\alpha_i\leq 1$. Then $u=\sum_{i=1}^n\alpha_iu(i)\in\Pcoh X$ satisfies $u_{a_i}>0$ for each $i=1,\dots,n$. Therefore $\Prom u_b=u^b>0$. This shows that there is $U\in\Pcoh{(\Excl X)}$ such that $U_b>0$. Let now $A\in\Realp$ be such that $\forall u\in\Pcoh X\,\forall i\in\{1,\dots,n\}\ u_{a_i}\leq A$. For all $u\in\Pcoh X$ we have $u^b\leq A^n$. We have \begin{align*} \Orth{(\Pcoh{(\Excl X)})}=\Triorth{\{\Prom u\mid u\in\Pcoh X\}} =\Orth{\{\Prom u\mid u\in\Pcoh X\}}\,. \end{align*} Let $t\in\Realpto{\Web{\Excl X}}$ be defined by $t_c=0$ if $c\not=b$ and $t_b=A^{-n}>0$; we have $t\in\Orth{(\Pcoh{(\Excl X)})}$. We have exhibited an element $t$ of $\Orth{(\Pcoh{(\Excl X)})}$ such that $t_b>0$. By Lemma~\ref{lemma:pcoh-charact} it follows that $\Excl X$ is a PCS. \subsubsection{Kleisli morphisms as functions.}\label{subsec:Kleisli_fun} Let $s\in\Realpto{\Web{\Limpl{\Excl X}{Y}}}$. We define a function $\Fun s:\Pcoh X\to\Realpcto{\Web Y}$ as follows. Given $u\in\Pcoh X$, we set \begin{align*} \Fun s(u)=s\Compl {\Prom u}=\left(\sum_{c\in\Web{\Excl X}}s_{c,b}u^c\right)_{b\in\Web Y}\,. \end{align*} \begin{theorem}\label{prop:kleisli-morph-charact} One has $s\in\Pcoh{(\Limpl{\Excl X}{Y})}$ iff, for all $u\in\Pcoh X$, one has $\Fun s(u)\in\Pcoh Y$. \end{theorem} This is an immediate consequence of the definition. \begin{theorem}\label{th:pcoh-functional} Let $s\in\PCOH(\Excl X,Y)$. The function $\Fun s$ is Scott-continuous. Moreover, given $s,s'\in\PCOH(\Excl X,Y)$, one has $s=s'$ (as matrices) iff $\Fun s=\Fun{s'}$ (as functions $\Pcoh X\to\Pcoh Y$). \end{theorem} This is an easy consequence of the fact that two polynomials of $n$ variables with real coefficients are identical iff they are the same function on any open subset of $\Real^n$. \emph{Terminology.} We say that $s\in\Pcoh{(\Limpl{\Excl X}Y)}$ is a \emph{power series} whose monomial $u^c$ has coefficient $s_{c,b}$. Since $s$ is characterized by the function $\Fun s$ we sometimes say that $\Fun s$ is a power series. \medskip We can consider the elements of $\Kl\PCOH(X,Y)$ (the morphisms of the Kleisli category of the comonad $\Excl\_$ on the category $\PCOH$) as particular Scott continuous functions $\Pcoh X\to\Pcoh Y$ and this identification is compatible with the definition of identity maps and of composition in $\Kl\PCOH$, see Section~\ref{sec:EM-general}. Of course, not all Scott continuous function are morphisms in $\Kl\PCOH$. \begin{theorem}\label{prop:order-fun-kleiseli} Let $s,s'\in\Kl\PCOH(X,Y)$ be such that $s\leq s'$ (as elements of $\Pcoh{(\Limpl{\Excl X}Y)}$). Then $\forall u\in\Pcoh X\ \Fun s(u)\leq\Fun{s'}(u)$. Let $(s(i))_{i\in\Nat}$ be a monotone sequence of elements of $\Kl\PCOH(X,Y)$ and let $s=\sup_{i\in\Nat}s(i)$. Then $\forall u\in\Pcoh X\ \Fun s(u)=\sup_{i\in I}\Fun{s_i}(u)$. \end{theorem} The first statement is obvious. The second one results from the monotone convergence Theorem. Given a multiset $b\in\Mfin I$, we define its \emph{factorial} $\Factor b=\prod_{i\in I}\Factor{b(i)}$ and its \emph{multinomial coefficient} $\Multinom{}{b}=\Factor{(\Card b)}/\Factor b\in\Natnz$ where $\Card b=\sum_{i\in I}b(i)$ is the cardinality of $b$. Remember that, given an $I$-indexed family $a=(a_i)_{i\in I}$ of elements of a commutative semi-ring, one has the multinomial formula \begin{align*} \Big(\sum_{i\in I}a_i\Big)^n=\sum_{b\in\Mfinc nI}\Multinom{}b a^b \end{align*} where $\Mfinc nI=\{b\in\Mfin I\mid\Card b=n\}$. Given $c\in\Web{\Excl X}$ and $d\in\Web{\Excl Y}$ we define $\Mexpset cd$ as the set of all multisets $r$ in $\Mfin{\Web X\times\Web Y}$ such that \begin{align*} \forall a\in\Web X\ \sum_{b\in\Web Y}r(a,b)=c(a) \quad\text{and}\quad \forall b\in\Web Y\ \sum_{a\in\Web X}r(a,b)=d(b)\,. \end{align*} Let $t\in\PCOH(X,Y)$, we define $\Excl t\in\Realpto{\Limpl{\Excl X}{\Excl Y}}$ by \begin{align*} (\Excl t)_{c,d} =\sum_{r\in\Mexpset cd}\frac{\Factor d}{\Factor r}t^r\,. \end{align*} Observe that the coefficients in this sum are all non-negative integers. \begin{lemma}\label{lemma:excl-morph-app} For all $u\in\Pcoh X$ one has $\Excl t\Compl\Prom u=\Prom{(t\Compl u)}$. \end{lemma} This results from a simple computation applying the multinomial formula. \begin{theorem} For all $t\in\PCOH(X,Y)$ one has $\Excl t\in\PCOH(\Excl X,\Excl Y)$ and the operation $t\mapsto\Excl t$ is functorial. \end{theorem} Immediate consequences of Lemma~\ref{lemma:excl-morph-app} and Theorem~\ref{th:pcoh-functional}. \subsubsection{Description of the exponential comonad.} We equip now this functor with a structure of comonad: let $\Der X\in\Realpto{\Web{\Limpl{\Excl X}X}}$ be given by $(\Der X)_{b,a}=\Kronecker{\Mset a}{b}$ (the value of the Kronecker symbol $\Kronecker ij$ is $1$ if $i=j$ and $0$ otherwise) and $\Digg X\in\Realpto{\Web{\Limpl{\Excl X}{\Excl{\Excl X}}}}$ be given by $(\Digg X)_{b,\Mset{\List b1n}}=\Kronecker{\sum_{i=1}^n b_i}{b}$. Then we have $\Der X\in\PCOH(\Excl X,X)$ and $\Digg X\in\PCOH(\Excl X,\Excl{\Excl X})$ simply because \begin{align*} \Fun{\Der X}(u)=u\quad\text{and}\quad\Fun{\Digg X}(u)=\Prom{(\Prom u)} \end{align*} for all $u\in\Pcoh X$, as easily checked. Using these equations, one also checks easily the naturality of these morphisms, and the fact that $(\Excl\_,\Der{},\Digg{})$ is a comonad. As to the monoidality of this comonad, we introduce $\Expmonisoz\in\Realpto{\Web{\Limpl{\One}{\Excl\top}}}$ by $\Expmonisoz_{\Onelem,\Mset{}}=1$ and $\Expmonisob XY\in\Realpto{\Web{\Limpl{\Tens{\Excl X}{\Excl Y}}{\Excl{(\With XY)}}}}$ by $(\Expmonisob XY)_{b,c,d}=\Kronecker{d}{\Injms 1b+\Injms 2c}$ where $\Injms i{\Mset{\List a1n}}=\Mset{(i,a_1),\dots,(i,a_n)}$. It is easily checked that the required commutations hold (again, we refer to~\cite{Mellies09}). \subsubsection{Fixpoints in $\Kl\PCOH$.} For any object $Y$ of $\PCOH$, a morphism $t\in\Kl\PCOH(Y,Y)$ defines a Scott-continuous function $f=\Fun t:\Pcohp Y\to\Pcohp Y$ which has a least fixpoint $\sup_{n\in\Nat}f^n(0)$. Let $X$ be an object of $\PCOH$ and set $Y=\Limpl{\Exclp{\Limpl{\Excl X}{X}}}X$. Then we have a morphism $t=\Curlin s\in\Kl\PCOH(Y,Y)$ where $s\in\PCOH( \Tens{\Excl Y} {\Exclp{\Limpl{\Excl X}{X}}} ,X)$ is defined as the following composition of morphisms in $\PCOH$: \begin{center} \begin{tikzpicture}[->, >=stealth] \node (1) {$\Tens{\Excl Y}{\Excl{(\Limpl{\Excl X}{X})}}$}; \node (2) [right of=1, node distance=60mm] {$\Excl Y\ITens\Excl{(\Limpl{\Excl X}{X})} \ITens\Excl{(\Limpl{\Excl X}{X})}$}; \node (3) [below of=2, node distance=12mm] {$\Excl X\ITens(\Limpl{\Excl X}{X})$}; \node (4) [below of=2, node distance=24mm] {$X$}; \tikzstyle{every node}=[midway,auto,font=\scriptsize] \draw (1) -- node [below] {$s$} (4); \draw (1) -- node {$\Tens{\Excl Y}{\Contr{\Limpl{\Excl X}{X}}}$} (2); \draw (2) -- node {$\Tens {\Prom{(\Evlin\Compl(\Tens{\Der Y}{\Excl{(\Limpl{\Excl X}{X})}}))}} {\Der{(\Limpl{\Excl X}{X})}}$} (3); \draw (3) -- node {$\Evlin\Compl\Sym$} (4); \end{tikzpicture} \end{center} Then $\Fun t$ is a Scott continuous function $\Pcoh Y\to\Pcoh Y$ whose least fixpoint is $\Sfix$, considered as a morphism $\Sfix\in\Kl\PCOH(\Limpl{\Excl X}{X},X)$, satisfies $\Fun\Sfix(u)=\sup_{n=0}^\infty\Fun u^n(0)$. \subsubsection{The partially ordered class of probabilistic coherence spaces} \newcommand\Subobj{\subseteq} \newcommand\Subwit[2]{\eta_{#1,#2}} We define the category $\Embr\PCOH$. This category is actually a partially ordered class whose objects are those of $\PCOH$. The order relation, denoted as $\Subobj$, is defined as follows: $X\Subobj Y$ if $\Web X\subseteq\Web Y$ and the matrices $\Emb{\Subwit XY}$ and $\Ret{\Subwit XY}$ defined, for $a\in\Web X$ and $b\in\Web Y$, by $(\Emb{\Subwit XY})_{a,b}=(\Ret{\Subwit XY})_{b,a}=\Kronecker ab$ satisfy $\Emb{\Subwit XY}\in\PCOH(X,Y)$ and $\Ret{\Subwit XY}\in\PCOH(Y,X)$. In other words: given $u\in\Pcoh X$, the element $\Matapp{\Emb{\Subwit XY}}{u}$ of $\Realpto{\Web Y}$ obtained by extending $u$ with $0$'s outside $\Web X$ belongs to $\Pcoh Y$. And conversely, given $v\in\Pcoh Y$, the element $\Matapp{\Ret{\Subwit XY}}{v}$ of $\Realpto{\Web X}$ obtained by restricting $v$ to $\Web X$ belongs to $\Pcoh X$. Considering $\Embr\PCOH$ as a category, $\Subwit XY$ is a notation for the unique element of $\Embr\PCOH(X,Y)$ when $X\Subobj Y$, in accordance with the notations of Paragraph~\ref{parag:sem-fix-points}. \begin{lemma}\label{lemma:sem-orth-subobj} If $X\Subobj Y$ then $\Orth X\Subobj\Orth Y$, $\Emb{\Subwit{\Orth X}{\Orth Y}}=\Orth{(\Ret{\Subwit XY})}$ and $\Ret{\Subwit{\Orth X}{\Orth Y}}=\Orth{(\Emb{\Subwit XY})}$. \end{lemma} The proof is a straightforward verification. We contend that $\Embr\PCOH$ is directed co-complete. Let $(X_\gamma)_{\gamma\in\Gamma}$ be a countable directed family in $\Embr\PCOH$ (so $\Gamma$ is a countable directed poset and $\gamma\leq\gamma'\Rightarrow X_\gamma\Subobj X_{\gamma'}$), we have to check that this family has a least upper bound $X$. We set $\Web X=\bigcup_{\gamma\in\Gamma}\Web{X_\gamma}$ and $\Pcoh X=\{w\in\Realpto{\Web X}\mid\forall\gamma\in\Gamma\ \Matapp{\Ret{\Subwit XY}}{w}\in\Pcoh{X_\gamma}\}$. This defines an object of $\PCOH$ which satisfies $\Pcoh X=\Biorth{\{\Matapp{\Emb{\Subwit{X_\gamma}{X}}}{u}\mid\gamma\in\Gamma\text{ and }u\in\Pcoh{X_\gamma}\}}$ and is therefore the lub of the family $(X_\gamma)_{\gamma\in\Gamma}$ in $\Embr\PCOH$. This object $X$ is denoted $\bigcup_{\gamma\in\Gamma}X_\gamma$. One checks easily that $\Orth{(\bigcup_{\gamma\in\Gamma}X_\gamma)} =\bigcup_{\gamma\in\Gamma}\Orth{X_\gamma}$. Then the functor $\Embf:\Embr\PCOH\to\PCOH$ defined by $\Embf(X)=X$ and $\Embf(\Subwit XY)=\Emb{\Subwit XY}$ is continuous: given a directed family $(X_\gamma)_{\gamma\in\Gamma}$ whose lub is $X$ and given a collection of morphisms $t_\gamma\in\PCOH(X_\gamma,Y)$ such that $t_{\gamma'}\Compl\Emb{\Subwit{X_{\gamma}}{X_{\gamma'}}}=t_\gamma$ for any $\gamma,\gamma'\in\Gamma$ such that $\gamma\leq\gamma'$, there is exactly one morphism $t\in\PCOH(X,Y)$ such that $t\Compl\Emb{\Subwit{X_\gamma}{X}}=t_\gamma$ for each $\gamma\in\Gamma$. Given $a\in\Web X$ and $b\in\Web Y$, $t_{a,b}=(t_\gamma)_{a,b}$ for any $\gamma$ such that $a\in\Web{X_\gamma}$ (our hypothesis on the $t_\gamma$'s means that $(t_\gamma)_{a,b}$ does not depend on the choice of $\gamma$). All the operations of Linear Logic define monotone continuous functionals on $\Embr\PCOH$ which moreover commute with the functor $\Funofemb$. This means for instance that if $X\Subobj Y$ then $\Excl X\Subobj\Excl Y$, $\Emb{\Subwit{\Excl X}{\Excl Y}}=\Excl{(\Emb{\Subwit{X}{Y}})}$, $\Ret{\Subwit{\Excl X}{\Excl Y}}=\Excl{(\Ret{\Subwit{X}{Y}})}$ and $\Excl{(\bigcup_{\gamma\in\Gamma}X_\gamma)}= \bigcup_{\gamma\in\Gamma}\Excl{X_\gamma}$ and similarly for $\ITens$ and $\IPlus$. As a consequence, and as a consequence of Lemma~\ref{lemma:sem-orth-subobj}, if $X_i\Subobj Y_i$ for $i=1,2$ then $\Limpl{X_1}{X_2}\Subobj\Limpl{Y_1}{Y_2}$, $\Emb{\Subwit{\Limpl{X_1}{X_2}}{\Limpl{Y_1}{Y_2}}} =\Limpl{\Ret{\Subwit{X_1}{Y_1}}}{\Emb{\Subwit{X_1}{Y_1}}}$ and $\Ret{\Subwit{\Limpl{X_1}{X_2}}{\Limpl{Y_1}{Y_2}}} =\Limpl{\Emb{\Subwit{X_1}{Y_1}}}{\Ret{\Subwit{X_1}{Y_1}}}$ and $\multimap$ commutes with directed colimits in $\Embr\PCOH$. \newcommand\Coalgmt[1]{\Coalgm{#1}} This notion of inclusion on probabilistic coherence spaces extends to coalgebras as outlined in Section~\ref{parag:sem-fix-points} (again, we refer to~\cite{Ehrhard16a} for more details). We describe briefly this notion of inclusion in the present concrete setting. Let $P$ and $Q$ be object of $\EM\PCOH$, we have $P\Subobj Q$ in $\Embr{\EM\PCOH}$ if $\Coalgc P\Subobj\Coalgc Q$ and $\Coalgm Q\Compl\Emb{\Subwit{\Coalgc P}{\Coalgc Q}}=\Excl{(\Emb{\Subwit{\Coalgc P}{\Coalgc Q}})}\Compl\Coalgm P$. The lub of a directed family $(P_\gamma)_{\gamma\in\Gamma}$ of coalgebras (for this notion of substructure) is the coalgebra $P=\bigcup_{\gamma\in\Gamma}P_\gamma$ defined by $\Coalgc P=\bigcup_{\gamma\in\Gamma}\Coalgc{P_\gamma}$ and $\Coalgm P$ is characterized by the equation $\Coalgm P\Compl\Emb{\Subwit{\Coalgc{P_\gamma}}{\Coalgc P}}=\Excl{\Emb{\Subwit{\Coalgc{P_\gamma}}{\Coalgc P}}}\Compl\Coalgm{P_\gamma}$ which holds for each $\gamma\in\Gamma$. As outlined in Section~\ref{sec:cbpv-interpretation}, this allows to interpret any type $\sigma$ as an object $\Tsem\sigma$ of $\PCOH$ and any positive type $\phi$ as an object $\Tsemca\phi$ such that $\Coalgc{\Tsemca\phi}=\Tsem\phi$, in such a way that the coalgebras $\Tsemca{\TREC\zeta\phi}$ and $\Tsemca{\Subst{\phi}{\TREC\zeta\phi}{\phi}}$ are \emph{exactly the same objects of $\EM\PCOH$}. We use $\Coalgmt\phi$ for $\Coalgm{\Tsemca\phi}$. \subsubsection{Dense coalgebras}\label{sec:reg-coalgebras} Let $P$ be an object of $\EM\PCOH$, so that $P=(\Coalgc P,\Coalgm P)$ where $\Coalgc P$ is a probabilistic coherence space and $\Coalgm P\in\PCOH(\Coalgc P,\Excl{\Coalgc P})$ satisfies $\Digg{\Coalgc P}\Compl\Coalgm P=\Excl{\Coalgm P}\Compl\Coalgm P$. Given coalgebras $P$ and $Q$, a morphism $t\in\PCOH(\Coalgc P,\Coalgc Q)$ is coalgebraic (that is $t\in\EM\PCOH(P,Q)$) if $\Coalgm Q\Compl t=\Excl t\Compl\Coalgm P$. In particular, we say that $u\in\Pcoh{(\Coalgc P)}$ is coalgebraic if, considered as a morphism from $\One$ to $\Coalgc P$, $u$ belongs to $\EM\PCOH(\One,P)$. This means that $\Prom u={\Coalgm P}\Compl u$. \begin{definition}\label{def:dense} Given an object $P$ of $\EM\PCOH$, we use $\PcohEM{(P)}$ for the set of coalgebraic elements of $\Pcoh{(\Coalgc P)}$. \end{definition} The following lemma is useful in the sequel and holds in any model of Linear Logic. \begin{lemma}\label{lemma:sem-values} Let $X$ be a probabilistic coherence space, one has $\PcohEM{(\Excl X)}=\{\Prom u\mid u\in\Pcoh X\}$. Let $P_\ell$ and $P_r$ be objects of $\EM\PCOH$. $\Tens{P_\ell}{P_r}$ is the cartesian product of $P_\ell$ and $P_r$ in $\EM\PCOH$. The function $\PcohEM{(P_\ell)}\times\PcohEM{(P_r)}\to\PcohEM{(\Tens{P_\ell}{P_r})}$ which maps $(u,v)$ to $\Tens uv$ is a bijection. The projections $\Projt i\in\EM\PCOH(\Tens{P_\ell}{P_r},P_i)$ are characterized by $\Projt i(\Tens{u_\ell}{u_r})=u_i$. The function $\{\ell\}\times\PcohEM{(P_\ell)}\cup\{r\}\times\PcohEM{(P_r)} \to\PcohEM{(\Plus{P_\ell}{P_r})}$ which maps $(i,u)$ to $\Inj i(u)$ is a bijection. The injection $u\mapsto\Inj i(u)$ has a left inverse $\Proj i\in\PCOH(\Plus{\Coalgc{P_\ell}}{\Coalgc{P_r}},\Coalgc{P_i})$ defined by $(\Proj i)_{(j,a),b}=\Kronecker{i}{j}\Kronecker{a}{b}$, which is not a coalgebra morphism in general. \end{lemma} \begin{proof} Let $v\in\PcohEM{(\Excl X)}$, we have $\Prom v={\Coalgm{\Excl X}}\Compl v=\Digg X\Compl v$ hence $\Prom{(\Der X\Compl v)}=\Excl{\Der X}\Compl\Prom v=\Excl{\Der X}\Compl\Digg X v=v$. The other properties result from the fact that the Eilenberg-Moore category $\EM\PCOH$ is cartesian and co-cartesian with $\ITens$ and $\IPlus$ as product and co-product, see~\cite{Mellies09} for more details. \end{proof} Because of these properties we write sometimes $(u_\ell,u_r)$ instead of $\Tens{u_\ell}{u_r}$ when $u_i\in\PcohEM{P_i}$ for $i\in\{\ell,r\}$. \begin{definition}\label{def:coalg-dense} An object $P$ of $\EM\PCOH$ is \emph{dense} if, for any object $Y$ of $\PCOH$ and any two morphisms $t,t'\in\PCOH(\Coalgc P,Y)$, if $t\Compl u=t'\Compl u$ for all $u\in\PcohEM{(P)}$, then $t=t'$. \end{definition} \begin{theorem}\label{th:coalg-dense-colsed} For any probabilistic coherence space $X$, $\Excl X$ is a dense coalgebra. If $P_\ell$ and $P_r$ are dense coalgebras then $\Tens{P_\ell}{P_r}$ and $\Plus{P_\ell}{P_r}$ are dense. The colimit in $\Embr{(\EM\PCOH)}$ of a directed family of dense coalgebras is dense. \end{theorem} \begin{proof} Let $X$ be an object of $\PCOH$, one has $\PcohEM{(\Excl X)}=\{\Prom u\mid u\in\Pcoh{X}\}$ by Lemma~\ref{lemma:sem-values}. It follows that $\Excl X$ is a dense coalgebra by Theorem~\ref{th:pcoh-functional}. Assume that $P_\ell$ and $P_r$ are dense coalgebras. Let $t,t'\in\PCOH(\Tens{\Coalgc{P_\ell}}{\Coalgc{P_r}},Y)$ be such that $t\Compl w=t'\Compl w$ for all $w\in\PcohEM{(\Tens{P_\ell}{P_r})}$. We have $\Curlin{(t)},\Curlin{(t')}\in\PCOH(\Coalgc{P_\ell},\Limpl{\Coalgc P_r}{Y})$ so, using the density of $P_\ell$, it suffices to prove that $\Curlin{(t)}\Compl u_\ell=\Curlin{(t')}\Compl u_\ell$ for each $u_\ell\in\PcohEM{(P_\ell)}$. So let $u_\ell\in\PcohEM{(P_\ell)}$ and let $s=\Curlin{(t)}\Compl u_\ell$ and $s'=\Curlin{(t')}\Compl u_\ell$. Let $u_r\in\PcohEM{(P_r)}$, we have $s\Compl u_r=t\Compl(\Tens{u_\ell}{u_r})=t'\Compl(\Tens{u_\ell}{u_r})=s'\Compl u_r$ since $\Tens{u_\ell}{u_r}\in\PcohEM{(\Tens{P_\ell}{P_r})}$ and therefore $s=s'$ since $P_r$ is dense. Let now $t,t'\in\PCOH(\Plus{\Coalgc{P_\ell}}{\Coalgc{P_r}},Y)$ be such that $t\Compl w=t'\Compl w$ for all $w\in\PcohEM{(\Plus{P_\ell}{P_r})}$. To prove that $t=t'$, it suffices to prove that $t\Compl\Inj i=t'\Compl\Inj i$ for $i\in\{\ell,r\}$. Since $P_i$ is dense, it suffices to prove that $t\Compl\Inj i\Compl u=t'\Compl\Inj i\Compl u$ for each $u\in\PcohEM{(P_i)}$ which follows from the fact that $\Inj i\Compl u\in\PcohEM{P_i}$. Last let $(P_\gamma)_{\gamma\in\Gamma}$ be a directed family of dense coalgebras (in $\Embr{\EM\PCOH}$) and let $P=\bigcup_{\gamma\in\Gamma}P_\gamma$, and let $t,t'\in\EM\PCOH(\Coalgc P,Y)$ be such that $t\Compl w=t'\Compl w$ for all $w\in\PcohEM{(P)}$. It suffices to prove that, for each $\gamma\in\Gamma$, one has $t\Compl\Emb{\Subwit{\Coalgc{P_\gamma}}{\Coalgc P}}=t'\Compl\Emb{\Subwit{\Coalgc{P_\gamma}}{\Coalgc P}}$ and this results from the fact that $P_\gamma$ is dense and $\Emb{\Subwit{\Coalgc{P_\gamma}}{\Coalgc P}}$ is a coalgebra morphisms (and therefore maps $\PcohEM{(P_\gamma)}$ to $\PcohEM{(P)}$). \end{proof} The sub-category $\EM\PCOH$ of dense coalgebras is cartesian and co-cartesian and is well-pointed by Theorem~\ref{th:coalg-dense-colsed}. We use $\EMR\PCOH$ for this sub-category and $\Embr{(\EMR\PCOH)}$ for the sub-class of $\Embr{\EM\PCOH}$ whose objects are the dense coalgebras (with the same order relation). \subsubsection{Interpreting types and terms in $\PCOH$}\label{subsec:sem-type-term} Given a type $\sigma$ with free type variables contained in the repetition-free list $\Vect\zeta$, and given a sequence $\Vect P$ of length $n$ of objects of $\EM\PCOH$, we define $\Tsem\sigma_{\Vect\zeta}(\Vect P)$ as an object of $\PCOH$ and when $\phi$ is a positive type (whose free variables are contained in $\Vect\zeta$) we define $\Tsemca\phi_{\Vect\zeta}(\Vect P)$ as an object of $\EM\PCOH$. These operations are continuous and their definition follows the general pattern described in Section~\ref{sec:cbpv-interpretation}. \begin{theorem}\label{th:pos-types-dense} Let $\phi$ be a positive type and let $\Vect\zeta=(\List\zeta 1n)$ be a repetition-free list of type variables which contains all the free variables of $\phi$. Let $\Vect P$ be a sequence of $n$ dense coalgebras. Then $\Tsemca\phi_{\Vect\zeta}(\Vect P)$ is a dense coalgebra. In particular, when $\phi$ is closed, the coalgebra $\Tsemca\phi$ is dense. \end{theorem} This is an immediate consequence of the definition of $\Tsemca\phi$ and of Theorem~\ref{th:coalg-dense-colsed} \begin{remark} This means that we have \emph{only one model}, where positive types are interpreted as objects of $\EM\PCOH$, and not two models as one could expect (one where positive types are interpreted in $\EM\PCOH$, and one where they are interpreted in $\EMR\PCOH$); it simply turns out that the interpretation of positive types in the model $\PCOH/\EM\PCOH$ are dense coalgebras. This is mainly due to the fact that the colimit of a directed family of dense coalgebras \emph{in the partially ordered class $\Embr{\EM\PCOH}$} is dense, see Theorem~\ref{th:coalg-dense-colsed}. From the viewpoint of Levy's CBPV~\cite{LevyP04}, whose semantics is described in terms of adjunctions, we are using a resolution of the comonad ``$\oc$'' through the category $\EM\PCOH$ (or, equivalently, through the category $\EMR\PCOH$). As pointed out to us by one of the referees and already mentioned in the introduction, there is another resolution through a category of families and introduced in~\cite{Abramsky1998}, which is initial among all resolutions that model CBPV. This other option will be explored in further work. \end{remark} Then $\BOOL=\PLUS\ONE\ONE$ satisfies $\Web{\Tsem\BOOL}=\{(\ell,\Onelem),(r,\Onelem)\}$ and $u\in\Realpto{\Web{\Tsem\BOOL}}$ satisfies $u\in\Pcoh{\Tsem\BOOL}$ iff $u_{(\ell,\Onelem)}+u_{(r,\Onelem)}\leq 1$. The coalgebraic structure of this object is given by \begin{align*} (\Coalgmt\BOOL)_{(j,\Onelem),\Mset{(j_1,\Onelem),\dots,(j_k,\Onelem)}}= \begin{cases} 1 & \text{if }j=j_1=\dots=j_k\\ 0 & \text{otherwise.} \end{cases} \end{align*} The object $\Snat=\Tsem\NAT$ satisfies $\Snat=\PLUS\One\Snat$ so that $\Web\SNat=\{(\ell,\Onelem),(r,(\ell,\Onelem)),(r,(r,(\ell,\Onelem))),\dots\}$ and we use $\Snum n$ for the element of $\Web\Snat$ which has $n$ occurrences of $r$. Given $u\in\Realpto{\Web\Snat}$, we use $l(u)$ for the element of $\Realpto{\Web\Snat}$ defined by $l(u)_{\Snum n}=u_{\Snum{n+1}}$. By definition of $\Snat$, we have $u\in\Pcoh\Snat$ iff $u_{\Snum 0}+\Norm{l(u)}_\Snat\leq 1$, and then $\Norm u_\Snat=u_{\Snum 0}+\Norm{l(u)}_\Snat$. It follows that $u\in\Pcoh\Snat$ iff $\sum_{n=0}^\infty u_{\Snum n}\leq 1$ and of course $\Norm u_\Snat=\sum_{n=0}^\infty u_{\Snum n}$. Then the coalgebraic structure $\Coalgmt\NAT$ is defined exactly as $\Coalgmt\BOOL$ above. In the sequel, we identify $\Web\Snat$ with $\Nat$. Given a typing context $\cP=(x_1:\phi_1,\dots,x_k:\phi_k)$, a type $\sigma$ and a term $M$ such that $\TSEQ\cP M\sigma$, $M$ is interpreted as a morphism $\Psem M^\cP\in\PCOH(\Tsem\cP,\Tsem\sigma)$. For all constructs of the language but probabilistic choice, this interpretation uses the generic structures of the model described in Section~\ref{sec:LL-semantics-short}, the description of this interpretation can be found in~\cite{Ehrhard16a}. We set $\Psem{\COIN p}{}=p\Base{(\ell,*)}+(1-p)\Base{(r,*)}$. If $\TSEQ{x_1:\phi_1,\dots,x_k:\phi_k}{M}{\sigma}$, the morphism $\Psem{M}^\cP$ is completely characterized by its values on $(u_1,\dots,u_k)\in\PcohEM{(\Tsemca\cP)}$. We describe now the interpretation of terms using this observation. \begin{itemize} \item $\Psem\ONELEM=1\in\Pcoh{\One}=[0,1]$. \item $\Psem{x_i}^\cP(\List u1k)=u_i$. \item $\Psem{\STOP N}^\cP(\List u1k)=\Prom{(\Psem N^\cP(\List u1k))}$. \item $\Psem{\PAIR{M_\ell}{M_r}}^\cP(\List u1k)= \Tens{\Psem{M_\ell}^\cP(\List u1k)}{\Psem{M_r}^\cP(\List u1k)}$. \item $\Psem{\IN i{N}}^\cP(\List u1k)=\Inj i(\Psem N^\cP(\List u1k))$, $i\in\{\ell,r\}$. \item $\Psem{\GO N}^{\cP}(\List u1k)=\Der{\Tsem\sigma}(\Psem{N}^\cP(\List u1k))$, assuming that $\TSEQ{\cP}{N}{\EXCL\sigma}$. \item If $\TSEQ{\cP}{N}{\LIMPL{\phi}{\sigma}}$ and $\TSEQ{\cP}{R}{\phi}$ then $\Psem{N}^\cP(\List u1k)\in\Pcoh{(\Limpl{\Tsem\phi}{\Tsem\sigma})}$, and $\Psem{R}^\cP(\List u1k)\in\Pcoh{(\Tsem\phi)}$ and using the application of a matrix to a vector we have $\Psem{\LAPP NR}^\cP(\List u1k)=\Psem{N}^\cP(\List u1k)\Compl\Psem{R}^\cP(\List u1k)$. \item If $\TSEQ{\cP,x:\phi}{N}{\sigma}$ then $\Psem{\ABST{x}{\phi}{N}}^\cP(\List u1k)\in\Pcoh{(\Limpl{\Tsem\phi}{\Tsem\sigma})}$ is completely described by the fact that, for all $u\in\PcohEM{(\Tsemca\phi)}$, one has $\Psem{\ABST{x}{\phi}{N}}^\cP(\List u1k)\Compl u=\Psem{N}^{\cP,x:\phi}(\List u1k,u)$. This is a complete characterization of this interpretation by Theorem~\ref{th:pos-types-dense}. \item If $\TSEQ{\cP}{N}{\PLUS{\phi_\ell}{\phi_r}}$ and $\TSEQ{\cP,y_i:\phi_i}{R_i}{\sigma}$ for $i\in\{\ell,r\}$, then\\ $\Psem{\CASE{N}{y_\ell}{R_\ell}{y_r}{R_r}}^\cP(\List u1k)=\Psem{R_\ell}^{\cP,y_\ell:\phi_\ell}(\List u1k,\Proj \ell(\Psem{N}^\cP(\List u1k)))+\Psem{R_r}^{\cP,y_r:\phi_r}(\List u1k,\Proj r(\Psem{N}^\cP(\List u1k)))$ where $\Proj i\in\PCOH{(\Plus{P_\ell}{P_r},P_i)}$ is the $i$th ``projection'' introduced in~\ref{sec:reg-coalgebras}, left inverse for $\Inj i$. \item If $\TSEQ{\cP,x:\EXCL\sigma}{N}{\sigma}$ then $\Psem{N}^{\cP,x:\EXCL\sigma} \in\PCOH(\Tens{\Coalgc{\Tsem\cP}}{\Excl{\Tsem\sigma}},\Tsem\sigma)$ and $\Psem{\FIXT x\sigma N}^\cP(\List u1k)=\sup_{n=0}^\infty f^n(0)$ where $f:\Pcoh{\Tsem\sigma}\to \Pcoh{\Tsem\sigma}$ is the Scott-continuous function given by $f(u)=\Psem{N}^{\cP,x:\Excl\sigma}(\List u1k,\Prom u)$. \item If $\TSEQ{\cP}{N}{\Subst\psi{\TREC\zeta\psi}\zeta}$ then $\Psem{\FOLD N}^\cP=\Psem N^\cP$ which makes sense since $\Tsem{\Subst\psi{\TREC\zeta\psi}\zeta}=\Tsem{\TREC\zeta\psi}$. \item If $\TSEQ{\cP}{N}{\TREC\zeta\psi}$ then $\Psem{\UNFOLD N}^\cP=\Psem N^\cP$. \end{itemize} \begin{theorem}[Soundness]\label{th:soundness} If $M$ satisfies $\TSEQ{\cP}{M}{\sigma}$ then \begin{align*} \Psem M^\cP=\sum_{\TSEQ{\cP}{M'}{\sigma}}\Redmats_{M,M'}\Psem{M'}^{\cP} \end{align*} \end{theorem} The proof is done by induction and is a straightforward verification. \begin{corollary}\label{th:soundness-ineq} Let $M$ be a term such that $\Tseq{}{M}{\One}$ so that $\Psem M\in\Izu$. Then $\Psem M\geq\Redmats^\infty_{M,\ONELEM}$. \end{corollary} This is an immediate consequence of Theorem~\ref{th:soundness} and of the definition of $\Redmats^\infty$, see Section~\ref{sec:obs-eq}. \subsection{Examples of term interpretations}\label{subsec:exden} We give the interpretation of terms that we gave as examples in Subsection~\ref{subsec:exsyn}. \begin{itemize} \item $\Psem{\LOOP\sigma}=\Psem{\FIXT x\sigma{\GO x}}=0$ \item $\Psem{\True}= \Base{(\ell,*)}$ and $\Psem{\False}=\Base{(r,*)}$ \item $\Psem{\IFB M{N_\ell}{N_r}}^\cP(\List u1k)=\Psem{M}^\cP_{(\ell,*)}(\List u1k)\Psem{N_\ell}^\cP(\List u1k)\\+\Psem{M}^\cP_{(r,*)}(\List u1k)\Psem{N_r}^\cP(\List u1k)$ \item $\Psem{\DICE p{M_\ell}{M_r}}^\cP(\List u1k)=p\Psem{M_\ell}^\cP(\List u1k)+(1-p)\Psem{M_r}^\cP(\List u1k)$ \item $\Psem{\NUM n}=\Snum n$ for $n\in\Nat$ \item $\Psem{\TSUCC M}^\cP_{n+1}(\List u1k)= \Psem{M}_n(\List u1k)$ \item $\Psem{\IFV M{N_\ell}x{N_r}}^\cP(\List u1k)=\Psem{M}^\cP_0(\List u1k)\Psem{N_\ell}^\cP(\List u1k)\\+\sum_{n=0}^\infty\Psem{M}^\cP_{n+1}(\List u1k)\Psem{N_r}^\cP(\List u1k)(\Snum n)$ \item $\Psem{\Ran{\Vect p}}=\sum_{i=1}^n p_i \Base{\Snum i}$ \item $\Psem{\LAPP{\Probe_\ell}M}^\cP(\List u1k)= \Psem{M}^\cP(\List u1k)_\ell\Base{*}$ \item $\Psem{M_0\cdot N}^\cP(\List u1k)=\Psem{M_0}^\cP(\List u1k)\Psem N^\cP(\List u1k)$ \item $\Psem{M_0\AND\cdots\AND M_l}^\cP(\List u1k)=\prod_{i=0}^l\Psem{M_i}^\cP(\List u1k)$ \item $\Psem{\LAPP{\Pchoose^\sigma_{l+1}(\List N0l)}P}^\cP(\List u1k)=\sum_{i=0}^l \Psem{P}_i^\cP(\List u1k)\cdot \Psem{N_i}^\cP(\List u1k)$ \item $\forall u \in\Pcoh{(\Tsem{\Tnat})},\ \Psem{\Pext lr}(u)=\sum_{i=l}^ru_i \Base{\Snum i}$ and \[\Psem{\Pwin l{\Vect n}}(u)=\sum_{i=n_1+\cdots+n_{l-1}+1}^{n_1+\cdots+n_l}u_i \Base{\Snum{i}}\] \end{itemize} \section{Adequacy}\label{sec:adequacy} Our goal is to prove the converse of Corollary~\ref{th:soundness-ineq}: for any closed term $M$ such that $\TSEQ{}{M}{\ONE}$, the probability that $M$ reduces to $\ONELEM$ is larger than or equal to $\Psem M{}\in\Pcoh{\Tsem\ONE}\Isom[0,1]$, so that we shall know that these two numbers are actually equal. In spite of its very simple statement, the proof of this property is rather long mainly because we have to deal with the recursive type definitions allowed by our syntax. As usual, the proof is based on the definition of a logical relations between terms and elements of the model (more precisely, given any type $\sigma$, we have to define a relation between closed terms of types $\sigma$ and elements of $\Pcoh{\Tsem\sigma}$; let us call such a relation a \emph{$\sigma$-relation}). Since we have no positivity restrictions on the occurrence of type variables wrt.~which recursive types are defined so that types are neither covariant nor contravariant wrt.~these type variables, we use a very powerful technique introduced in~\cite{Pitts93} for defining this logical relation. Indeed a type variable $\zeta$ can have positive and negative occurrences in a positive\footnote{Warning: the word ``positive'' has two different meanings here!} type $\phi$, consider for instance the case $\phi=\EXCL{(\LIMPL\zeta\zeta)}$ where the type variable $\zeta$ has a positive (on the right of the $\LIMPL{}{}$) and a negative occurrence (on the left). To define the logical relation associated with $\TREC\zeta\phi$, we have to find a fixpoint for the operation which maps a $(\TREC\zeta{\phi})$-relation $\cR$ to the relation $\Phi(\cR)=\EXCL{(\LIMPL{\cR}{\cR})}$ (which can be defined using $\cR$ as a ``logical relation'' in a fairly standard way). Relations are naturally ordered by inclusion, and this strongly suggests to define the above fixpoint using this order relation by \Eg{}~Tarski's Fixpoint Theorem. The problem however is that $\Phi$ is neither a monotone nor an anti-monotone operation on relations, due to the fact that $\zeta$ has a positive and a negative occurrence in $\phi$. It is here that Pitts's trick comes in: we replace the relations $\cR$ with pairs of relations $\cR=(\Nrel\cR,\Prel\cR)$ ordered as follows: $\cR\Subrel\cS$ if $\Prel\cR\subseteq\Prel\cS$ and $\Nrel\cS\subseteq\Nrel\cR$. Then we define accordingly $\Phi(\cR)$ as a pair of relations by $\Nrel{\Phi(\cR)}=\EXCL{(\LIMPL{\Prel\cR}{\Nrel\cR})}$ and $\Prel{\Phi(\cR)}=\EXCL{(\LIMPL{\Nrel\cR}{\Prel\cR})}$. Now the operation $\Phi$ is monotone wrt.~the $\Subrel$ relation and it becomes possible to apply Tarski's Fixpoint Theorem to $\Phi$ and get a pair of relations $\cR$ such that $\cR=\Phi(\cR)$. The next step consists in proving that $\Nrel\cR=\Prel\cR$. This is obtained by means of an analysis of the definition of the interpretation of fixpoints of types as colimits in the category $\Embr\PCOH$. One is finally in position of proving a fairly standard ``Logical Relation Lemma'' from which adequacy follows straightforwardly. In this short description of our adequacy proof, many technicalities have obviously been hidden, the most important one being that values are handled in a special way so that we actually consider two kinds of pairs of relations. Also, a kind of ``biorthogonality closure'' plays an essential role in the handling of positive types, no surprise for the readers acquainted with Linear Logic, see for instance the proof of normalization in~\cite{Girard87}. \subsection{Pairs of relations and basic operations} Given a \emph{closed} type $\sigma$, we define $\Rels\sigma$ as the set of all pairs of relations $\cR=(\Nrel\cR,\Prel\cR)$ such that, for $\epsilon\in\{\POS,\NEG\}$, each element of $\Srel\cR\epsilon$ is a pair $(M,u)$ where $\TSEQ{}{M}{\sigma}$ and $u\in\Pcoh{\Tsem\sigma}$. For a \emph{closed} positive type $\phi$, we also define $\Relsv\phi$ as the set of all pairs of relations $\cV=(\Nrel\cV,\Prel\cV)$ such that, for $\epsilon\in\{\POS,\NEG\}$, each element of $\Srel\cV\epsilon$ is a pair $(V,v)$ where $\TSEQ{}{V}{\phi}$ is a value and $v\in\PcohEM{\Tsem\phi}$. Given $\cR,\cS\in\Rels\sigma$, we write $\cR\Subrel\cS$ if $\Prel\cR\subseteq\Prel\cS$ and $\Nrel\cS\subseteq\Nrel\cR$. We define similarly $\cV\Subrel\cW$ for $\cV,\cW\in\Relsv\phi$. Then $\Rels\sigma$ is a complete meet-lattice, the infimum of a collection $(\cR_i)_{i\in I}$ being $\Infrel_{i\in I}\cR_i=(\bigcup_{i\in I}\Nrel\cR_i,\bigcap_{i\in I}\Prel\cR_i)$. The same holds of course for $\Relsv\phi$ and we use the same notations. We define $\cR(\ONE)$ as the set of all pairs $(M,p)$ such that $\TSEQ{}{M}{\ONE}$, $p\in[0,1]$ and $\Redmats^\infty_{M,\ONELEM}\geq p$. We define in Figure~\ref{fig:adeq-rel-log-op} logical operations on these pairs of relations. The last one is the aforementioned biorthogonality closure operation on pairs of relations. \begin{figure} \centering \begin{itemize} \item Let $\cR\in\Rels\sigma$, we define $\EXCL\cR\in\Relsv{\EXCL\sigma}$ by: $\Srel{\EXCL\cR}\epsilon=\{(\STOP M,\Prom u)\mid (M,u)\in\Srel\cR\epsilon\}$ for $\epsilon\in\{\NEG,\POS\}$. \item Let $\cV_i\in\Relsv{\phi_i}$ for $i\in\{\ell,r\}$. We define $\Srel{(\TENS{\cV_\ell}{\cV_r})}\epsilon=\{(\PAIR{V_\ell}{V_r},\Tens{v_\ell}{v_r})\}\mid (V_i,v_i)\in\Srel{\cV_i}\epsilon\}$ for $\epsilon\in\{\NEG,\POS\}$, so that $\TENS{\cV_\ell}{\cV_r}\in\Relsv{\TENS{\phi_\ell}{\phi_r}}$. \item Let $\cV_i\in\Relsv{\phi_i}$ for $i\in\{\ell,r\}$. We define $\Srel{(\PLUS{\cV_\ell}{\cV_r})}\epsilon=\{(\IN i{V},\Inj i(v))\}\mid i\in\{\ell,r\}\text{ and } (V,v)\in\Srel{\cV_i}\epsilon\}$ for $\epsilon\in\{\NEG,\POS\}$, so that $\PLUS{\cV_\ell}{\cV_r}\in\Relsv{\PLUS{\phi_\ell}{\phi_r}}$. \item Let $\cV\in\Relsv\phi$ and $\cR\in\Rels\sigma$. We define $\LIMPL\cV\cR\in\Rels{\LIMPL\phi\sigma}$ as follows: $\Srel{(\LIMPL\cV\cR)}\epsilon=\{(M,u)\mid\ \TSEQ{}{M}{\LIMPL\phi\sigma},\ u\in\Pcoh{\Tsem{\LIMPL\phi\sigma}}\text{ and }\forall (V,v)\in\Srel\cV{-\epsilon}\ (\LAPP MV,u\Matapp v)\in\Srel\cR\epsilon\}$. \item Last, given $\cV\in\Relsv{\phi}$, we define $\Crel\cV\in\Rels{\phi}$ as follows: $\Srel{\Crel\cV}\epsilon$ is the set of all $(M,u)$ such that $\TSEQ{}{M}{\phi}$, $u\in\Pcoh{\Tsem\phi}$ and, for all $(T,t)\in\Srel{(\LIMPL\cV\cR(\ONE))}{-\epsilon}$, one has $(\LAPP TM,\Matapp tu)\in\cR(\ONE)$. \end{itemize} \caption{Logical operations for pairs of relations} \label{fig:adeq-rel-log-op} \end{figure} Observe that all these operations are monotone wrt.~$\Subrel$. For instance $\cV\Subrel\cW\Andc\cR\Subrel\cS\Rightarrow(\LIMPL\cV\cR)\Subrel(\LIMPL\cW\cS)$, and $\cV\Subrel\cW\Rightarrow\Crel\cV\Subrel\Crel\cW$. \subsection{Fixpoints of pairs of relations} To deal with fixpoint types $\TREC\zeta\phi$, we need to consider types parameterized by relations as follows. Let $\sigma$ be a type and let $\Vect\zeta=(\List\zeta 1n)$ be a list of type variables without repetitions and which contains all free variables of $\sigma$. For all list $\Vect\phi=(\List\phi 1n)$ of \emph{closed} positive types, we define \begin{align*} \Trel\sigma{\Vect\zeta}:\prod_{i=1}^n\Relsv{\phi_i} \to\Rels{\Subst\sigma{\Vect\phi}{\Vect\zeta}}\,. \end{align*} Let also $\phi$ be a positive type whose free variables are contained in $\Vect\zeta$, we define \begin{align*} \Trelv\phi{\Vect\zeta}:\prod_{i=1}^n\Relsv{\phi_i} \to\Relsv{\Subst\phi{\Vect\phi}{\Vect\zeta}}\,. \end{align*} The definition is by simultaneous induction on $\sigma$ and $\phi$. All cases but one consist in applying straightforwardly the above defined logical operations on pairs of relations, for instance \begin{align*} \Trel{\LIMPL\phi\tau}{\Vect\zeta}(\Vect\cV) =\LIMPL{\Trelv\phi{\Vect\zeta}(\Vect\cV)}{\Trel\sigma{\Vect\zeta}(\Vect\cV)} \quad\text{and}\quad \Trel{\phi}{\Vect\zeta}(\Vect\cV)=\Crel{\Trelv\phi{\Vect\zeta}(\Vect\cV)}\,. \end{align*} We are left with the case of recursive definitions of types, so assume that $\phi=\TREC\zeta\psi$. Let $\Vect\phi=(\List\phi 1n)$ be a list of closed positive types and let $\Vect\cV\in\prod_{i=1}^n\Relsv{\phi_i}$, we set \begin{align} \label{eq:bi-rel-fixpoint-def} \Trelv\phi{\Vect\zeta}(\Vect\cV)= \Infrel\{\cV\in\Relsv{\Subst\phi{\Vect\phi}{\Vect\zeta}}\mid \FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}(\Vect\cV,\cV)}\Subrel\cV\} \end{align} where we use the following notation: given $\cW\in\Relsv{\Substbis\psi{\Vect\phi/\Vect\zeta,\Subst\phi{\Vect\phi}{\Vect\zeta}/\zeta}}$, $\FOLD{\cW}\in\Relsv{\Subst{\phi}{\Vect\phi}{\Vect\zeta}}$ is given by $\Srel{\FOLD\cW}{\epsilon}=\{(\FOLD W,v)\mid(W,v)\in\Srel\cW\epsilon\}$ for $\epsilon\in\{\POS,\NEG\}$. We recall the statement of Tarski's fixpoint theorem. \begin{theorem} Let $S$ and $T$ be complete meet semi-lattices and let $f:S\times T\to T$ be a monotone function. For $x\in S$, let $g(x)$ be the meet of the set $\{y\in T\mid f(x,y)\leq y\}$. Then the function $g$ is monotone and satisfies $f(x,g(x))=g(x)$ for each $x\in S$. \end{theorem} Applying this theorem we obtain, by induction on types, the following property. \begin{lemma}\label{lemma:bi-rel-monotone} For any type $\sigma$ and any positive type $\phi$, the maps $\Trel\sigma{\Vect\zeta}$ and $\Trelv\phi{\Vect\zeta}$ are monotone wrt.~the $\Subrel$ order relation. If $\psi$ is a positive type, $\Vect\zeta=(\List\zeta 1n,\zeta)$ is a repetition-free list of type variables containing all the free variables of $\psi$ and $\phi=\TREC\zeta\psi$ and $\Vect\cV=(\List\cV 1n)$ is a list of pairs of relations such that $\cV_i\in\Relsv{\phi_i}$ for each $i$, then $\cV=\Trel\phi{\Vect\zeta}(\Vect\cV)$ satisfies $\cV=\FOLD{\Trel{\psi}{\Vect\zeta,\zeta}}(\Vect\cV,\cV)$. \end{lemma} \subsection{Some useful closeness lemmas} We state and prove a series of lemmas expressing that our pairs of relations are closed under various syntactic and semantic operations. \begin{lemma}\label{lemma:prob-conv-indep-det} Let $M$ and $M'$ be terms such that $\TSEQ{}{M}{\ONE}$ and $\TSEQ{}{M'}{\ONE}$. If $M\Rel\Wred M'$ then $\Redmats^\infty_{M,\ONELEM} =\Redmats^\infty_{M',\ONELEM}$. \end{lemma} This is straightforward since any reduction path from $M$ to $\ONELEM$ must start with the step $M\Rel\Wred M'$, and this is a probability $1$ step. \begin{lemma}\label{lemma:rel-app-closeness} Let $\phi$ be a closed positive type and let $\sigma$ be a closed type. Let $(M,u)\in\Srel{\Trel\phi{}}{-\epsilon}$ and $(R,r)\in\Srel{\Trel{\LIMPL\phi\sigma}{}}{\epsilon}$. Then $(\LAPP RM,\Matapp ru)\in\Srel{\Trel{\sigma}{}}{\epsilon}$. \end{lemma} \begin{proof} We can write $\sigma=\LIMPL{\phi_1}{\LIMPL\cdots{\LIMPL{\phi_n}\psi}}$ for some $n$ and $\List\phi 1n,\psi$ positive and closed. Given $(V_i,v_i)\in\Srel{\Trelv{\phi_i}{}}{-\epsilon}$ for $i=1,\dots,n$, we have to prove that \begin{align*} (\LAPP R{M\,V_1\cdots V_n},\Matapp ru\Appsep v_1\cdots v_n) \in\Srel{(\Crel{\Trelv{\psi}{}})}{\epsilon} \end{align*} so let $(T,t)\in\Srel{(\LIMPL{\Trelv\psi{}}{\Trel\ONE{}})}{-\epsilon}$, we have to prove that \begin{align*} (\LAPP{T}{(\LAPP R{M\,V_1\cdots V_n})},t(\Matapp ru\Appsep v_1\cdots v_n)) \in\Trel{\ONE}{}\,. \end{align*} % Let $S=\ABST{x}{\phi}{\LAPP T{(\LAPP R{x\,V_1\cdots V_n})}}$ so that $\TSEQ{}{S}{\LIMPL\phi\ONE}$. Similarly let $s\in\Pcoh{\Tsem{\LIMPL\phi\ONE}}$ be the linear morphism defined by $\Matapp s{u'}=t(\Matapp r{u'}\Appsep v_1\cdots v_n)$ (the fact that $s$ so defined is actually a morphism in $\PCOH$ results from the symmetric monoidal closeness of that category and from the fact that $r$ and $t$ are morphisms in $\PCOH$). % Let $(V,v)\in\Srel{\Trelv\phi{}}{-\epsilon}$, we have $(\LAPP RV,\Matapp rv)\in\Srel{\Trel\sigma{}}{\epsilon}$ and hence $(\LAPP R{V\Appsep V_1\cdots V_n}),\Matapp rv\Appsep v_1\cdots v_n\in\Srel{\Trel{\psi}{}}{\epsilon}$. Therefore \begin{align*} (\LAPP T{(\LAPP R{V\Appsep V_1\cdots V_n})}),t(\Matapp rv\Appsep v_1\cdots v_n)\in\Trel{\ONE}{} \end{align*} since we have assumed that $(T,t)\in\Srel{(\LIMPL{\Trelv\psi{}}{\Trel\ONE{}})}{-\epsilon}$. Since $t(\Matapp rv\Appsep v_1\cdots v_n)=\Matapp sv$, and by Lemma~\ref{lemma:prob-conv-indep-det}, it follows that $(\LAPP SV,\Matapp sv)\in\Trel\ONE{}$. Hence $(S,s)\in\Srel{\Trel{\LIMPL\phi\ONE}{}}{\epsilon}$ and therefore $(\LAPP SM,\Matapp su)\in\Trel\ONE{}$ since we have $(M,u)\in\Srel{\Trel\phi{}}{-\epsilon}$. We finish the proof by observing that $\Matapp su=t(\Matapp ru\Appsep v_1\cdots v_n)$ and by showing that \begin{align*} \Redmats^\infty_{\LAPP T{\LAPP R{M\,V_1\cdots V_n}},\ONELEM} =\Redmats^\infty_{\LAPP SM,\ONELEM} \end{align*} For this it suffices to observe (by inspection of the reduction rules) that each reduction path $\pi$ from $\LAPP T{(\LAPP R{M\,V_1\cdots V_n})}$ to $\ONELEM$ is of shape $\pi=\lambda\rho$ where \begin{itemize} \item $\lambda$ is a reduction path \[\LAPP T{(\LAPP R{M_1\,V_1\cdots V_n})}\Rel{\Redone{p_1}}\LAPP T{(\LAPP R{M_2\,V_1\cdots V_n})}\Rel{\Redone{p_2}}\cdots\Rel{\Redone{p_k}}\LAPP T{(\LAPP R{M_{k+1}\,V_1\cdots V_n})}\] where $M_1=M$, $M_{k+1}$ is a value $V$ and $M=M_1\Rel{\Redone{p_1}}M_2\Rel{\Redone{p_2}}\cdots\Rel{\Redone{p_{k}}} M_{k+1}=V$ \item and $\rho$ is a reduction path from $\LAPP T{(\LAPP R{V\,V_1\cdots V_n})}$ to $\ONELEM$. \end{itemize} Then we have $\LAPP SM=\LAPP S{M_1}\Rel{\Redone{p_1}}\LAPP S{M_2}\Rel{\Redone{p_2}}\cdots\Rel{\Redone{p_{k}}}\LAPP SV\Rel{\Redone 1} \LAPP T{(\LAPP R{V\,V_1\cdots V_n})}$, the last step resulting from the definition of $S$. In that way, we have defined a probability preserving bijection between the reduction paths from $\LAPP T{(\LAPP R{M_1\,V_1\cdots V_n})}$ to $\ONELEM$ and the reduction paths from $\LAPP SM$ to $\ONELEM$, proving our contention. \end{proof} \begin{lemma}\label{lemma:tens-crel} Let $\phi_i$ be closed positive types and $(M_i,u_i)\in\Srel{\Trel{\phi_i}{}}{\epsilon}$ for $i\in\{\ell,r\}$. Then $(\PAIR{M_\ell}{M_r},\TENS{u_\ell}{u_r}) \in\Srel{\Trel{\TENS{\phi_\ell}{\phi_r}}{}}{\epsilon}$. \end{lemma} \begin{proof} Let $(T,t)\in\Srel{(\Limpl{\Trelv{\TENS{\phi_\ell}{\phi_r}}{}} {\Trel{\ONE}{}})}{-\epsilon}$, we must prove that $(\LAPP T{\PAIR{M_\ell}{M_r}},t(\Tens{u_\ell}{u_r}))\in\Trel{\ONE}{}$. Let $S=\ABST{x_\ell}{\phi_\ell}{\ABST{x_r}{\phi_r}{\LAPP T{\PAIR{x_\ell}{x_r}}}}$ and $s\in\Pcoh{\Tsem{\LIMPL{\phi_\ell}{(\LIMPL{\phi_r}{\ONE})}}}$ be defined by $\Matapp s{u_\ell}\Appsep u_r=t(\Tens{u_\ell}{u_r})$ (again, $s$ is a morphism in $\PCOH$ by symmetric monoidal closeness of that category). It is clear that $(S,s)\in\Srel{(\LIMPL{\Trelv{\phi_\ell}{}} {(\LIMPL{\Trelv{\phi_r}{}}{\Trel{\ONE}{}}{})}{})}{-\epsilon}$. By Lemma~\ref{lemma:rel-app-closeness} we get $(\LAPP S{M_\ell},\Matapp s{u_\ell})\in\Srel{(\LIMPL{\Trelv{\phi_r}{}}{\Trel{\ONE}{}})}{-\epsilon}$ and then $(\LAPP{S}{M_\ell\,M_r},t(\Tens{u_\ell}{u_r})\in\Trel{\ONE}{}$. Observing that there is a probability preserving bijection between the reduction paths from $\LAPP{S}{M_\ell\,M_r}$ to $\ONELEM$ and the reduction paths from $\LAPP T{\PAIR{M_\ell}{M_r}}$ to $\ONELEM$, we conclude that $(\LAPP T{\PAIR{M_\ell}{M_r}},t(\Tens{u_\ell}{u_r}))\in\Trel{\ONE}{}$ as contended (in both terms one has to reduce first $M_\ell$ and then $M_r$ to a value). \end{proof} \begin{lemma}\label{lemma:proj-crel} Let $\phi_\ell$ and $\phi_r$ be closed positive types. If $(M,u)\in\Srel{\Trel{\TENS{\phi_\ell}{\phi_r}}{}}\epsilon$ then $(\PR iM,\Matapp{\Proj i}u)\in\Srel{\Trel{\phi_i}{}}\epsilon$. \end{lemma} \begin{proof} Let $(T,t)\in\Srel{(\Limpl{\Trelv{\phi_i}{}}{\Trel\ONE{}})}{-\epsilon}$, we have to prove that $(\LAPP T{\PR iM},\Matapp t{(\Matapp{\Proj i}u)})\in\Trel\ONE{}$. Let $S=\ABST x{\TENS{\phi_\ell}{\phi_r}}{\LAPP T{\PR ix}}$ and $s\in\Pcoh{\Tsem{\LIMPL{\TENS{\phi_\ell}{\phi_r}}\ONE}}$ be defined by $\Matapp s{u_0}=\Matapp t{(\Matapp{\Proj i}{u_0})}$ for all $u_0\in\Pcoh{\Tsem{\TENS{\phi_\ell}{\phi_r}}}$. Let $(W,w)\in\Srel{\Trelv{\TENS{\phi_\ell}{\phi_r}}{}}\epsilon$, which means that $W=\PAIR{V_\ell}{V_r}$ and $w=\Tens{v_\ell}{v_r}$ with $(V_j,v_j)\in\Srel{\Trelv{\phi_j}{}}\epsilon$ for $j\in\{\ell,r\}$. We have $\LAPP SW\Rel\Wred\LAPP T{V_i}$ and $\Matapp sw=\Matapp t{v_i}$ and we know that $(\LAPP T{V_i},\Matapp t{v_i})\in\Trel\ONE{}$ from which it follows by Lemma~\ref{lemma:prob-conv-indep-det} that $(\LAPP SW,\Matapp sw)\in\Trel{\ONE}{}$. So we have proven that $(S,s)\in\Srel{(\Limpl{\Trelv{\TENS{\phi_\ell}{\phi_r}}{}} {\Trel\ONE{}})}{-\epsilon}$ and hence $(\LAPP SM,\Matapp su)\in\Trel{\ONE}{}$. We have $\Matapp su=\Matapp t{(\Matapp{\Proj i}u)}$. Moreover we have a probability preserving bijection between the reduction paths from $\LAPP T{\PR iM}$ to $\ONELEM$ and the reduction paths from $\LAPP SM$ to $\ONELEM$, and hence we have $(\LAPP T{\PR iM},\Matapp t{(\Matapp{\Proj i}u)})\in\Trel\ONE{}$ as contended. Indeed, any reduction path $\pi$ from $\LAPP T{\PR iM}$ to $\ONELEM$ has shape $\pi=\lambda\rho$ where $\lambda$ is a reduction path $\LAPP T{\PR iM}=\LAPP T{\PR i{M_\ell}}\Rel{\Redone{p_1}}\LAPP T{\PR i{M_r}}\Rel{\Redone{p_2}}\cdots\Rel{\Redone{p_k}}\LAPP T{\PR i{W}}\Rel{\Redone 1}\LAPP T{V_i}$ (with $W=\PAIR{V_\ell}{V_r}$) and $\rho$ is a reduction path from $\LAPP T{V_i}$ to $\ONELEM$. The first steps $\lambda$ of this reduction are determined by the reduction path $M=M_\ell\Rel{\Redone{p_1}}\cdots\Rel{\Redone{p_k}}W$ from $M$ to the value $W$. This reduction path determines uniquely the reduction path $\LAPP SM=\LAPP S{M_\ell}\Rel{\Redone{p_1}}\cdots\Rel{\Redone{p_k}}\LAPP SW\Rel{\Redone 1}\LAPP T{\PR iW}\Rel{\Redone 1}\LAPP T{V_i}$ followed by the reduction $\rho$ from $\LAPP T{V_i}$ to $\ONELEM$ by $\rho$. \end{proof} \begin{lemma}\label{lemma:plus-crel} Let $\phi_\ell$ and $\phi_r$ be closed positive types and let $(M,u)\in\Srel{\Trel{\phi_i}{}}{\epsilon}$ for $i=\ell$ or $i=r$. Then $(\IN i{M},\Matapp{\Inj i}u)\in\Srel{\Trel{\PLUS{\phi_\ell}{\phi_r}}{}}{\epsilon}$. \end{lemma} \begin{proof} Let $(T,t)\in\Srel{(\Limpl{\Trelv{\PLUS{\phi_\ell}{\phi_r}}{}} {\Trel{\ONE}{}})}{-\epsilon}$, we must prove that $(\LAPP T{\IN i{M}},\Matapp t{(\Matapp{\Inj i}u)})\in\Trel{\ONE}{}$. Let $S=\ABST x{\phi_i}{\LAPP t{\IN i(x)}}$ and let $s\in\Pcoh{\Tsem{\LIMPL{\phi_i}{\ONE}}}$. It is clear that $(S,s)\in\Srel{(\Limpl{\Trelv{\phi_i}{}}{\Trel{\ONE}{}}{})}{-\epsilon}$ and it follows that $(\LAPP S{M},\Matapp su)\in\Trel{\ONE}{}$ which implies $(\LAPP T{\IN i{M}},\Matapp t{(\Matapp{\Inj i}u)})\in\Trel{\ONE}{}$ by the usual bijective and probability preserving bijection on reductions. \end{proof} The next lemma uses notations introduced in Section~\ref{sec:sem-additives}. \begin{lemma}\label{lemma:case-crel} Let $\phi_\ell$ and $\phi_r$ be closed positive type and $\sigma$ be a closed type. Let $(M,u)\in\Srel{\Trel{\PLUS{\phi_\ell}{\phi_r}}{}}\epsilon$. For $i\in\{\ell,r\}$, let $R_i$ be a term such that $\TSEQ{y_i:\phi_i}{R_i}{\sigma}$ and assume that $(\ABST{x_i}{\phi_i}{R_i},r_i) \in\Srel{\Trel{\LIMPL{\phi_i}{\sigma}}{}}{-\epsilon}$. Then $(\CASE{M}{y_\ell}{R_\ell}{y_r}{R_r}, \Matapp{\Case(r_\ell,r_r)}u)\in\Srel{\Trel{\sigma}{}}{-\epsilon}$. \end{lemma} \begin{proof} We can write $\sigma=\LIMPL{\psi_1}{\LIMPL\cdots{\LIMPL{\psi_k}{\psi}}}$ where $\psi$ and the $\psi_j$'s are closed and positive types. Given $(W_j,w_j)\in\Srel{\Trelv{\psi_j}{}}{\epsilon}$ for $j=1,\dots,k$, we have to prove that \begin{align}\label{eq:adeq-case-test1} (\LAPP{\CASE{M}{y_\ell}{R_\ell}{y_r}{R_r}}{\Vect W}, \Matapp{\Case(r_\ell,r_r)}u\Appsep\Vect w) \in\Srel{\Trelv{\psi}{}}{-\epsilon} \end{align} so let $(T,t)\in\Srel{(\Limpl{\Trelv{\psi}{}}{\Trel\ONE{}})}{\epsilon}$, our goal is to prove that \begin{align}\label{eq:adeq-case-test2} (\LAPP T{\LAPP{\CASE{M}{y_\ell}{R_\ell}{y_r} {R_r}}{\Vect W}}, \Matapp t{(\Matapp{\Case(r_\ell,r_r)}u\Appsep\Vect w)})\in\Trel\ONE{}\,. \end{align} Let $S=\ABST x{\PLUS{\phi_\ell}{\phi_r}}{\LAPP T{\LAPP{\CASE{x}{y_\ell}{R_\ell}{y_r} {R_r}}{\Vect W}}}$ and $s\in\Pcoh{\Tsem{\LIMPL{\PLUS{\phi_\ell}{\phi_r}}{\ONE}}}$ be defined by $\Matapp s{u_0}=\Matapp t{(\Matapp{\Case(r_\ell,r_r)}{u_0}\Appsep\Vect w)}$ for each $u_0\in\Pcoh{\Tsem{\PLUS{\phi_\ell}{\phi_r}}}$. Then we have $(S,s)\in\Srel{(\Limpl{\Trelv{\PLUS{\phi_\ell}{\phi_r}}{}} {\Trel\ONE{}})}{\epsilon}$. Let indeed $i\in\{\ell,r\}$ and let $(V,v)\in\Srel{\Trelv{\phi_i}{}}{-\epsilon}$ so that $(\IN iV,\Matapp{\Inj i}v)\in\Srel{\Trelv{\PLUS{\phi_\ell}{\phi_r}}{}}{-\epsilon}$. We have $\LAPP S{\IN iV}\Rel{\Transcl{\Wred}}\LAPP{T} {\LAPP{\Subst{R_i}{V}{y_i}}{\Vect W}}$ and $\Matapp s{(\Matapp{\Inj i}v)}=\Matapp t{(\Matapp{r_i}v\Appsep\Vect w)}$ and, by our assumptions and Lemma~\ref{lemma:prob-conv-indep-det}, $(\Subst{R_i}{V}{y_i},\Matapp{r_i}v)\in\Srel{\Trel{\sigma}{}}{-\epsilon}$ and hence $(\LAPP{\Subst{R_i}{V}{y_i}}{\Vect W},\Matapp{r_i}v\Appsep\Vect w)\in\Srel{\Trelv{\psi}{}}{-\epsilon}$. By Lemma~\ref{lemma:prob-conv-indep-det} it follows that $(\LAPP S{\IN iV},\Matapp s{(\Matapp{\Inj i}v)})\in\Trel\ONE{}$ and hence $(S,s)\in\Srel{(\Limpl{\Trelv{\PLUS{\phi_\ell}{\phi_r}}{}} {\Trel\ONE{}})}{\epsilon}$ as contended. Therefore $(\LAPP SM,\Matapp su)\in\Trel\ONE{}$. There is a bijective and probability preserving correspondence between the reductions from $\LAPP SM$ to $\ONELEM$ and the reductions from $\LAPP T{\LAPP{\CASE{M}{x_\ell}{\LAPP{R_\ell}{x_\ell}}{x_r} {\LAPP{R_r}{x_r}}}{\Vect W}}$ to $\ONELEM$: as usual such reductions start with a reduction $M=M_1\Rel{\Redone{p_1}}M_2\Rel{\Redone{p_2}}\cdots\Rel{\Redone{p_k}}M_k=\IN iV$ where $i\in\{\ell,r\}$ and $V$ is a value of type $\phi_i$ and (after a few $\Wred$-steps) continue with a reduction from $\LAPP{T}{\LAPP{R_i}{V\,\Vect W}}$ to $\ONELEM$. Therefore~\Eqref{eq:adeq-case-test2} holds and hence we have~\Eqref{eq:adeq-case-test1}, this ends the proof of the lemma. \end{proof} \begin{lemma}\label{lemma:der-crel} Let $\sigma$ be a closed type and let $(M,u)\in\Srel{\Trel{\EXCL\sigma}{}}\epsilon$. We have $(\GO M,\Matapp{\Der{}}u)\in\Srel{\Trel\sigma{}}\epsilon$. \end{lemma} \begin{proof} We can write $\sigma=\LIMPL{\psi_1}{\LIMPL\cdots{\LIMPL{\psi_k}{\psi}}}$ where $\psi$ and the $\psi_j$'s are closed and positive types. Given $(W_j,w_j)\in\Srel{\Trelv{\psi_i}{}}{-\epsilon}$, we have to prove that \begin{align}\label{eq:adeq-der-test1} (\LAPP{\GO M}{\Vect W}, \Matapp{\Der{}}u\Appsep\Vect w)\in\Srel{\Trelv{\psi}{}}{\epsilon} \end{align} so let $(T,t)\in\Srel{(\Limpl{\Trelv{\psi}{}}{\Trel\ONE{}})}{-\epsilon}$, our goal is to prove that \begin{align}\label{eq:adeq-der-test2} (\LAPP T{\LAPP{\GO M}{\Vect W}}, \Matapp t{(\Matapp{\Der{}}u\Appsep\Vect w)})\in\Trel\ONE{}\,. \end{align} We set $S=\ABST x{\EXCL\sigma}{\LAPP T{\LAPP{\GO x}{\Vect W}}}$ and we define $s\in\Pcoh{\Tsem{\LIMPL{\EXCL\sigma}{\ONE}}}$ by $\Matapp s{u_0}=\Matapp t{(\Matapp{\Der{}}{u_0}\Appsep\Vect w)}$ for all $u_0\in\Pcoh{\Tsem{\EXCL\sigma}}$, and we prove that $(S,s)\in\Srel{(\Limpl{\Trelv{\EXCL\sigma}{}}{\Trel\ONE{}})}{-\epsilon}$ as in the proof of Lemme~\ref{lemma:case-crel} (for instance). We finish the proof in the same way, showing~\Eqref{eq:adeq-der-test2} by establishing a bijective and probability preserving correspondence between reductions. Our contention~\Eqref{eq:adeq-der-test1} follows. \end{proof} \begin{lemma}\label{lemma:unfold-crel} Let $\phi$ be a closed positive type of shape $\phi=\TREC\zeta\psi$. If $(M,u)\in\Srel{\Trel{\phi}{}}\epsilon$ then $(\UNFOLD M,u)\in\Srel{\Trel{\Subst\psi\phi\zeta}{}}\epsilon$. \end{lemma} \begin{proof} Let $(T,t)\in\Srel{(\Limpl{\Trelv{\Subst\psi\phi\zeta}{}}{\Trel\ONE{}})} {-\epsilon}$, we must prove that $(\LAPP T{\UNFOLD M},u)\in\Trel{\ONE}{}$. As usual one defines $S=\ABST x{\phi}{\LAPP T{\UNFOLD x}}$ and one proves that $(S,t)\in\Srel{(\Limpl{\Trelv{\phi}{}}{\Trel\ONE{}})} {-\epsilon}$. This results from the fact that if $(V,v)\in\Srel{\Trelv{\phi}{}}{\epsilon}$ then $V=\FOLD W$ with $(W,v)\in\Srel{\Trelv{\Subst\psi\phi\zeta}{}}{\epsilon}$, from Lemma~\ref{lemma:prob-conv-indep-det} and from the fact that $\UNFOLD{\FOLD W}\Rel\Wred W$ (and of course from our assumption on $(T,t)$). It follows that $(\LAPP SM,\Matapp tu)\in\Trel{\ONE}{}$ from which we deduce $(\LAPP T{\UNFOLD M},u)\in\Trel{\ONE}{}$ by the usual reasoning involving a bijective probability preserving correspondence on reductions. \end{proof} \begin{lemma}\label{lemma:fold-crel} Let $\phi$ be a closed positive type of shape $\phi=\TREC\zeta\psi$. If $(M,u)\in\Srel{\Trel{\Subst\psi\phi\zeta}{}}\epsilon$ then $(\FOLD M,u)\in\Srel{\Trel{\phi}{}}\epsilon$. \end{lemma} \begin{proof} Let $(T,t)\in\Srel{(\Limpl{\Trelv{\phi}{}}{\Trel\ONE{}})} {-\epsilon}$, we must prove that $(\LAPP T{\FOLD M},u)\in\Trel{\ONE}{}$. As usual one defines $S=\ABST x{\Subst\psi\phi\zeta}{\LAPP T{\FOLD x}}$ and one proves that $(S,t)\in\Srel{(\Limpl{\Trelv{\Subst\psi\phi\zeta}{}}{\Trel\ONE{}})} {-\epsilon}$. This results easily from the fact that, if $(V,v)\in\Srel{\Trelv{\Subst\psi\phi\zeta}{}}{\epsilon}$ then $(\FOLD V,v)\in\Srel{\Trelv{\phi}{}}{\epsilon}$, from Lemma~\ref{lemma:prob-conv-indep-det} and from our assumption about $(T,t)$. Therefore we have $(\LAPP SM,\Matapp tu)\in\Trel{\ONE}{})$ from which we deduce $(\LAPP T{\FOLD M},u)\in\Trel{\ONE}{}$ by the usual reasoning. \end{proof} \begin{lemma}\label{lemma:Trel-sem-closeness} Let $\sigma$ be a closed type and let $M$ be a closed term of type $\sigma$. Then $(M,0)\in\Srel{\Trel{\sigma}{}}{\epsilon}$ and, if $D\subseteq\Pcoh{\Tsem\sigma}$ is directed and satisfies $\forall u\in D\ (M,u)\in\Srel{\Trel{\sigma}{}}{\epsilon}$ then $(M,\sup D)\in\Srel{\Trel{\sigma}{}}{\epsilon}$. Last, if $(M,u)\in\Srel{\Trel{\sigma}{}}{\epsilon}$ and $u'\leq u$ then $(M,u')\in\Srel{\Trel{\sigma}{}}{\epsilon}$. \end{lemma} \begin{proof} We can write $\sigma=\LIMPL{\phi_1}{\LIMPL\cdots{\LIMPL{\phi_n}\psi}}$ for some $n$ and $\List\phi 1n,\psi$ positive and closed. Let us prove the second statement. For $i=1,\dots,n$, let $(V_i,v_i)\in\Srel{\Trelv{\phi_i}{}}{-\epsilon}$, we must prove that $(\LAPP M{V_1\cdots V_n},\Matapp{(\sup D)}{v_1}\cdots v_n)\in\Srel{\Crel{\Trelv{\psi}{}}}{\epsilon}$, knowing that \[\forall u\in D\quad (\LAPP M{V_1\cdots V_n},u\Appsep v_1\cdots v_n)\in\Srel{\Crel{\Trelv{\psi}{}}}{\epsilon}\,. \] This results from the fact that, given $t\in\Pcoh{\Tsem{\LIMPL\psi\ONE}}$, the map $u\mapsto \Matapp t{(u\Appsep v_1\cdots v_n)}$ is Scott continuous from $\Pcoh{\Tsem\phi}$ to $[0,1]$. The first statement of the lemma results from the fact that this function maps $0$ to $0$. The last one results from the fact that this function is monotone. \end{proof} \subsection{Uniqueness of the relation} With any closed type $\sigma$ we have associated a pair of relations $\Trel\sigma{}$. We need now to prove that this pair satisfies $\Srel{\Trel\sigma{}}\POS=\Srel{\Trel\sigma{}}\NEG$ so that we have actually associated a unique relation with any type. To this end we prove first that $\Srel{\Trel\sigma{}}\POS\subseteq\Srel{\Trel\sigma{}}\NEG$. Defining, for any pair of relations $\cR$, the relation $\Op\cR$ as $(\Srel\cR\POS,\Srel\cR\NEG)$, this amounts to proving that ${\Trel\sigma{}}\Subrel\Op{\Trel\sigma{}}$. We use the same notation for the elements of $\Relsv\phi$ for $\phi$ positive. For the next lemma, we use the same notational conventions as above. \begin{lemma} Let $\Vect\cV$ be a list of pairs of relations such that $\cV_i\in\Relsv{\phi_i}$ and $\cV_i\Subrel\Op{\cV_i}$ for each $i$. Then $\Trel\sigma{\Vect\zeta}(\Vect\cV) \Subrel\Op{\Trel\sigma{\Vect\zeta}(\Vect\cV)}$ and $\Trelv\phi{\Vect\zeta}(\Vect\cV) \Subrel\Op{\Trelv\phi{\Vect\zeta}(\Vect\cV)}$. \end{lemma} \begin{proof} The proof is by induction on types. All cases result straightforwardly from the monotony of the logical operations on pairs of relations, but the case of fixpoints of types. So assume that $\phi=\TREC\zeta\psi$, let $\cV=\Trelv\phi{\Vect\zeta}(\Vect\cV)$ and let us prove that $\cV \Subrel\Op\cV$. For this, because of the definition of $\cV$ as a glb~\Eqref{eq:bi-rel-fixpoint-def}, it suffices to show that \begin{align*} \FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}(\Vect\cV,\Op\cV)}\Subrel\Op\cV\,. \end{align*} By the first statement of Lemma~\ref{lemma:bi-rel-monotone} and our assumption on the $\cV_i$'s we have \[ \FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}({\Vect\cV},\Op\cV)} \Subrel\FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}(\Op{\Vect\cV},\Op\cV)}\,. \] By inductive hypothesis we have $\FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}(\Op{\Vect\cV},\Op\cV)} \Subrel\Op{\FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}({\Vect\cV},\cV)}}=\Op\cV$ since $\cV=\FOLD{\Trelv{\psi}{\Vect\zeta,\zeta}({\Vect\cV},\cV)}$ by Lemma~\ref{lemma:bi-rel-monotone}. \end{proof} We are left with proving the converse property, namely that $\Op{\Trel\sigma{}}\Subrel\Trel\sigma{}$ for each closed type $\sigma$. This requires a bit more work, and is based on a notion of ``finite'' approximation of elements of the model, that we define by syntactic means as follows. \subsubsection{Restriction operators} We define\footnote{This definition as well as our reasoning below features some similarities with \emph{step-indexing} that we would like to understand better.} closed terms $\Restr n\sigma$ and $\Restrv n\phi$ (for $n\in\Nat$, $\sigma$ a type and $\phi$ a positive type) typed as follows: $\TSEQ{}{\Restr n\sigma}{\LIMPL{\EXCL\sigma}{\sigma}}$ and $\TSEQ{}{\Restrv n\phi}{\LIMPL{\phi}{\phi}}$. \begin{align*} \Restr n\phi &= \ABST x{\Excl\phi}{\LAPP{\Restrv n\phi}{\GO x}} \\ \Restr n{\LIMPL\phi\sigma} &= \ABST f{\EXCL{(\LIMPL\phi\sigma)}} {\ABST{x}{\phi}{\LAPP{\Restr{n}{\sigma}} {\STOP{(\LAPP{\GO f}{\LAPP{\Restrv n\phi}{x}})}}}}\\ \Restrv n\ONE &= \ABST x\ONE x\\ \Restrv{n}{\EXCL\sigma} &= \ABST{x}{\EXCL\sigma}{\STOP{(\LAPP{\Restr n\sigma}{x})}}\\ \Restrv{n}{\TENS{\phi_\ell}{\phi_r}} &= \ABST x{\TENS{\phi_\ell}{\phi_r}}{\PAIR{\LAPP{\Restrv n{\phi_\ell}}{\PR \ell x}}{\LAPP{\Restrv n{\phi_r}}{\PR r x}}}\\ \Restrv n{\PLUS{\phi_\ell}{\phi_r}} &= \ABST x{\PLUS{\phi_\ell}{\phi_r}}{\CASE{x}{x_\ell} {\IN \ell {\LAPP{\Restrv n{\phi_\ell}}{x_\ell}}}{x_r} {\IN r {\LAPP{\Restrv n{\phi_r}}{x_r}}}}\\ \Restrv 0{\TREC\zeta\phi} &= \ABST x{\TREC\zeta\phi}{\LOOP{\TREC\zeta\phi}}\\ \Restrv{n+1}{\TREC\zeta\phi} &= \ABST{x}{\TREC\zeta\phi}{\FOLD{\LAPP{\Restrv n{\Subst\phi{\TREC\zeta\phi}\zeta}}{\UNFOLD x}}} \end{align*} This is a well-founded lexicographic inductive definition on triples $(n,\sigma,l)$ (where $l\in\{\Vsymb,\Gsymb\}$) if we order the symbols $\Vsymb$ and $\Gsymb$ by $\Vsymb<\Gsymb$. We consider actually only triples $(n,\sigma,l)$ such that $\sigma$ is positive if $l=\Vsymb$. We describe similarly the interpretation of these terms: we give an explicit description of the matrices $\Psem{\Restr n\sigma}{}$ and $\Psem{\Restrv n\phi}{}$. To this end, we define a family of sets $\Srestrp n\sigma l\subseteq\Web{\Tsem\sigma}$ by induction on $(n,\sigma,l)$ (where $n\in\Nat$, $l\in\{\Vsymb,\Gsymb\}$ and $\sigma$ is a closed type which is positive if $l=\Vsymb$). \begin{itemize} \item $\Srestrv n{\Excl\sigma}=\Mfin{\Srestr n\sigma}$ \item $\Srestrv n{\TENS{\phi_\ell}{\phi_r}}=\Srestrv n{\phi_\ell}\times\Srestrv n{\phi_r}$ \item $\Srestrv n{\PLUS{\phi_\ell}{\phi_r}}= \{\ell\}\times\Srestrv n{\phi_\ell}\cup\{r\}\times\Srestrv n{\phi_r}$ \item $\Srestrv 0{\TREC\zeta\psi}=\emptyset$ \item $\Srestrv{n+1}{\TREC\zeta\psi}=\Srestrv{n}{\Subst\phi{\TREC\zeta\psi}\zeta}$ \item $\Srestr n\phi=\Srestrv n\phi$ \item $\Srestr n{\LIMPL\phi\sigma}=\Srestrv n\phi\times\Srestr n\sigma$. \end{itemize} \begin{lemma}\label{lemma:restr-expression} Let $n\in\Nat$, $\phi$ be a closed positive type and $\sigma$ be a closed type. One has \begin{align*} \Psem{\Restrv n\phi}{}_{(a,b)}= \begin{cases} 1 & \text{if }a=b\in\Srestrv n\phi\\ 0 & \text{otherwise.} \end{cases} \quad \Psem{\Restr n\sigma}{}_{(c,b)}= \begin{cases} 1 & \text{if }c=\Mset b\text{ and }b\in\Srestr n\sigma\\ 0 & \text{otherwise.} \end{cases} \end{align*} \end{lemma} \begin{proof} By Theorem~\ref{th:pos-types-dense}, for a closed positive type $\phi$ and for $u\in\PcohEM{\Tsemca\phi}$, it suffices to prove that \begin{align*} \Psem{\Restrv n\phi}(u)_a= \begin{cases} u_a & \text{if }a\in\Srestrv n\phi\\ 0 & \text{otherwise} \end{cases} \end{align*} And for a closed type $\sigma$ and for $u\in\Pcoh{\Tsem\sigma}$, it suffices to prove \begin{align*} \Psem{\Restr n\sigma}(\Prom u)_a= \begin{cases} u_a & \text{if }a\in\Srestr n\sigma\\ 0 & \text{otherwise} \end{cases} \end{align*} Both statements are easily proved by induction. \end{proof} We need now to prove that, given $u\in\Pcoh{\Tsem{\sigma}}$, the sequence $\Psem{\Restr n\sigma}{(\Prom u)}$ is monotone and has $u$ as lub. \begin{lemma}\label{lemma:restr-monotone-limit} For any triple $(n,\sigma,l)$ where $\sigma$ is positive if $l=\Vsymb$, one has $\Srestrp n\sigma l\subseteq\Srestrp{n+1}\sigma l$. Moreover \begin{align*} \bigcup_{n=0}^\infty\Srestrp n\sigma l=\Web{\Tsem\sigma}\,. \end{align*} \end{lemma} \begin{proof} The first statement is straightforward, by induction on $(n,\sigma,l)$. For the second statement, we only have to prove the right-to-left inclusions. We define the \emph{height} $\Height a$ of an element $a$ of $\Web{\Tsem\sigma}$ as follows. \begin{itemize} \item $\Height\Onelem=1$ \item $\Height{a_1,a_2}=1+\max{(\Height{a_1},\Height{a_2})}$ (this definition is used when $\phi$ is a tensor and when $\sigma$ is a linear implication) \item $\Height{i,a}=1+\Height a$ \item $\Height{\Mset{\List a1k}}=1+\max{(\Height{a_1},\dots,\Height{a_k})}$ \end{itemize} Then by induction on $\Height a$ one proves that \begin{align*} \forall a\in\Web{\Tsem\sigma}\,\exists n\in\Nat\ a\in\Srestrp n\sigma l \end{align*} We deal only with the statement relative to $\Srestrv n\phi$. The closed positive type $\phi$ is of shape \begin{align*} \phi=\TREC{\zeta_1}{\cdots\TREC{\zeta_k}{\psi}} \end{align*} where $\psi$ is not of shape $\TREC\zeta\rho$. We introduce auxiliary closed types $\List \phi 1k$ as follows: \begin{align*} \phi_1&=\phi=\TREC{\zeta_1}{\cdots\TREC{\zeta_k}{\psi}}\\ \phi_2&= \TREC{\zeta_2}{\cdots\TREC{\zeta_k}{\Subst\psi{\phi_1}{\zeta_1}}}\\ &\ \ \vdots\\ \phi_{k+1}&= \Substbis\psi{\phi_1/\zeta_1,\phi_2/\zeta_2,\dots,\phi_k/\zeta_k} \end{align*} and all these types have the same interpretation in $\EM\PCOH$. The type $\psi$ cannot be one of the type variables $\zeta_i$ as otherwise we would have $\Web{\Tsem\phi}=\emptyset$, contradicting our assumption that $a$ belongs to this set. Assume that $\psi=\EXCL\sigma$ so that we must have $a=\Mset{\List b1l}$ with $b_i\in\Web{\Tsem{\sigma'}}$ (where $\sigma'=\Substbis\sigma{\phi_1/\zeta_1,\phi_2/\zeta_2,\dots,\phi_k/\zeta_k}$) for each $i=1,\dots,l$. We have $\Height{b_i}<\Height a$ for each $i$ so that we can apply the inductive hypothesis: for each $i$ there is $n_i$ such that $b_i\in\Srestr{n_i}{\sigma'}$. Using the monotonicity property (first statement of the lemma) and setting $n=\max(\List n1l)$ we have $b_i\in\Srestr n{\sigma'}$ and hence $a\in\Srestrv n{\Excl{\sigma'}}$. Therefore $a\in\Srestrv {n+k}{\phi}$ (coming back to the definition of this set). The other cases are dealt with similarly. \end{proof} \begin{lemma}\label{lemma:restr-sup-id} Let $\sigma$ be a closed type and let $\phi$ be a closed positive type. If $u\in\Pcoh{\Tsem\sigma}$ then the sequence $(\Matapp{\Psem{\Restr n\sigma}{}}{\Prom u})_{n\in\Nat}$ is monotone (in $\Pcoh{\Tsem\sigma}$) and has $u$ as lub. If $u\in\Pcoh{\Tsem\phi}$ then the sequence $(\Matapp{\Psem{\Restrv n\phi}{}}u)_{n\in\Nat}$ is monotone and has $u$ as lub. \end{lemma} \begin{proof} Immediate consequence of Lemmas~\ref{lemma:restr-expression} and~\ref{lemma:restr-monotone-limit}. \end{proof} \subsubsection{Main Inclusion Lemma} Now we are in position of proving the key lemma in the proof of the uniqueness of relations. \begin{lemma}\label{lemma:restr-minus-sub-plus} Let $\sigma$ be a closed type, $n\in\Nat$ and $l\in\{\Vsymb,\Gsymb\}$. If $l=\Gsymb$ and $(M,u)\in\Nrel{\Trel\sigma{}}$ then $(M,\Matapp{\Psem{\Restr n\sigma}{}}{\Prom u})\in\Prel{\Trel\sigma{}}$. If $l=\Vsymb$ and $\sigma$ is a closed positive type $\phi$ then $(V,v)\in\Nrel{\Trelv{\phi}{}}\Rightarrow(V,\Matapp{\Psem{\Restrv n\phi}{}}v)\in\Prel{(\Crel{\Trelv{\phi}{}})}=\Prel{\Trel{\phi}{}}$. \end{lemma} \begin{proof} By lexicographic induction on triples $(n,\sigma,l)$ (with $\sigma$ positive when $l=\Vsymb$). Until further notice, we assume that $l=\Vsymb$. The only case where ``$n$ decreases'' in this induction is when $\phi=\TREC\zeta\psi$, we start with this case. Assume that $\phi=\TREC\zeta\psi$ and that $(V,v)\in\Nrel{\Trelv{\phi}{}}$. If $n=0$ we have $\Matapp{\Psem{\Restrv n\phi}{}}v=0$ and hence $(V,\Matapp{\Psem{\Restrv n\phi}{}}v)\in\Prel{\Trel{\phi}{}}$ by Lemma~\ref{lemma:Trel-sem-closeness}. Assume that the implication holds for $n$ and let us prove it for $n+1$. Let $(V,v)\in\Nrel{\Trelv{\phi}{}}$, that is $V=\FOLD W$ with $(W,v)\in\Nrel{\Trelv{\Subst\psi\phi\zeta}{}}$. We have $\Matapp{\Psem{\Restrv{n+1}{\TREC\zeta\psi}}{}}v= \Matapp{\Psem{\Restrv{n}{\Subst\psi\phi\zeta}}{}}v$ by definition. By inductive hypothesis we have \begin{align}\label{eq:hyprec-unfold} (W,\Matapp{\Psem{\Restrv{n}{\Subst\psi\phi\zeta}}{}}v) \in\Prel{\Trel{\Subst\psi\phi\zeta}{}} \end{align} and we must prove that $(\FOLD W,\Matapp{\Psem{\Restrv{n}{\Subst\psi\phi\zeta}}{}}v) \in\Prel{\Trel{\phi}{}}$. Let $(T,t)\in\Nrel{(\LIMPL{\Trelv\phi{}}{\Trel\ONE{}})}$, we must prove that $(\LAPP T{\FOLD W},t(\Matapp{\Psem{\Restrv{n}{\Subst\psi\phi\zeta}}{}}v))\in\Trel\ONE{}$. Let $S=\ABST x{\Subst\psi\phi\zeta}{\LAPP T{\FOLD x}}$, we have $(S,t)\in\Nrel{(\LIMPL{\Trelv{\Subst\psi\phi\zeta}{}}{\Trel\ONE{}})}$ by Lemma~\ref{lemma:prob-conv-indep-det} and therefore \[ (\LAPP SW,t(\Matapp{\Psem{\Restrv{n}{\Subst\psi\phi\zeta}}{}}v))\in\Trel\ONE{} \] by~\Eqref{eq:hyprec-unfold} and Lemma~\ref{lemma:rel-app-closeness} and this implies $(\LAPP T{\FOLD W},\Matapp t{(\Matapp{\Psem{\Restrv{n}{\Subst\psi\phi\zeta}}{}}v)}) \in\Trel\ONE{}$ by Lemma~\ref{lemma:prob-conv-indep-det}. Assume that $\phi=\EXCL\sigma$ and that $(V,v)\in\Nrel{\Trelv{\EXCL\sigma}{}}$, that is $V=\STOP M$ and $v=\Prom u$ with $(M,u)\in\Nrel{\Trel\sigma{}}$. By inductive hypothesis we have $(M,\Matapp{\Psem{\Restr{n}{\sigma}}}{\Prom u}) \in\Prel{\Trel{\sigma}{}}$ and hence $(\STOP M,\Prom{(\Matapp{\Psem{\Restr{n}{\sigma}}}{\Prom u})})\in\Prel{\Trelv{\EXCL\sigma}{}}$ and since $\Prom{(\Matapp{\Psem{\Restr{n}{\sigma}}}{\Prom u})}=\Matapp{\Psem{\Restrv{n}{\EXCL\sigma}}}{\Prom u}$ we get \[ (V,\Matapp{\Psem{\Restrv{n}{\EXCL\sigma}}}v) \in\Prel{\Trelv{\EXCL\sigma}{}}\subseteq\Prel{\Trel{\EXCL\sigma}{}} \] as expected. Assume that $\phi=\TENS{\phi_\ell}{\phi_r}$ and that $(V,v)\in\Nrel{\Trelv{\TENS{\phi_\ell}{\phi_r}}{}}$, that is $V=\PAIR{V_\ell}{V_r}$ and $v=\Tens{v_\ell}{v_r}$ with $(V_i,v_i)\in\Nrel{\Trelv{\phi_i}{}}$ for $i\in\{\ell,r\}$. By inductive hypothesis we have $(V_i,\Matapp{\Psem{\Restr{n}{\phi_i}}}{v_i})\in\Prel{\Trel{\phi_i}{}}$. By Lemma~\ref{lemma:tens-crel} we get $(\PAIR{V_\ell}{V_r},\Tens{\Matapp{\Psem{\Restr{n}{\phi_\ell}}} {v_\ell}}{\Matapp{\Psem{\Restr{n}{\phi_r}}}{v_r}}) \in\Prel{\Trel{\TENS{\phi_\ell}{\phi_r}}{}}$, that is $(\PAIR{V_\ell}{V_r},\Psem{\Restr{n}{\TENS{\phi_\ell}{\phi_r}}}{}(\Tens{v_\ell}{v_r})) \in\Prel{\Trel{\TENS{\phi_\ell}{\phi_r}}{}}$. Assume that $\phi=\PLUS{\phi_\ell}{\phi_r}$ and $(V,v)\in\Nrel{\Trelv{\PLUS{\phi_\ell}{\phi_r}}{}}$. This means that for some $i\in\{\ell,r\}$, one has $V=\IN i{W}$ and $v=\Matapp{\Inj i}w$ for $(W,w)\in\Nrel{\Trelv{\phi_i}{}}$. By inductive hypothesis we have $(W,\Matapp{\Psem{\Restr{n}{\phi_i}}{}}w)\in\Prel{\Trel{\phi_i}{}}$ and hence $(\IN i{W},\Matapp{\Inj i}{(\Matapp{\Psem{\Restr{n}{\phi_i}}{}}w)}) \in\Prel{\Trel{\PLUS{\phi_\ell}{\phi_r}}{}}$ by Lemma~\ref{lemma:plus-crel}, that is $(\IN i{W}, \Matapp{\Psem{\Restr{n}{\PLUS{\phi_\ell}{\phi_r}}}{}}w) \in\Prel{\Trel{\PLUS{\phi_\ell}{\phi_r}}{}}$. We assume now that $l=\Gsymb$. If $\sigma$ is a closed positive type $\phi$ and let $(M,u)\in\Nrel{\Trel{\sigma}{}}$, we must prove that \[ (M,\Matapp{\Psem{\Restr{n}{\sigma}}{}}{\Prom u})\in\Prel{\Trel\sigma{}} \] which follows directly from the definition of $\Restr{n}{\phi}$ and from the inductive hypothesis applied to $(n,\phi,\Vsymb)$. Assume last that $\sigma=\LIMPL\phi\tau$ and that $(M,u)\in\Nrel{\Trel{\LIMPL\phi\tau}{}}$, we must prove that $(M,\Matapp{\Psem{\Restr{n}{\LIMPL\phi\tau}}{}}{\Prom u})\in\Prel{\Trel{\IMPL\phi\tau}{}}$. Let $(V,v)\in\Nrel{\Trelv{\phi}{}}$, we must prove that \begin{align}\label{eq:goal-app} (\LAPP MV,\Matapp{\Psem{\Restr{n}{\LIMPL\phi\tau}}{}}{\Prom u}\Appsep v) \in\Prel{\Trel\tau{}} \end{align} which follows from the fact that $\Matapp{\Psem{\Restr{n}{\LIMPL\phi\tau}}{}}{\Prom u}\Appsep v= \Psem{\Restr{n}{\tau}}{\Prom{(u(\Matapp{\Psem{\Restrv{n}{\phi}}{}}v))}}$. Indeed the inductive hypothesis applied to $(n,\phi)$ yields $(V,\Matapp{\Psem{\Restrv n\phi}{}}v)\in\Prel{\Trel{\phi}{}}$ and hence $(\LAPP MV,\Matapp u{(\Matapp{\Psem{\Restrv n\phi}{}}v)}) \in\Prel{\Trel{\tau}{}}$ by Lemma~\ref{lemma:rel-app-closeness}, from which we derive~\Eqref{eq:goal-app} by Lemma~\ref{lemma:restr-sup-id} and Lemma~\ref{lemma:Trel-sem-closeness}. \end{proof} \begin{lemma} For any closed type $\sigma$ one has $\Nrel{\Trel\sigma{}}=\Prel{\Trel\sigma{}}$. \end{lemma} \begin{proof} Immediate consequence of lemmas~\ref{lemma:Trel-sem-closeness}, \ref{lemma:restr-sup-id} and~\ref{lemma:restr-minus-sub-plus}. \end{proof} From now on we simply use the notation $\Trel\sigma{}$ instead of $\Nrel{\Trel\sigma{}}$ and $\Prel{\Trel\sigma{}}$. \subsection{Logical relation lemma} We can prove now the main result of this section. \begin{theorem}[Logical Relation Lemma]\label{lemma:adequacy} Assume that $\TSEQ{x_1:\phi_1,\dots,x_k:\phi_k}{M}{\sigma}$ and let $(V_i,v_i)\in\Trel{\phi_i}{}$ (where $V_i$ is a value and $v_i\in\PcohEM{\Tsem{\phi_i}}$) for $i=1,\dots,k$. Then $(\Substbis M{V_1/x_1,\dots,V_k/x_k},\Psem{M}^{\List x1k}{\Vect v})\in\Trel\sigma{}$ where $\Vect v=(\List v1k)$. \end{theorem} \begin{remark} One would expect to have rather assumptions of the shape ``$(V_i,v_i)\in\Trelv{\phi_i}{}$''; the problem is that we don't know whether $\Srel{\Trelv{\phi_i}{}}\POS=\Srel{\Trelv{\phi_i}{}}\NEG$. \end{remark} \begin{proof} By induction on the typing derivation of $M$, that is, on $M$. We set $\cP=(x_1:\phi_1,\dots,x_k:\phi_k)$ and, given a term $R$, we use $R'$ for the term $\Substbis R{V_1/x_1,\dots,V_k/x_k}$. We also use $\Vect v$ for the sequence $\List v1k$ and $\Vect x$ for the sequence $\List x1k$. The case $M=x_i$ is straightforward. Assume that $M=\STOP N$ and that $\phi=\EXCL\sigma$ with $\TSEQ\cP N\sigma$. By inductive hypothesis we have $(N',\Matapp{\Psem{N}^{\Vect x}}{\Vect v})\in\Trel\sigma{}$. Therefore $(\STOP{(N')},\Prom{(\Matapp{\Psem{N}^{\Vect x}}{\Vect v})})\in\Srel{\Trelv{\EXCL\sigma}{}}{\epsilon}$ (for $\epsilon=\POS$ or $\epsilon=\NEG$)\footnote{It is not clear whether $\Nrel{\Trelv\phi{}}=\Prel{\Trelv\phi{}}$ for any closed positive type $\phi$, but we don't need this property in this proof, so we leave this technical question unanswered.}. We have $\Srel{\Trelv{\EXCL\sigma}{}}{\epsilon} \subseteq\Crel{\Srel{\Trelv{\EXCL\sigma}{}}{\epsilon}} ={\Srel{\Trel{\EXCL\sigma}{}}{\epsilon}}=\Trel{\EXCL\sigma}{}$ and hence $(M',\Matapp{\Psem{M}^{\Vect x}}{\Vect v}))\in\Trel{\EXCL\sigma}{}$ as contended. Assume that $M=\PAIR{N_\ell}{N_r}$ and $\sigma=\TENS{\psi_\ell}{\psi_r}$ with $\TSEQ{\cP}{N_i}{\psi_i}$ for $i\in\{\ell,r\}$. By inductive hypothesis we have $(N'_i,\Matapp{\Psem{N_i}^{\Vect x}}{\Vect v})\in\Trel{\psi_i}{}$. By Lemma~\ref{lemma:tens-crel} we get \[(\PAIR{N'_\ell}{N'_r},\Matapp{\Psem{\PAIR{N_\ell}{N_r}}^{\Vect x}}{\Vect v})\in\Trel{\TENS{\psi_\ell}{\psi_r}}{}\] as contended, since $\Matapp{\Psem{\PAIR{N_\ell}{N_r}}^{\Vect x}}{\Vect v} =\Tens{\Matapp{\Psem{N_\ell}^{\Vect x}}{\Vect v}}{\Matapp{\Psem{N_r}^{\Vect x}}{\Vect v}}$. The case $M=\IN iN$ (and $\sigma=\PLUS{\psi_\ell}{\psi_r}$) is handled similarly, using Lemma~\ref{lemma:plus-crel}. Assume that $M=\FOLD N$ and $\sigma=\phi=\TREC\zeta\psi$ with $\TSEQ{\cP}{N}{\Subst\psi\phi\zeta}$. By inductive hypothesis we have $(N',\Matapp{\Psem{N}^{\Vect x}}{\Vect v}) \in\Trel{\Subst\psi\phi\zeta}{}$ which implies $(\FOLD{N'},\Matapp{\Psem{N}^{\Vect x}}{\Vect v})\in\Trel{\phi}{}$ by Lemma~\ref{lemma:fold-crel}. Assume that $M=\ABST x\phi N$ and $\sigma=\LIMPL\phi\tau$, with $\TSEQ{\cP,x:\phi}{N}{\tau}$. We must prove that $(\ABST{x}{\phi}{N'},\Matapp{\Psem{\ABST x\phi N}^{\Vect x}}{\Vect v})\in\Srel{(\Limpl{\Trelv\phi{}}{\Trel\tau{}})}\epsilon$ for an arbitrary $\epsilon\in\{\NEG,\POS\}$. So let $(V,v)\in\Srel{\Trelv{\phi}{}}{-\epsilon}$. Since $\Srel{\Trelv{\phi}{}}{-\epsilon}\subseteq\Trel{\phi}{}$, we have $(\Subst{N'}{V}{x},\Psem{N}^{\Vect x,x}(\Vect v,v))\in\Trel{\tau}{}$ by inductive hypothesis. It follows that $(\LAPP{\ABST{x}{\phi}{N'}}{V},\Matapp{\Psem{\ABST x\phi N}^{\Vect x}}{\Vect v}\Appsep v)\in\Trel\tau{}$ by Lemma~\ref{lemma:prob-conv-indep-det}, proving our contention. Assume that $M=\LAPP RN$ with $\TSEQ{\cP}{R}{\LIMPL\phi\sigma}$ and $\TSEQ{\cP}{N}{\phi}$ where $\phi$ is a closed positive type. By inductive hypothesis we have $(R',\Psem R^{\Vect x}(\Vect v))\in\Trel{\LIMPL\phi\sigma}{}$ and $(N',\Matapp{\Psem N^{\Vect x}}{\Vect v})\in\Trel{\phi}{}$ and hence $(\LAPP{R'}{N'},\Matapp{\Psem R^{\Vect x}}{\Vect v}\Appsep(\Matapp{\Psem N^{\Vect x}}{\Vect v}))\in\Trel{\sigma}{}$ by Lemma~\ref{lemma:rel-app-closeness}, that is $(M',\Matapp{\Psem M^{\Vect x}}{\Vect v})\in\Trel{\sigma}{}$. Assume that $M=\FIXT x\sigma N$ with $\TSEQ{\cP,x:\EXCL\sigma}{N}\sigma$. The function $f:\Pcoh{\Tsem\sigma}\to\Pcoh{\Tsem\sigma}$ defined by \begin{align*} f(u)=\Psem N^{\Vect x,x}(\Vect v,\Prom u) \end{align*} is Scott continuous and we have $\Matapp{\Psem M^{\Vect x}}{\Vect v}=\sup_{k\in\Nat}f^k(0)$. By induction on $k$, we prove that \begin{align}\label{eq:adeq-fixpoint-rec} \forall k\in\Nat\quad (M',f^k(0))\in\Trel{\sigma}{}\,. \end{align} The base case is proven by Lemma~\ref{lemma:Trel-sem-closeness}. Assume that $(M',f^k(0))\in\Trel{\sigma}{}$. Choosing an arbitrary $\epsilon$, we have $(\STOP{(M')},\Prom{(f^k(0))}) \in\Srel{\Trelv{\EXCL\sigma}{}}\epsilon\subseteq\Trel{\EXCL\sigma}{}$ and hence by our ``outermost'' inductive hypothesis we have $(\Subst{N'}{\STOP{(M')}}{x},f^{k+1}(0))\in\Trel\sigma{}$ from which we get $(M',f^{k+1}(0))\in\Trel\sigma{}$ by Lemma~\ref{lemma:prob-conv-indep-det} and this ends the proof of~\Eqref{eq:adeq-fixpoint-rec}. We conclude that $(M',\Matapp{\Psem M^{\Vect x}}{\Vect v})\in\Trel\sigma{}$ by Lemma~\ref{lemma:Trel-sem-closeness}. Assume that $M=\GO N$ with $\TSEQ{\cP}{N}{\EXCL\sigma}$. By inductive hypothesis we have $(N',\Matapp{\Psem{N}^{\Vect x}}{\Vect v})\in\Trel{\EXCL\sigma}{}$ which implies $(\GO{N'},\Der{}(\Matapp{\Psem N^{\Vect x}}{\Vect v}))\in\Trel{\sigma}{}$ by Lemma~\ref{lemma:der-crel}, that is $(M',\Matapp{\Psem{M}^{\Vect x}}{\Vect v})\in\Trel{\sigma}{}$. Assume that $M=\PR jN$ with $j\in\{\ell,r\}$, $\sigma=\Tens{\phi_\ell}{\phi_r}$ and $\TSEQ{\cP}{M}{\TENS{\phi_\ell}{\phi_r}}$. By inductive hypothesis we have $(N'\Matapp{,\Psem{N}^{\Vect x}}{\Vect v})\in\Trel{\TENS{\phi_\ell}{\phi_r}}{}$ and hence $(\PR j{N'},\Matapp{\Proj j}{(\Matapp{\Psem{N}^{\Vect x}}{\Vect v})}) \in\Trel{\phi_j}{}$ by Lemma~\ref{lemma:proj-crel} that is $(M',\Matapp{\Psem{M}^{\Vect x}}{\Vect v})\in\Trel{\phi_j}{}$. Assume that $M=\CASE{N}{y_\ell}{R_\ell}{y_r}{R_r}$ with $\TSEQ{\cP}{N}{\PLUS{\phi_\ell}{\phi_r}}$ and $\TSEQ{\cP,y_j:\phi_j}{R_j}{\sigma}$ for $j\in\{\ell,r\}$. By inductive hypothesis we have $(N',\Matapp{\Psem N^{\Vect x}}{\Vect v})\in\Trel{\PLUS{\phi_\ell}{\phi_r}}{}$ and $(\ABST{y_j}{\phi_j}{R'_j},\Matapp{\Psem{\ABST{y_j}{\phi_j}{R_j}}^{\Vect x}}{\Vect v})\in\Trel{\LIMPL{\phi_j}{\sigma}}{}$ for $j\in\{\ell,r\}$ (to prove this latter fact, one chooses $\epsilon\in\{\NEG,\POS\}$ and considers an arbitrary $(V,v)\in\Srel{\Trelv{\phi_j}{}}{-\epsilon}$, we have $(V,v)\in\Trel{\phi_j}{}$ and hence $(\Subst{R'_j}{V}{y_j},\Psem{R_j}^{\Vect x,y_j}(\Vect v,v))\in\Trel{\sigma}{}$ by inductive hypothesis, which implies \[ (\LAPP{\ABST{y_j}{\phi_j}{R'_j}}V, \Matapp{\Psem{\ABST{y_j}{\phi_j}{R_j}}^{\Vect x}}{\Vect v}\Appsep v)\in\Trel{\sigma}{} \] by Lemma~\ref{lemma:prob-conv-indep-det}). By Lemma~\ref{lemma:case-crel} we get \[(\CASE{N'}{y_\ell}{R'_\ell}{y_r}{R'_r}, \Matapp{\Case(\Psem{\ABST{y_\ell}{\phi_\ell}{R_\ell}}^{\Vect x}}{\Vect v}),\Matapp{\Psem{\ABST{y_r}{\phi_r}{R_r}}^{\Vect x}}{\Vect v}))(\Matapp{\Psem N^{\Vect x}}{\Vect v}))\in\Trel{\sigma}{} \] that is $(M',\Psem{M}^{\Vect x}(\Vect v))\in\Trel\sigma{}$, by Lemma~\ref{lemma:prob-conv-indep-det}. Assume that $M=\UNFOLD N$ where $\TSEQ{\cP}{N}{\phi}$ with $\phi=\TREC\zeta\psi$. We apply Lemma~\ref{lemma:unfold-crel} straightforwardly. Assume that $M=\ONELEM$ and the typing derivation consists of the axiom $\TSEQ{\cP}{\ONELEM}{\ONE}$ so that $\sigma=\ONE$. We have $(M,\Psem M{})\in\Trel{\ONE}{}$ by definition since $\Redmats^\infty_{M,\ONELEM}=1=\Psem M{}$. Assume last that $M=\COIN p$ for some $p\in[0,1]\cap\Rational$ and the typing derivation consists of the axiom $\TSEQ{\cP}{\COIN p}{\PLUS\ONE\ONE}$ so that $\sigma=\PLUS\ONE\ONE$. We must prove that $(\COIN p,\Psem{\COIN p}{})\in\Trel{\PLUS\ONE\ONE}{}$. Remember that $\Psem{\COIN p}{}=p\Base{(\ell,*)}+(1-p)\Base{(r,*)}$. Let $\epsilon\in\{\NEG,\POS\}$ and let $(T,t)\in \Srel{(\Limpl{\Trelv{\PLUS\ONE\ONE}{}}{{\Trel{\ONE}{}}})}{\epsilon}$, we must prove that $(\LAPP T{\COIN p},t(p\Base{(\ell,*)}+(1-p)\Base{(r,*)}))\in\Trel{\ONE}{}$. We have \begin{align*} \Redmats^\infty_{\LAPP T{\COIN p},\ONELEM} = p\Redmats^\infty_{\LAPP T{\IN \ell \ONELEM}, \ONELEM}+(1-p)\Redmats^\infty_{\LAPP T{\IN r \ONELEM},\ONELEM} \end{align*} since the first reduction step must be $\COIN p\Redone p\IN \ell \ONELEM$ or $\COIN p\Redone{1-p}\IN r \ONELEM$. By our assumption on $(T,t)$ we have $\Redmats^\infty_{\LAPP T{\IN i\ONELEM}}\geq t(\Base{(i,*)})$ and hence $\Redmats^\infty_{\LAPP T{\COIN p},\ONELEM}\geq t(p\Base{(\ell,*)}+(1-p)\Base{(r,*)})$ as contended, by linearity of $t$. \end{proof} \begin{theorem}[Adequacy]\label{th:rel-ad-lemma} Let $M$ be a closed term such that $\TSEQ{}{M}{\ONE}$. Then $\Psem M{}=\Redmats^\infty_{M,\ONELEM}$. \end{theorem} By Corollary~\ref{th:soundness-ineq} and Theorem~\ref{lemma:adequacy}. \section{Full Abstraction}\label{sec:FA} We prove now the Full Abstraction Theorem~\ref{thm:fa}, that is the converse of the Adequacy Theorem. \subsection{Outline of the proof.} We reason by contradiction and assume that two closed terms $M_1$ and $M_2$ have different semantics. Remember from Section~\ref{subsec:model-pcoh}, that a closed term of type $\sigma$ is interpreted as a vector with indices in the web $\Web{\Psem\sigma}$, so that there is $a\in\Web{\Psem\sigma}$ such that $\Psem{M_1}_a\neq\Psem{M_2}_a$. We want to design a term that will separate $M_1$ and $M_2$ observationally. We define a \emph{testing term} $\TSEQ{}{\Testt a}{\LIMPL{\EXCL\Tnat}{(\LIMPL{\EXCL\sigma}\ONE)}}$ that will depend only on the structure of the element $a$ of the web. We then use properties of the semantics (namely that terms of type $\LIMPL{\EXCL\Tnat}\tau$ can be seen as power series) to find reals $\vec p$ such that $\LAPP{\Testt a}{\STOP{\Ran{\vec p}}}$ separates $M_1$ and $M_2$: $$ \Redmats^\infty_{\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\vec p}}}}{\STOP M_1},\ONELEM}\neq \Redmats^\infty_{\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\vec p}}}}{\STOP M_2},\ONELEM} $$ \medskip \subsubsection*{Let us detail the key points of the proof.} Remember from Section~\ref{subsec:Kleisli_fun} that, because $\TSEQ{}{\Testt a}{\LIMPL{\EXCL\Tnat}{(\LIMPL{\EXCL\sigma}\ONE)}}$, its interpretation $\Psem{\Testt a}$ can be seen as a power series \begin{equation*} \Fun{\Psem{\Testt a}}(\vec\zeta) =\Psem{\Testt a}\Compl {\Prom{\vec\zeta\ }}=\left(\sum_{[\List k1n]\in\Web{\Excl \Tnat}}\Psem{\Testt a}_{[\List k1n],b}\prod_{i=1}^n\zeta_{k_i}\right)_{b\in\Web{\LIMPL{\EXCL\sigma}\ONE}} \end{equation*} with infinitely many parameters $\vec\zeta=(\zeta_0,\dots,\zeta_n,\dots)$. Moreover, if $\sum_{i=0}^\infty \zeta_i\le 1$, then $\vec\zeta\in\Pcoh{\Psem\Tnat}$ and $\Fun{\Psem{\Testt a}}(\vec\zeta)\in\Pcoh{\Psem{\LIMPL{\EXCL\sigma}\ONE}}$ (see Theorem~\ref{prop:kleisli-morph-charact}). The first key point is to remark that the testing term $\Testt a$ is defined in such a way that ${\Psem{\Testt a}}$ has actually only finitely many parameters $\List\zeta 0{\Lent a}$ (meaning that if the support of the multiset $c$ is not included in $\{0,\dots,\Lent a\}$, then $\Psem{\Testt a}_{(c,b)}=0$). Now, for any $u\in\Pcoh{\Psem\sigma}$, $\Psem{\ABST y{\EXCL\sigma}{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}y}}} \Prom u\in\Pcoh{\Psem {\LIMPL{\EXCL\Tnat}\ONE}}$. It is also a power series that depends on the same finitely many parameters $\List\zeta 0{\Lent a}$. The second key point is a separation property of $\Psem{\Testt a }$: we prove that, in the power series $\Psem{\ABST y{\EXCL\sigma}{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}y}}} \Prom u$, the coefficient of the unitary monomial\footnote{That is, the monomial where each exponent is equal to one.} $\prod_{k=0}^{\Lent a}\zeta_k$ is equal to $\coeff - a\,u_a$ with a coefficient $\coeff - a\neq 0$ which depends only on $a$. Now, by assumption, $\Psem{M_1}_a$ and $\Psem{M_2}_a$ have different coefficient. For $i=1,2$, we have ${\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}{\STOP M_i}}}}= \Psem{\ABST y{\EXCL\sigma}{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}y}}}{\Prom{\Psem{M_i}}}$. Thus, the power series ${\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}{\STOP M_i}}}}$ (for $i=1,2$) have \emph{different coefficients}. The last key point uses classical analysis: if two power series with non-negative real coefficients and finitely many parameters have different coefficients, then they differ on non-negative arguments $\vec p$ close enough to zero: $\sum p_i\le 1$, so that $\vec p\in\Pcoh{\Psem\Tnat}$ and ${\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}{\STOP M_1}}}}\Prom{\vec p\,}\neq{\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}{\STOP M_2}}}}\Prom{\vec p\,}$. Finally, in order to substitute in $\Testt a$ the parameters $\vec \zeta$ with the reals $\vec p$, we use $\Ran{\vec p}$ as introduced in Paragraph~\ref{subsec:exsyn}. Indeed, $\Psem{\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\vec p}}}}{\STOP M_i}}= \Fun{\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}{ x}}{\STOP M_i}}}}\Prom{\vec p\,}$. We conclude thanks to the Adequacy Theorem~\ref{th:rel-ad-lemma} that ensures that $$\Redmats^\infty_{\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\vec p}}}}{\STOP M_1},\ONELEM}\neq \Redmats^\infty_{\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\vec p}}}}{\STOP M_2},\ONELEM}$$ \subsection{Notations.} In order to define the testing term $\Testt a$, we will reason by induction and we will need to associate three kinds of \emph{testing terms} with the points of the webs. More precisely: \begin{itemize} \item Given a positive type $\phi$ and $a\in\Web{\Tsem\phi}$, we define a term $\Testv a$ such that \begin{align*} &\TSEQ{}{\Testv a}{\LIMPL{\EXCL\Tnat}{\LIMPL\phi\ONE}}. \end{align*} \item Given a general type $\sigma$ and $a\in\Web{\Tsem\sigma}$, we define terms $\Testa a$ and $\Testt a$ such that \begin{align*} &\TSEQ{}{\Testa a}{\LIMPL{\EXCL\Tnat}\sigma} &\TSEQ{}{\Testt a}{\LIMPL{\EXCL\Tnat}{\LIMPL{\EXCL\sigma}\ONE}}. \end{align*} \end{itemize} We also introduce natural numbers $\Lenv a$, $\Lent a$ and $\Lena a$ depending only on $a$. They represent the finite numbers of parameters on which the power series $\Fun{\Psem{\Testv a}}$, $\Fun{\Psem{\Testt a}}$ and $\Fun{\Psem{\Testa a}}$ depend respectively. We denote as $\coeffv a$, $\coefft a$ and $\coeffa a$ natural numbers depending only on $a$ and that will appear as the coefficient of the unitary monomial $\prod_{k=0}^{\Lenv a}\zeta_k$, $\prod_{k=0}^{\Lent a}\zeta_k$ and $\prod_{k=0}^{\Lena a}\zeta_k$ respectively of the corresponding power series. These numbers are all $>0$. \medskip We use the terms introduced in the probabilistic tests paragraph of Subsection~\ref{subsec:exsyn} and whose semantics are given in Subsection~\ref{subsec:exden}: \begin{itemize} \item $\Ran{\Vect p}$ which reduces to $\Num i$ with probability $p_i$ for $\sum_{i=0}^{k-1}p_i\le 1$, \item $M_0\cdot N$ which reduces to $V$ with probability $ p\,q$ if $M_0$ reduces to $\ONELEM$ with probability $p$ and $N$ reduces to $V$ with probability $q$, \item $M_0\AND\dots\AND M_{k-1}$ which reduces to $\ONELEM$ with probability $\prod_{i=0}^{k-1}p_i$ if $M_i$ reduces to $\ONELEM$ with probability $p_i$, \item $\LAPP{\Pchoose_k^\sigma(\List M0{k-1})}P$ which reduces to $M_i$ with probability $p_i$ if $p_i$ is the probability of $P$ to reduce to $\NUM i$, \item $\LAPP{\Pext lr}{\GO Z}$ and $\LAPP{\Pwin k{\vec n}}{\GO Z}$ to partition the parameters $\GO Z$. Indeed $Z$ will denote a variable of type $\EXCL\Tnat$ and $\GO Z$ has to be considered as the sequence of parameters $\Vect\zeta$ of the power series interpreting testing terms. $\Lenv a$, $\Lent a$ and $\Lena a$ represent the number of parameters on which the respective testing terms depend. We use $\LAPP{\Pwin i{\vec n}}{\GO Z}$ to extract subsequences of $\Vect \zeta$ that will be given as arguments to subterms in the inductive definition of the testing terms. Remember that $\LAPP{\Pwin i{n_0,\dots,n_k}}{\GO Z} \Redone {p_l} \Num l$ if $l$ is in the $i$th window of size $n_i$, that is $n_0+\cdots+n_{i-1}\le l\le n_0+\cdots+n_{i}-1$ and $p_l$ is the probability that $\GO Z$ reduces to $\Num l$, that is the non-negative real parameter $\zeta_l$. This is a key ingredient in the computation of the coefficient of the unitary monomial of the interpretation of testing terms by induction on type and on the structure of $a$. \end{itemize} \subsection{Testing terms.} We define the terms $\Testv a$, $\Testa a$ and $\Testt a$ and the associated natural numbers $\Lenv a$, $\Lena a$ and $\Lent a$, by induction on the structure of the point $a$. \medskip {Let $\phi$ be a positive type and $a\in\Web{\Tsem\phi}$. We define $\Testv a$ and $\Testa a$ by induction on the size of $a$ using the structure of $\phi$} \paragraph{Let $\phi=\EXCL\tau$ and $a=\Mset{\List b0{k-1}}$ with $b_i\in\Web{\Tsem\tau}$.} By inductive hypothesis, we have built terms $\TSEQ{}{\Testt{b_i}}{\LIMPL{\EXCL\Tnat}{\LIMPL{\EXCL\tau}\ONE}}$ and $\TSEQ{}{\Testa{b_i}}{\LIMPL{{\EXCL\Tnat}}\tau}$. Then we set \begin{multline*} \Testv a= \ABST{Z}{\EXCL\Tnat}{\ABST{x}{\EXCL\tau}{\LAPP{\EAPP{\Testt{b_0}}{{\left(\LAPP{\Pwin 0{\Lent{b_0},\dots,\Lent{b_{k-1}}}}\GO Z\right)}}}x\AND\cdots\\ \AND\LAPP{\EAPP{\Testt{b_{k-1}}}{\left(\LAPP{\Pwin {k-1}{\Lent{b_0},\dots,\Lent{b_{k-1}}}}\GO Z\right)}}x}},\\ \coeffv a =\prod_{i=0}^{k-1} \coefft{b_i},\ \text{ and } \Lenv a = \Lent{b_0} + \cdots + \Lent{b_{k-1}}\,. \end{multline*} \begin{multline*} \Testa a = \ABST{Z}{\EXCL\Tnat}{\STOP{\left(\langle\Pchoose^{\tau}_k\left(\EAPP{\Testa{b_0}}{\left(\LAPP{\Pwin{1}{k,\Lena{b_0},\cdots,\Lena{b_{k-1}}}}\GO Z\right)}\right.\right.,\dots,\\\left.\left.\EAPP{\Testa{b_{k-1}}}{\left(\LAPP{\Pwin{k}{k,\Lena{b_0},\cdots,\Lena{b_{k-1}}}}\GO Z\right)}\right)\rangle{\GO Z}\right)}}, \\ \coeffa a =\Factor a \,\prod_{i=0}^{k-1} \coeffa{b_i},\ \text{ and } \Lena a =k+ \Lena{b_0} + \cdots + \Lena{b_{k-1}}\,. \end{multline*} Remember that the factorial $\Factor a$ of a multiset $a$ has been defined in Paragraph~\ref{subsec:Kleisli_fun} as the number of permutations that fix $a$. \paragraph{If $\phi=\TENS{\phi_\ell}{\phi_r}$ and $a=(b_\ell,b_r)$ with $b_i\in\Web{\Tsem{\phi_i}}$ for $i\in\{\ell,r\}$,} then we set \begin{multline*} \Testv a = \ABST{Z}{\EXCL\Tnat}{\ABST{x}{\phi } {\LAPP{\EAPP{\Testv{b_\ell}}{\left(\LAPP{\Pwin 0{\Lenv{b_\ell},\Lenv{b_r}}}\GO Z\right)}}{\PR \ell x}\AND\LAPP{\EAPP{\Testv{b_r}}{\left(\LAPP{\Pwin 1{\Lenv{b_\ell},\Lenv{b_r}}}\GO Z\right)}}{\PR r x}}},\\ \coeffv a = \coeffv{b_\ell}\, \coeffv{b_r},\ \text{ and } \Lenv a=\Lenv{b_\ell}+\Lenv{b_r}\,. \end{multline*} \begin{multline*} \Testa a = \ABST{Z}{\EXCL\Tnat}\PAIR{\EAPP{\Testa{b_\ell}}{\left(\LAPP{\Pwin 0{\Lena{b_\ell},\Lena{b_r}}}\GO Z\right)}}{\EAPP{\Testa{b_r}}{\left(\LAPP{\Pwin 1{\Lena{b_\ell},\Lena{b_r}}}\GO Z\right)}},\\ \coeffa a = \coeffa{b_\ell}\,\coeffa{b_r} \text{ and } \Lena a=\Lena{b_\ell}+\Lena{b_r}\,. \end{multline*} \paragraph{If $\phi=\PLUS{\phi_\ell}{\phi_r}$ and $a=(\ell,a_\ell)$ with $b\in\Web{\Tsem{\phi_\ell}}$ (the case $a=(r,a_r)$ is similar),} then we set \begin{align*} \Testv a = \ABST{Z}{\EXCL\Tnat}{\ABST{x} {\PLUS{\phi_\ell}{\phi_r}}{\CASE{x}{y_\ell} {\LAPP{\LAPP{\Testv {a_\ell}}Z}{y_\ell}}{y_r}{\LOOP\ONE}}}, \quad \coeffv a = \coeffv{a_\ell} \quad\text{and } \Lenv a = \Lenv{a_\ell}\,. \end{align*} \begin{align*} \Testa a =\ABST{Z}{\EXCL\Tnat}{ \IN \ell {\LAPP{\Testa a_\ell}{Z}}}, \quad \coeffa a = \coeffa{a_\ell}\quad\text{and } \Lena a = \Lena{a_\ell}\,. \end{align*} \medskip Finally, for a general type $\sigma$ and $a\in\Web{\Tsem\sigma}$, we define $\Testa a$ and $\Testt a$. \paragraph{If $\sigma=\phi$ is positive,} then we have already defined $\Testa a$. \\ Let us now define $\Testt a$. This term does not depend on the structure of $\phi$: \begin{align*} \Testt a = \ABST Z{\EXCL\Tnat}{\ABST x{\EXCL\phi}{\LAPP{\LAPP{\Testv a}Z}{\GO x}}} , \quad \coefft a = \coeffv{a}\quad\text{and } \Lent a = \Lenv{a} \,. \end{align*} \paragraph{If $\sigma=\LIMPL\phi\tau$ and $a=(b,c)$ with $b\in\Web{\Tsem\phi}$ and $c\in\Web{\Tsem\tau}$,} then we set \begin{multline*} \Testa a =\ABST{Z}{\EXCL\Tnat}{ \ABST{x}{\phi} {{{\LAPP{\EAPP{\Testv b}{\left(\LAPP{\Pwin 0{\Lenv b,\Lena c}}\GO Z\right)}}x}\,\cdot\,{{\EAPP{\Testa c}{\left({\LAPP{\Pwin 1{\Lenv b,\Lena c}}\GO Z}\right)}}}}}}\,,\\ \coeffa a = \coeffv{b}\,\coeffa{c}\,, \text{ and }\Lena a=\Lenv b+\Lena c\,. \end{multline*} \begin{multline*} \Testt a = \ABST Z{\EXCL\Tnat}{\ABST{f}{\EXCL{(\LIMPL\phi\tau)}}{ {\LAPP{\EAPP{\Testt c}{\left(\LAPP{\Pwin 1{\Lena b,\Lent c}}{\GO Z}\right)}}{\STOP{\left(\LAPP{\GO f}{\EAPP{\Testa b}{\left(\LAPP{\Pwin 0{\Lena b,\Lent c}}\GO Z\right)}}\right)}}}}}\\ \coefft a = \coeffa{b}\,\coefft{c},\, \text{ and }\Lent a=\Lena b+\Lent c\,. \end{multline*} It is easy to check that these terms satisfy the announced typing judgments. It is also clear that $\coeffv a$, $\coeffa a$ and $\coefft a$ are non zero natural numbers. \medskip We will now tackle the proof of the main observation: that is that the semantics of $\Testt a$ is a power series with finitely many parameters and whose coefficient of the unitary monomial can be seen as a morphism in $\Pcoh{\Tsem{\LIMPL{\EXCL\sigma}\ONE}}$. Lemma~\ref{lem:coeff-morph} introduces notations for the unitary monomials and provides useful properties for proving the key Lemma~\ref{lem:point-test} which gives the coefficients of these monomials. \begin{lemma}\label{lem:coeff-morph} Let $\sigma$ be a general type and $t\in\Pcoh{\Tsem{\LIMPL{\EXCL\Tnat}{\sigma}}}$. \begin{enumerate} \item Assume that there is $k\in\Nat$ such that for any $c\in\Web{\Tsem{\sigma}}$, the power series $\Fun t_c$ over $\Pcoh{\Tsem\Tnat}$ depends on the $k$ first parameters. For any $c\in\Web{\Tsem\sigma}$, let us denote as $\mon{\vec\zeta}{\Fun t}_c$ the coefficient of the monomial $\Listbis\zeta 0{k-1}$ of $\Fun t_c$. Then, $k^{-k}\,\mon{\vec\zeta}{\Fun t}\in\Pcoh{\Tsem\sigma}$. \label{item1:coeff-morph} \item Assume moreover that $\sigma=\LIMPL\phi\tau$ where $\phi$ is a positive type and $\tau$ a general type. Let $m\in\Pcoh{\Tsem\tau}$ and $a\in\Web{\Tsem{\phi}}$. If $\forall u\in \PcohEM{\Tsemca\phi}\ \mon{\vec\zeta}{\Fun t}u =mu_a$ then $\forall u\in \Pcoh{\Tsem\phi}\ \mon{\vec\zeta}{\Fun t}u =mu_a$. \label{item2:coeff-morph} \end{enumerate} \end{lemma} \begin{proof} We prove~\Eqref{item1:coeff-morph}. First notice that $\forall c\in\Web{\Tsem\sigma}$, the coefficient of the monomial $\prod_{i=0}^{k-1}\zeta_i$ is $\mon{\vec\zeta}{\Fun t}_c=t_{([0,\dots,k-1],c)}$. Now, let $\Vect{\tfrac1k}$ be the sequence of $k$ coefficients all equal to $\frac1k$: \begin{equation*} \Fun{t}_c(\vec{\tfrac 1k})=\displaystyle\sum_{\substack{\mu\in\Web{\Tsem{\EXCL\Tnat}}\\\Supp\mu\subseteq\{0,\dots,k-1\}}}t_{(\mu,c)}{\tfrac 1k}^{\Card\mu} \end{equation*} so that $\mon{\vec\zeta}{\Fun t}_c\, k^{-k}\le \Fun{t}(\vec{\tfrac 1k})_c$. Since $\Vect{\tfrac 1k}\in\Pcoh{\Tsem\Tnat}$, $\Fun{t}(\vec{\tfrac 1k})\in\Pcoh{\Tsem\sigma}$ which is downward closed, we have that $k^{-k}\,\mon{\vec\zeta}{\Fun t}\in\Pcoh{\Tsem\sigma}$. Now we prove~\Eqref{item2:coeff-morph}. For any $a\in\Web{\Tsem{\phi}}$, there is $A$ such that $u_a\le A$ for any $u\in\Pcoh{\Tsem\phi}$ (see Subsection~\ref{subsec:model-pcoh}). Hence, for any $u\in\Pcoh{\Tsem\phi}$, $\tfrac{u_a}A m\in\Pcoh{\Tsem\tau}$ and we deduce thanks to Lemma~\ref{lemma:PCS-moprh-charact} that $u\mapsto u_a\,m\,A^{-1}$ is in $\Pcoh{\Tsem{\LIMPL\phi\tau}}$. Without loss of generality, we can choose $A\ge k^{k}$ so that $u\mapsto u_a\,m\,A^{-1}$ and $u\mapsto A^{-1}\mon{\vec\zeta}{\Fun t}u$ are both in $\Pcoh{\Tsem{\LIMPL\phi\tau}}$. Now, since $\Tsemca\phi$ is dense (see Theorem~\ref{th:pos-types-dense}), if $u\mapsto A^{-1}\,\mon{\vec\zeta}{\Fun t}u$ and $u\mapsto A^{-1}\,u_a\,m$ are equal on all coalgebraic points $u\in\PcohEM{\Tsemca\phi}$, they are equal (see Definition~\ref{def:dense}). Thus, for all $u\in\Pcoh{\Tsem\phi}$, $u_a\, m=\mon{\vec\zeta}{\Fun t}u$. \end{proof} We are now ready to prove that the coefficient of the unitary monomial of a testing term associated with a point $a$ of the web allows to extract the $a$-coefficient of an argument, up to a non-zero coefficient depending only on $a$. This is central in the proof of Full Abstraction. Let us first introduce some notations that will be used along this proof. Intuitively, for $s\in \Pcoh{\Tsem{\LIMPL{\EXCL\Tnat}\sigma}}$, we reason on power series $\Fun s$ with values in $\Pcoh{\Tsem{\sigma}}$. But formally, we reason on the non-negative real power series $\Fun s_{a'}$ defined for each $a'\in\Web{\Tsem\sigma}$ and for all parameters\footnote{\label{footnote:parameters}We follow the common mathematical practice of using the same notation $\Vect\zeta=(\List\zeta0n)$ to refer to the formal parameters of a power series and to real arguments of the corresponding function. } $\vec\zeta\in\Pcoh{\Tsem\iota}$ as $\Fun s_{a'}(\vec\zeta)=(\Fun s(\vec\zeta))_{a'}=(s\Prom{\left.\vec \zeta\,\right.})_{a'}$ (see Paragraph~\ref{subsec:Kleisli_fun}). We want to compute the unitary monomial of $\Fun s$ which will be in $\Pcoh{\Tsem\sigma}$. We define it for each $a'\in\Web{\Tsem\sigma}$ as $\mon{\vec\zeta}{s}_{a'}=\mon{\vec\zeta}{\Fun s_{a'}}$. We will also use the fact that a morphism $t\in\Pcoh{\Tsem{\LIMPL\phi\ONE}}$ is defined by the collection of $t_{(a',\ast)}$ for $a'\in\Web{\Tsem\phi}$ and is extensionally characterized by its values $t\,u$ on every $u\in \Pcoh{\Tsem\phi}$. Indeed, for any $a'\in\Web{\Tsem\phi}$, there is $\epsilon>0$ such that $\epsilon \Base{a'}\in\Pcoh{\Tsem\phi}$ and $t_{(a',\ast)}=\tfrac1\epsilon\,(t\,\epsilon\Base{a'})_{\ast}$ by linearity of matrix multiplication. \begin{lemma}\label{lem:point-test} Let $\sigma$ be a type and $a\in\Web{\Tsem\sigma}$. \begin{enumerate} \item\label{point-test0} Assume that $\sigma=\phi$ is positive. If $a'\in\Web{\Tsem\phi}$, then $\Fun{\Psem{\Testv a}}_{(a',\ast)}$ is a power series over $\Pcoh{\Tsem\Tnat}$ depending on $\Lenv a$ parameters, so we define $\mon{\vec\zeta}{\Psem{\Testv a}}_{(a',\ast)}=\mon{\vec\zeta}{\Fun{\Psem{\Testv a}}_{(a',\ast)}}$ and check that $\mon{\vec\zeta}{\Psem{\Testv a}}\in\Pcoh{\Tsem{\LIMPL\phi\ONE}}$ and that for any $u\in\Pcoh{(\Tsem\phi)}$, $\mon{\vec\zeta}{\Psem{\Testv a}}u=\coeffv a\,u_a$. \label{point-testa0} \item Assume that $\sigma$ is a general type. For any $a'\in\Web{\Tsem\sigma}$, $\Fun{\Psem{\Testa a}}_{a'}$ is a power series over $\Pcoh{\Tsem\Tnat}$ depending on $\Lena a$ parameters, so we define $\mon{\vec\zeta}{\Psem{\Testa a}}_{a'}=\mon{\vec\zeta}{\Fun{\Psem{\Testa a}}_{a'}}$ and check that $\mon{\vec\zeta}{\Psem{\Testa a}}\in\Pcoh{\Tsem\sigma}$ and that $\mon{\vec\zeta}{\Psem{\Testa a}}=\coeffa a\,\Base a$ where $\Base a$ is the base vector such that $(\Base a)_{a'}=\Kronecker{a'}a$ for $a'\in\Web{\Tsem\sigma}$. \label{point-testa+} \item Let $\sigma$ be a general type. For any $a'\in\Web{\Tsem{\EXCL\sigma}}$, $\Fun{\Psem{\Testt a}}_{(a',\ast)}$ is a power series over $\Pcoh{\Tsem\Tnat}$ depending on $\Lent a$ parameters, so we define $\mon{\vec\zeta}{\Psem{\Testt a}}_{(a',\ast)}=\mon{\vec\zeta}{\Fun{\Psem{\Testt a}}_{(a',\ast)}}$ and check that $\mon{\vec\zeta}{\Psem{\Testt a}}\in\Pcoh{\Tsem{\LIMPL{\EXCL\sigma}\ONE}}$ and that for any $u\in \Pcoh{\Tsem\sigma}$, $\mon{\vec\zeta}{\Psem{\Testt a}}\Prom u=\coefft a\,u_a$. \label{point-testa-} \end{enumerate} \end{lemma} \begin{proof} Let us argue by mutual induction on the size of $a$ and the structure of $\phi$. \medskip Let $\phi$ be a positive type and $a\in\Web{\Tsem\phi}$. We prove~\Eqref{point-testa+} and~\Eqref{point-testa0} by induction on the structure of $\phi$ \paragraph{Assume that $\phi=\EXCL\tau$ and that $a=\Mset{\List b0{k-1}}$} with $b_i\in\Web{\Tsem\tau}$. \smallskip We prove~\Eqref{point-testa0}. Let $a'=\Mset{\List{b'}0{k'-1}}\in\Tsem{\Excl\tau}$ with $b'_j\in\Web{\Tsem\tau}$ and $\vec\zeta\in\Pcoh{\Tsem\Tnat}$ be the concatenation of the finite sequences\footnote{We assume that the support of indices of the sequences are disjoint even if this requires some renaming.} $\vec\zeta^i\in\Pcoh{\Tsem\Tnat}$ such that the length of $\vec\zeta^i$ is $\Lent{b_i}$. By Theorem~\ref{th:pos-types-dense}, $\Fun{\Psem{\Testv a}}_{(a',\ast)}(\vec\zeta)=(\Psem{\Testv a}\,\Prom{\left.\vec\zeta\right.})_{(a',\ast)}\in\Pcoh{\Tsem{\LIMPL{\EXCL\tau}\ONE}}$ is completely determined by the function $u\mapsto \Psem{\Testv a}\,\Prom{\left.\vec\zeta\right.}\,\Prom u$ defined on $\Pcoh{\Psem{\tau}}$. By inductive hypothesis, $\Fun{\Psem{\Testv a}}_{(a',\ast)}$ depends on finitely many parameters $\Lenv a=\Lent{b_0}+\dots+\Lent{b_{k-1}}$, since \begin{align*} \Psem{\Testv a}\Prom{\left.\vec\zeta\right.}\Prom u = \prod_{i=0}^{k-1}\Psem{\Testt{b_i}}\Prom{\left.\vec\zeta^i\right.}\Prom u \qquad\text{and therefore}\qquad \mon{\vec\zeta}{ \Psem{\Testv a}}\Prom u = \prod_{i=0}^{k-1}\mon{\vec\zeta^i}{\Psem{\Testt{b_i}}}\Prom{u}. \end{align*} Again, by inductive hypothesis it follows that \[\mon{\vec\zeta}{\Psem{\Testv{a}}}\Prom u=\prod_{i=0}^{k-1}\coefft{b_i}\,u_{b_i}=\coeffv a\,(\Prom u)_a. \] We can apply~\Eqref{item2:coeff-morph} of Lemma~\ref{lem:coeff-morph}, so that we have $\mon{\vec\zeta}{\Psem{\Testv{a}}}u=\coeffv a\,u_a$ for all $u\in\Pcoh{\Tsem{\EXCL\tau}}$. \smallskip We prove~\Eqref{point-testa+}. Let $a'=\Mset{\List{b'}0{k'-1}}\in\Tsem{\Excl\tau}$ and $\vec\zeta\in\Pcoh{\Tsem\Tnat}$ be the concatenation of the finite sequences $\vec\zeta^\ast,\vec\zeta^0,\dots,\vec\zeta^{k-1}\in\Pcoh{\Tsem\Tnat}$ such that the length of $\Vect\zeta^{\ast}$ is $k$ and the length of $\vec\zeta^{i}$ is $\Lena{b_i}$ for $i\ge 0$. By inductive hypothesis, $\Fun{\Psem{\Testa a}}_{a'}$ depends on finitely many parameters $\Lenv a=k+\Lena{b_0}+\dots+\Lena{b_{k-1}}$, since \begin{align} \label{eq:aplus-expression} \Fun{\Psem{\Testa a}}_{a'}(\vec \zeta)=(\Psem{\Testa a}\Prom{\left.\vec\zeta\right.})_{a'} &= \Promp{\sum_{i=0}^{k-1} {\zeta^\ast_i}\ \Psem{\Testa{b_i}}\,\Prom{\left.\vec\zeta^{i}\right.}}_{a'}=\prod_{j=0}^{k'-1}\left(\sum_{i=0}^{k-1} {\zeta^\ast_i}\ \Psem{\Testa{b_i}}\,\Prom{\left.\vec\zeta^{i}\right.}\right)_{b'_j}\,. \end{align} We want to compute the coefficient of the unitary monomial, which contains exactly one copy of each parameter of each $\Vect\zeta^i$. If $k'\neq k$ then expression~\Eqref{eq:aplus-expression} contains no monomial where each parameter of $\Vect\zeta^\ast$ appears exactly once, so that $\mon{\vec\zeta}{\Psem{\Testa a}}_{a'}=0$ in that case. If $k'=k$ and $\mathfrak{S}_k$ is the set of permutations over $k$, then by using the fact that factorial $\Factor a=\Card\{\rho\in\mathfrak S_k \mid \forall i\, b_i=b_{\rho(i)}\}$, by denoting the Kronecker symbol as $\Kronecker{a}{a'}$ and by the inductive hypothesis, we get: \begin{align*} \mon{\vec\zeta}{\Psem{\Testa a}}_{a'}=\sum_{\rho\in\mathfrak{S}_k}\prod_{i=0}^{k-1}\mon{\vec\zeta^i}{\Psem{\Testa{b_i}}}_{b'_{\rho(i)}}=\Factor a\ \prod_{i=0}^{k-1} \coeffa{b_i}\Kronecker{b_i}{b'_i}=\coeffa a\, \Kronecker{a}{a'}=\coeffa a\, (\Base a)_{a'}\,. \end{align*} \paragraph{ Assume that $\phi=\TENS{\phi_\ell}{\phi_r}$ and that $a=(b_\ell,b_r)$ with $a_i\in\Web{\Tsem{\phi_i}}$.} Let $\vec\zeta\in\Pcoh{\Tsem\Tnat}$ be the concatenation of the finite sequences $\vec\zeta^\ell,\vec\zeta^r\in\Pcoh{\Tsem\Tnat}$ such that the length of $\vec\zeta^i$ is $\Lena{b_i}$. \smallskip We prove~\Eqref{point-testa0}. Let $a'=(b'_\ell,b'_r)\in\Web{\Tsem{\TENS{\phi_\ell}{\phi_r}}}$. By Theorem~\ref{th:pos-types-dense}, $\Fun{\Psem{\Testv a}}_{(a',\ast)}(\vec\zeta)=(\Psem{\Testv a}\,\Prom{\left.\vec\zeta\right.})_{(a',\ast)}\in\Pcoh{\Tsem{\LIMPL{\TENS{\phi_\ell}{\phi_r}}\ONE}}$ is completely determined by the function $u\mapsto \Psem{\Testv a}\,\Prom{\left.\vec\zeta\right.}\,\Prom u$ defined on $\Pcoh{\Psem{\Tens{\phi_\ell}{\phi_r}}}$. Besides, if $u\in\PcohEM{(\Tsemca{\TENS{\phi_\ell}{\phi_r}})}$, then $u=\Tens{u_\ell}{u_r}$ where $u_i=\Projt i(u)\in\PcohEM{(\Tsemca{\phi_i})}$ for $i\in\{\ell,r\}$ (see Lemma~\ref{lemma:sem-values}). Therefore, by inductive hypothesis, $\Fun{\Psem{\Testv a}}_{(a',\ast)}$ depends on finitely many parameters $\Lenv a=\Lenv{b_\ell}+\Lenv{b_r}$, since \begin{align*} \Psem{\Testv a}\Prom{\left.\vec\zeta\right.}u &= \Psem{\Testv{b_\ell}}\Prom{\left.\vec\zeta^\ell\right.}u_\ell\,\Psem{\Testv{b_r}}\Prom{\left.\vec\zeta^r\right.}u_r \quad \text{ and therefore }\quad \mon{\vec\zeta}{ \Psem{\Testv a}}u= \mon{\vec\zeta^\ell}{\Psem{\Testv{b_\ell}}}u_\ell\,\mon{\vec\zeta^r}{\Psem{\Testv{b_r}}}u_r. \end{align*} Hence, by inductive hypothesis $\mon{\vec\zeta}{ \Psem{\Testv a}}u=\coeffv{b_\ell}(u_\ell)_{b_\ell}\,\coeffv{b_r}(u_r)_{b_r}=\coeffv{a}u_{a}$ for $u\in\PcohEM{\Tsemca\phi}$. We conclude by Lemma~\ref{lem:coeff-morph}, that this holds also for $u\in\Pcoh{\Tsem\phi}$. \smallskip We prove~\Eqref{point-testa+}. Let $a'=(b'_\ell,b'_r)\in\Web{\Tsem{\TENS{\phi_\ell}{\phi_r}}}$. By inductive hypothesis, $\Fun{\Psem{\Testa a}}_{a'}$ depends on finitely many parameters $\Lena a=\Lena{b_\ell}+\Lena{b_r}$, since \begin{align*} \Fun{\Psem{\Testa a}}_{a'}(\vec\zeta)= (\Psem{\Testa a}\Prom{\vec\zeta\,})_{a'} = \Tens{(\Psem{\Testa{b_\ell}}\Prom{\left.\vec\zeta^\ell\right.})_{b'_\ell}}{(\Psem{\Testa{b_r}}\Prom{\left.\vec\zeta^r\right.})_{b'_r}} \end{align*} We deduce using inductive hypothesis that \begin{align*} \mon{\vec\zeta}{\Psem{\Testa a}}_{a'} = \mon{\vec\zeta^\ell}{\Psem{\Testa{b_\ell}}}_{b'_\ell}\,\mon{\vec\zeta^r}{\Psem{\Testa{b_r}}}_{b'_r}=\coeffa{b_\ell}\,\Kronecker{b_\ell}{b'_\ell}\ \coeffa{b_r}\,\Kronecker{b_r}{b'_r}=\coeffa{a}\,\Kronecker{a}{a'}. \end{align*} \paragraph{ Assume that $\phi=\PLUS{\phi_\ell}{\phi_r}$ and that $a=(\ell,a_\ell)$ with $a_\ell\in\Web{\Tsem{\phi_\ell}}$}(the case $a=(r,a_r)$ is similar). \smallskip We prove~\Eqref{point-testa0}. Let $a'=(i,a'_i)\in \Web{\Tsem{\PLUS{\phi_\ell}{\phi_r}}}$. By Theorem~\ref{th:pos-types-dense}, $\Fun{\Psem{\Testv a}}_{(a',\ast)}(\vec\zeta)=(\Psem{\Testv a}\,\Prom{\left.\vec\zeta\right.})_{(a',\ast)}\in\Pcoh{\Tsem{\LIMPL{\PLUS{\phi_\ell}{\phi_r}}\ONE}}$ is completely determined by the function $u\mapsto \Psem{\Testv a}\,\Prom{\left.\vec\zeta\right.}\,u$ defined on $\PcohEM{\Tsemca{\PLUS{\phi_\ell}{\phi_r}}}$. Besides, if $u\in\PcohEM{(\Tsemca{\PLUS{\phi_\ell}{\phi_r}})}$, then there is $i\in\{\ell,r\}$ such that $u=\Inj i {u_i}$ with $u_i\in\PcohEM{(\Tsemca{\phi_i})}$ (see Lemma~\ref{lemma:sem-values}). Therefore, by inductive hypothesis, $\Fun{\Psem{\Testv a}}_{(a',\ast)}$ depends on finitely many parameters $\Lenv a=\Lenv{a_\ell}$, since if $i=\ell$, then \begin{align*} \Psem{\Testv a}\Prom{\left.\vec\zeta\right.}u=\Psem{\ABST{x} {\PLUS{\phi_\ell}{\phi_r}}{\CASE{x}{y_\ell} {\LAPP{\Testv{a_\ell}}{y_\ell}}{y_r}{\LOOP\ONE}}}\Prom{\left.\vec\zeta\right.}u=\Psem{\Testv{a_\ell}}\Prom{\left.\vec\zeta\right.}u_\ell \end{align*} and if $i=r$ then $\Psem{\Testv a}\Prom{\left.\vec\zeta\right.}u=\Psem{\Loopt\One}=0$. So we can compute that $\mon{\vec\zeta}{\Psem{\Testv a}}u=\coeffv{a_\ell}(u_\ell)_{a_\ell}=\coeffv a\, u_a$ for $u\in\PcohEM{\Tsemca\phi}$ and this still holds for $u\in\Pcoh{\Tsem\phi}$ by Lemma~\ref{lem:coeff-morph}. \smallskip We prove~\Eqref{point-testa+}. Let $a'=(i,a'_i)\in \Web{\Tsem{\PLUS{\phi_\ell}{\phi_r}}}$. By inductive hypothesis, $\Fun{\Psem{\Testa a}}_{a'}$ depends on finitely many parameters $\Lena a=\Lena{a_\ell}$, since \begin{align*} \Fun{\Psem{\Testa a}}_{a'}(\vec\zeta)= (\Psem{\Testa a}\Prom{\left.\vec\zeta\right.})_{a'}=(\Psem{\IN{\ell}{\Testa{a_\ell}}}\Prom{\left.\vec\zeta\right.})_{(i,a'_i)}=\Injtr{\ell}(\Psem{\Testa{a_\ell}}\Prom{\left.\vec\zeta\right.})_{i,a'_i}=\Kronecker\ell i\, (\Psem{\Testa{a_\ell}}\Prom{\left.\vec\zeta\right.})_{a'_i}. \end{align*} We can therefore compute $ \mon{\vec\zeta}{\Psem{\Testa a}}_{(i,a'_i)} =\Kronecker\ell i\,\mon{\vec\zeta}{\Psem{{\Testa{a_\ell}}}}_{a'_i} =\coeffa a\,\Kronecker{a}{(i,a'_i)} $ by inductive hypothesis. \medskip Finally, for a general type $\sigma$ and $a\in\Web{\Tsem\sigma}$, we prove~\Eqref{point-testa+} and~\Eqref{point-testa-}. \paragraph{If $\sigma=\phi$ is positive,} we have already proved~\Eqref{point-testa+}. Let us prove~\Eqref{point-testa-}. Let $a\in\Web{\Tsem{\phi}}$. Let $a'\in\Web{\Tsem{\EXCL\phi}}$. By Theorem~\ref{th:pos-types-dense} and~Lemma~\ref{lemma:sem-values}, $\Fun{\Psem{\Testt a}}_{(a',\ast)}(\vec\zeta)=(\Psem{\Testt a}\,\Prom{\left.\vec\zeta\right.})_{(a',\ast)}\in\Pcoh{\Tsem{\LIMPL{\EXCL\tau}\ONE}}$ is completely determined by $u\mapsto\Psem{\Testt a}\,\Prom{\left.\vec\zeta\right.}\,\Prom u$ defined on $\Pcoh{\Tsem{\tau}}$. Therefore, by inductive hypothesis, $\Fun{\Psem{\Testt a}}_{(a',\ast)}$ depends on finitely many parameters $\Lent a=\Lenv{a}$, since \begin{align*} \Psem{\Testt a}\Prom{\left.\vec\zeta\right.}\Prom u =\Psem{\Testv a}\Prom{\left.\vec\zeta\right.}u \qquad\text{and therefore}\qquad \mon{\vec\zeta}{\Psem{\Testt a}}\Prom u=\mon{\vec\zeta}{\Psem{\Testv a}}u \end{align*} By inductive hypothesis, $\mon{\vec\zeta}{\Psem{\Testv a}}u= \coeffv a\, u_a=\coefft a\,u_a$. \paragraph{ Last, let $\sigma=\LIMPL\phi\tau$} Let $a=(b,c)\in\Web{\Tsem\sigma}$. \smallskip We prove~\Eqref{point-testa+}. Let $a'=(b',c')\in\Web{\Tsem\sigma}$. Let $\vec\zeta\in\Pcoh{\Tsem\Tnat}$ be the concatenation of the finite sequences $\vec\zeta^1,\vec\zeta^2\in\Pcoh{\Tsem\Tnat}$ such that the length of $\vec\zeta^1$ is $\Lenv{b}$ and the length of $\vec\zeta^2$ is $\Lena{c}$. By inductive hypothesis, for any $u\in\Pcoh{\Tsem\phi}$, \begin{align*} \Fun{\Psem{\Testa a}}(\vec\zeta)u= \Psem{\Testa a}\Prom{\left.\vec\zeta\right.}u=(\Psem{\Testv b}\Prom{\left.\vec\zeta^1\right.}u)_{\ast}\,(\Psem{\Testa c}\Prom{\left.\vec\zeta^2\right.}). \end{align*} Now, let $\epsilon>0$ such that $\epsilon\Bcanon{b'}\in\Pcoh{\Tsem{\phi}}$, we compute \begin{align*} \Fun{\Psem{\Testa a}}_{(b',c')}(\Prom{\left.\vec\zeta\right.})=\tfrac 1\epsilon (\Psem{\Testa a}\Prom{\left.\vec\zeta\right.} \,\epsilon\Bcanon{b'})_{c'}=\tfrac 1\epsilon(\Psem{\Testv b}\Prom{\left.\vec\zeta^1\right.}\,\epsilon\Bcanon{b'})_{\ast}\,(\Psem{\Testa c}\Prom{\left.\vec\zeta^2\right.})_{c'}=\Fun{\Psem{\Testv b}}_{(b',\ast)}(\vec\zeta^1)\,\Fun{\Psem{\Testa c}}_{c'}(\vec\zeta^2). \end{align*} Therefore, by inductive hypothesis, $\Fun{\Psem{\Testa a}}_{(b',c')}$ depends on finitely many coefficients $\Lena a=\Lenv b+\Lena c$. We compute using inductive hypothesis that \begin{align*} \mon{\vec\zeta}{\Psem{\Testa a}}_{(b',c')}=\mon{\vec\zeta^1}{\Psem{\Testv b}}_{(b',\ast)}\,\mon{\vec\zeta^2}{\Psem{\Testa c}}_{c'}=\coeffv b\,\Kronecker{b}{b'}\,\coeffa c\,\Kronecker{c}{c'}. \end{align*} We conclude that $\mon{\vec\zeta}{\Psem{\Testa a}}=\coeffv b\,\coeffa c\,\Bcanon{b,c}=\coeffa a\,\Bcanon a$. \smallskip We prove~\Eqref{point-testa-}. Let $a'=[(b'_0,c'_0),\dots,(b'_{k-1},c'_{k-1})])\in\Web{\Tsem{\EXCL\sigma}}$. Let $\vec\zeta\in\Pcoh{\Tsem\Tnat}$ be the concatenation of the finite sequences $\vec\zeta^1,\vec\zeta^2\in\Pcoh{\Tsem\Tnat}$ such that the length of $\vec\zeta^1$ is $\Lena{b}$ and the length of $\vec\zeta^2$ is $\Lent{c}$. For any $w\in\Pcoh{\Tsem{\EXCL{(\LIMPL{\phi}\tau}})}$, we have: \begin{align*} \Psem{\Testt a}\Prom{\left.\vec\zeta\right.} w = \Psem{\Testt c}\Prom{\left.\vec\zeta^2\right.}\Prom{\left(\Derel{\Tsem{\LIMPL\phi\tau}}(w)\Psem{\Testa b}\Prom{\left.\vec\zeta^1\right.}\right)}. \end{align*} Let $\epsilon>0$ such that $\epsilon\Bcanon{a'}\in\Pcoh{\Tsem{\EXCL{(\LIMPL\phi\tau})}}$, then \begin{align*} \Fun{\Psem{\Testt a}}_{(a',\ast)}(\vec\zeta)=\tfrac 1\epsilon(\Psem{\Testt a}\Prom{\left.\vec\zeta\right.}\epsilon\Bcanon{a'})_\ast=\tfrac 1\epsilon(\Psem{\Testt c}\Prom{\left.\vec\zeta^2\right.}\Prom{\left(\Derel{\Tsem{\LIMPL\phi\tau}}(\epsilon\Bcanon{a'})\Psem{\Testa b}\Prom{\left.\vec\zeta^1\right.}\right)})_\ast. \end{align*} By inductive hypothesis, we get that $\Fun{\Psem{\Testt a}}_{(a',\ast)}$ depends on $\Lent a=\Lenv b+\Lent c$ coefficients. Let now $u\in\Pcoh{\Psem{\LIMPL\phi\tau}}$, then by~Lemma~\ref{lemma:sem-values} $\Derel{\Tsem{(\LIMPL\phi\tau)}}(\Prom u)=u$ and we compute: \begin{align*} \mon{\vec\zeta}{\Psem{\Testt a}}{\Prom u}=\mon{\vec\zeta^1}{\mon{\vec\zeta^2}{\Psem{\Testt c}}\Prom{\left({u\Psem{\Testa b}\Prom{\left.\vec\zeta^1\right.}}\right)}}\,. \end{align*} By inductive hypothesis, we have $\mon{\vec\zeta^2}{\Psem{\Testt c}}\Prom{({u\Psem{\Testa b}\Prom{\left.\vec\zeta^1\right.}})}=\coefft c\, (u\Psem{\Testa b}\Prom{\left.\vec\zeta^1\right.})_c$. Moreover, notice that $u\in\Pcoh{\Tsem{\LIMPL\phi\tau}}$, seen as a morphism in $\PCOH(\Tsem\phi,\Tsem\tau)$ is linear, and there is $\epsilon>0$ such that $\epsilon \Bcanon b\in\Pcoh{\Tsem\phi}$), so that we can apply $u$ to $\Bcanon b$. Now, by using inductive hypothesis, we get that $\mon{\vec\zeta^1}{u\Psem{\Testa b}}_c=(u\,\mon{\vec\zeta^1}{\Psem{\Testa b}})_c= (u\,\coeffa b\,e_{b})_c=\coeffa b\,u_{(b,c)}$. Therefore, we have $\mon{\vec\zeta}{\Psem{\Testt a}}\Prom u=\coeffa b\,\coefft c\,u_{(b,c)}=\coefft a\, u_{a}$. \end{proof} \begin{theorem}[Full Abstraction]\label{thm:fa} \hspace{1cm} If $\TSEQ{}{M_1}{\sigma}$ and $\TSEQ{}{M_2}{\sigma}$ satisfy $M_1\Rel\Obseq M_2$ then $\Psem{M_1}=\Psem{M_2}$. \end{theorem} \begin{proof} Towards a contradiction, assume that $\Psem{M_1}\not=\Psem{M_2}$. There is $a\in\Web{\Tsem\sigma}$ such that $\Psem{M_1}_a\neq \Psem{M_2}_a$. Then by Lemma~\ref{lem:point-test}, $\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}x}{\STOP{M_i}}}}$, for $i\in \{1,2\}$, are power series with different coefficients, namely the coefficients of the monomial $\Listbis \zeta 0{\Lent a-1}$ are $\coefft a\,\Psem{M_i}_a$ for $i\in\set{1,2}$ as $\Psem{\ABST x{\EXCL\Tnat}{\LAPP{\LAPP{\Testt a}x}{\STOP{M_i}}}}{\Prom{\left.\Vect\zeta\right.}}=\Psem{\Testt a}{\Prom{\left.\Vect\zeta\right.}}{\Prom{\Psem{M_i}}}$. There is $\vec\zeta=(\List\zeta0{\Lent a-1})\in\Pcoh{\Tsem\Tnat}$ with $\zeta_i\in\mathbb Q\cap[0,1]$ such that $\Psem{\Testt a}\Prom{\left.\vec\zeta\right.}\Prom{\Psem{M_1}}\neq \Psem{\Testt a}\Prom{\left.\vec\zeta\right.}\Prom{\Psem{M_2}}$. Yet, $\Psem{\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\Vect\zeta}}}}{\STOP{M_i}}}= \Psem{\Testt a}\Prom{\left.\vec\zeta\right.}({\Psem{M_i}})$. By Theorem~\ref{th:rel-ad-lemma}, we get that $\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\Vect\zeta}}}}{\STOP{M_1}}$ and $\LAPP{\LAPP{\Testt a}{\STOP{\Ran{\Vect\zeta}}}}{\STOP{M_2}}$ converge to $\ONELEM$ with different probabilities. It follows that $M_1\Rel{\not\Obseq}M_2$. \end{proof} \subsection*{Acknowledgements} We would like to thank the referees for their many useful and constructive comments and suggestions. We are also grateful to Michele Pagani and Rapha\"elle Crubill\'e for numerous and deep discussions. This work has been partly funded by the ANR Project RAPIDO ANR-14-CE25-0007, by the French-Chinese project ANR-11-IS02-0002 and NSFC 61161130530 \emph{Locali}. \bibliographystyle{plain}
c215d4ec12df6b63836dbe479061476b7ee63b5b
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} Lately there was a growing interest in studying self-similarity and fractal properties of graphs, which is largely inspired by applications in biology, sociology and chemistry \cite{song2005self,shanker2007defining}. Such studies often employ statistical physics methods that use ideas from graph theory and general topology, but are not intended to approach the problems under consideration in a rigorous mathematical way. Several studies that translate certain notions of topological dimension theory to graphs using combinatorial methods are also known \cite{smyth2010topological,evako1994dimension}. However, to the best of our knowledge a rigorous combinatorial theory that defines and studies graph-theoretical analogues of topological fractals still has not been developed. In this paper we introduce and study graph analogues of Lebesgue and Hausdorff dimensions of topological spaces from the graph-theoretical point of view. We show that they are closely related to well-known graph characteristics such as rank dimension \cite{berge1984hypergraphs} and Prague (or Ne{\v s}et{\v r}il-R{\"o}dl) dimension \cite{hell2004graphs}. It allowed us to define fractal graphs and determine fractality of some graph classes. In particular, it occurred that fractal graphs in some sense could be considered as generalizations of class 2 graphs. We demonstrate that various properties of dimensions of compact topological spaces have graph-theoretical analogues. Moreover, we also show how these relations allow for reverse transfer of combinatorial results to the general topology by proving a new property of general compact metric spaces using machinery from theory of hypergraphs. Finally, we show how Hausdorff (and Prague) dimension of graphs is related to Kolmogorov complexity. This relation allowed us to find a lower bound for Prague dimension for almost all graphs using incompressibility method from the theory of Kolmogorov complexity. \section{Basic definitions and facts from measure theory, dimension theory and graph theory} \label{sec:examples} Let $X$ be a compact metric space. A family $\mathcal{C} = \{C_{\alpha} : \alpha \in A\}$ of open subsets of $X$ is a {\it cover}, if $X = \bigcup_{\alpha \in A} C_{\alpha}$. A cover $\mathcal{C}$ is {\it $k$-cover}, if every $x\in X$ belongs to at most $k$ sets from $\mathcal{C}$; {\it $\epsilon$-cover}, if for every set $C_i\in \mathcal{C}$ its diameter $diam(C_i)$ does not exceed $\epsilon$; $(\epsilon, k)$-cover, if it is both $\epsilon$-cover and $k$-cover. {\it Lebesgue dimension} ({\it cover dimension}) $dim_L(X)$ of the space $X$ is the minimal integer $k$ such that for every $\epsilon > 0$ there exists $(\epsilon, k+1)$-cover of $X$. Let $\mathcal{F}$ be a semiring of subsets of a set X. A function $m:\mathcal{F}\rightarrow \mathbb{R}^+_0$ is {\it a measure}, if $m(\emptyset) = 0$ and for any disjoint sets $A,B\in \mathcal{F}$ $m(A\cup B) = m(A) + m(B)$. Let now $X$ be a subspace of an Euclidean space $\mathbb{R}^d$. {\it Hyper-rectangle} $R$ is a Cartesian product of semi-open intervals: $R = [a_1,b_1)\times\dots\times [a_d,b_d)$, where $a_i < b_i$, $a_i,b_i \in \mathbb{R}$; the {\it volume} of a the hyper-rectangle $R$ is defined as $vol(R) = \prod_{i=1}^d (b_i - a_i)$. The {\it $d$-dimensional Jordan measure} of the set $X$ is the value $\mathcal{J}^d(X) = \inf\{\sum_{R\in \mathcal{C}} vol(R)\},$ where infimum is taken over all finite covers $\mathcal{C}$ of $X$ by disjoint hyper-rectangles. The {\it $d$-dimensional Lebesgue measure} of a measurable set $\mathcal{L}^d(X)$ is defined analogously, with the additional condition that infimum is taken over all countable covers $\mathcal{C}$ of $X$ by (not necessarily disjoint) hyper-rectangles. Let $s >0$ and $\epsilon > 0$. Consider the parameter $\mathcal{H}^s_{\epsilon}(X) = \inf\{\sum_{C\in \mathcal{C}}diam(C)^s\},$ where infimum is taken over all $\epsilon$-covers of $X$. The {\it $s$-dimensional Hausdorff measure} of the set $X$ is defined as $\mathcal{H}^s(X) = \lim_{\epsilon \rightarrow 0} \mathcal{H}^s_{\epsilon}(X)$. The aforementioned measures are related as follows. If Jordan measure of the set $X$ exists, then it is equal to its Lebesgue measure. For Borel sets Lebesgue measure and Hausdorff measure are equivalent in the sense, that for any Borel set $Y$ we have $\mathcal{L}^d(Y) = C_d\mathcal{H}^d(Y)$, where $C_d$ is a constant depending only on $d$. {\it Hausdorff dimension} $dim_H(X)$ of the set $X$ is the value \begin{equation}\label{hdimtop} dim_H(X) = \inf\{s \geq 0 :\mathcal{H}^s(X) < \infty\}. \end{equation} Lebesgue and Hausdorff dimension of $X$ are related as follows: \begin{equation}\label{lebvshaus} dim_L(X) \leq dim_H(X). \end{equation} The set $X$ is {\it a fractal} \cite{falconer2004fractal}, if the inequality (\ref{lebvshaus}) is strict. Let $G=(V(G),E(G))$ be a simple graph. The family of subgraphs $\mathcal{C} = \{C_1,...,C_m\}$ of $G$ is a {\it cover}, if every edge $uv\in E(G)$ belongs to at least one subgraph from $\mathcal{C}$. A cover $\mathcal{C}$ is {\it $k$-cover}, if every vertex $v\in V(G)$ belongs to at most $k$ subgraphs of $\mathcal{C}$; {\it clique cover}, if all subgraphs $C_i$ are cliques. A set $W\subseteq V(G)$ {\it separates} vertices $u,v\in V(G)$, if $|W\cap \{u,v\}| = 1$. For a hypergraph $\mathcal{H} = (\mathcal{V}(\mathcal{H}),\mathcal{E}(\mathcal{H}))$, its {\it rank} $r(\mathcal{H})$ is the maximal size of its edges. A hypergraph $\mathcal{H}$ is {\it strongly $k$-colorable}, if for every vertex a color from the set $\{1,...,k\}$ can be assigned in such a way that vertices of every edge receive different colors. Intersection graph $L = L(\mathcal{H})$ of a hypergraph $\mathcal{H}$ is a simple graph with a vertex set $V(L) = \{v_E : E\in \mathcal{E}(\mathcal{H})\}$ in a bijective correspondence with the edge set of $\mathcal{H}$ and two distinct vertices $v_E,v_F\in V(L)$ being adjacent, if and only if $E\cap F \ne \emptyset$. The following theorem establishes a connection between intersection graphs and clique $k$-covers: \begin{theorem}\cite{berge1984hypergraphs}\label{thm:covervsinter} A graph $G$ is an intersection graph of a hypergraph of rank $\leq k$ if and only if it has a clique $k$-cover. \end{theorem} {\it Rank dimension} \cite{metelsky2003} $dim_R(G)$ of a graph $G$ is the minimal $k$ such that $G$ satisfies conditions of Theorem \ref{thm:covervsinter}. In particular, graphs with $dim_R(G)=1$ are disjoint unions of cliques (such graphs are called {\it equivalence graphs} \cite{alon1986cover} or {\it $M$-graphs} \cite{tyshkevich1989matr}), and graphs with $dim_R(G)=2$ are line graphs of multigraphs. {\it Categorical product} of graphs $G_1$ and $G_2$ is the graph $G_1 \times G_2$ with the vertex set $V(G_1\times G_2) = V(G_1)\times V(G_2)$ with two vertices $(u_1,u_2)$ and $(v_1,v_2)$ being adjacent whenever $u_1v_1\in E(G_1)$ and $u_2v_2\in E(G_2)$. {\it Prague dimension} $dim_P(G)$ is the minimal integer $d$ such that $G$ is an induced subgraph of a categorical product of $d$ complete graphs. {\it Equivalent cover} $\mathcal{M} = \{M_1,...,M_k\}$ of the graph $G$ consists of spanning subgraphs such that each subgraph $M_i$ is an equivalence graph. Equivalent cover is {\it separating}, if every two distinct vertices of $G$ are separated by one of connected components in some subgraph from $\mathcal{M}$. Relations between Prague dimension, clique covers, vertex labeling and intersection graphs are described by the following theorem: \begin{theorem}\cite{hell2004graphs,babaits1996kmern}\label{thm:pdimcharact} The following statements are equivalent: 1) $dim_P(\overline{G}) \leq k$; 2) there exists a separating equivalent cover of $G$; 3) $G$ is an intersection graph of strongly $k$-colorable hypergraph without multiple edges; 4) there exists an injective mapping $\phi: V(G) \rightarrow \mathbb{N}^k$, $v \mapsto (\phi_1(v),\dots,\phi_k(v))$ such that $uv\not\in E(G)$ whenever $\phi_j(u) \ne \phi_j(v)$ for every $j=1,...,k$. \end{theorem} Numbers of vertices and edges of a graph $G$ are denoted by $n$ and $m$, respectively. A subgraph of $G$ induced by vertex subset $U\subseteq V(G)$ is denoted as $G[U]$. For two graphs $G_1$ and $G_2$ the notation $G_1 \leq G_2$ indicates, that $G_1$ is an induced subgraph of $G_2$. \section{Lebesgue dimension of graphs}\label{lebgraph} Lebesgue dimension of a metric space is defined through $k$-covers by sets of arbitrary small diameter. It is natural to transfer this definition to graphs using graph $k$-covers by subgraphs of smallest possible diameter, i.e. by cliques. Thus by Theorem \ref{thm:covervsinter} we define Lebesgue dimension of a graph through its rank dimension: \begin{equation}\label{kdimeqleb} dim_L(G) = dim_R(G)-1. \end{equation} An analogy between Lebesgue and rank dimensions is futher justified by the following Proposition \ref{topspacehyper}, that states that any compact metric spaces of bounded Lebesgue measure could be approximated by intersection graphs of (infinite) hypergraphs of bounded rank. To prove it, we will use the following fact: \begin{lemma}\label{lebnumber}\cite{edgar2007measure} Let $X$ be a compact metric space and $\mathcal{U}$ be its open cover. Then there exists $\delta>0$ (called a {\it Lebesgue number} of $\mathcal{U}$) such that for every subset $A\subseteq X$ with $diam(A) < \delta$ there is a set $U\in \mathcal{U}$ such that $A\subseteq U$. \end{lemma} \begin{proposition}\label{topspacehyper} Let $X$ be a compact metric space with a metric $\rho$. Then $dim_L(X) \leq k-1$ if and only if for any $\epsilon > 0$ there exists a number $0 < \delta < \epsilon$ and a hypergraph $\mathcal{H}(\epsilon)$ on a finite vertex set $V(\mathcal{H}(\epsilon))$ with an edge set $E(\mathcal{H}(\epsilon)) = \{e_x : x\in X\}$ with the following properties: 1) $rank(\mathcal{H}(\epsilon)) \leq k$; 2) $e_x\cap e_y \ne \emptyset$ for every $x,y\in X$ such that $\rho(x,y) < \delta$; 3) $\rho(x,y) < \epsilon$ for every $x,y\in X$ such that $e_x \cap e_y \ne \emptyset$; 4) for every $v\in V(\mathcal{H}(\epsilon))$ the set $X_v = \{x\in X : v\in e_x \}$ is open. \end{proposition} \begin{proof} The proof borrows some ideas from intersection graphs theory (see \cite{berge1984hypergraphs}). Suppose that $dim_L(X) \leq k$, $\epsilon > 0$ and let $\mathcal{C}$ be the corresponding $(\epsilon,k)$-cover of $X$. Since $X$ is compact, we can assume that $\mathcal{C}$ is finite, i.e. $\mathcal{C} = \{C_1,...,C_m\}$. Let $\delta$ be the Lebesgue number of $\mathcal{C}$. For a point $x\in X$ let $e_x = \{i\in [m] : x\in C_i\}$. Consider a hypergraph $\mathcal{H}$ with $V(\mathcal{H}) = [m]$ and $E(\mathcal{H}) = \{e_x : x\in X\}$. Then $\mathcal{H}$ satisfies conditions 1)-4). Indeed, $rank(\mathcal{H}) \leq k$, since $\mathcal{C}$ is $k$-cover. If $\rho(x,y) < \delta$, then by Lemma \ref{lebnumber} there is $i\in [m]$ such that $\{x,y\}\in C_i$, i.e. $i\in e_x\cap e_y$. Condition $j\in e_x\cap e_y$ means that $x,y\in C_j$, and so $\rho(x,y) < \epsilon$, since $diam(C_j) < \epsilon$. Finally, for every $i\in V(\mathcal{H})$ we have $X_v = C_v$, and thus $X_v$ is open. Conversely, let $\mathcal{H}$ be a hypergraph with $V(\mathcal{H}) = [m]$ satisfying conditions (1)-(4). Then it is straightforward to checked, that $\mathcal{C} = \{X_1,...,X_m\}$ is an open $(\epsilon,k)$-cover of $X$. \end{proof} So, $dim_L(X)\leq k$ whenever for any $\epsilon > 0$ there is a well-defined hypergraph $\mathcal{H}(\epsilon)$ of $rank(\mathcal{H}(\epsilon)) \leq k$ with edges in bijective correspondence with points of $X$ such that two points are close if and only if corresponding edges intersect. \section{Graph measure and Hausdorff dimension of graphs}\label{measuregraph} In order to rigorously define a graph analogue of Hausdorff dimension, we need to define first a corresponding measure. Note that in any meaningful finite graph topology every set is a Borel set. As mentioned above, for measurable Borel sets in $\mathbb{R}^n$ Jordan, Lebesgue and Hausdorff measures are equivalent. Thus further we will work with a graph analogue of Jordan measure. It is known, that every graph is isomorphic to an induced subgraph of a categorical product of complete graphs \cite{hell2004graphs}. Consider a graph $G$ embedded into a categorical product of $d$ complete graphs $K^1_{n_1}\times\dots \times K^d_{n_d}$. Without loss of generality we may assume that $n_1 = ... = n_d = n$, i.e. \begin{equation}\label{graphembed} G\cong G' \leq S = (K_n)^d, \end{equation} Suppose that $V(K^n) = \{1,\dots,n\}$. The graph $S$ will be referred to as {\it a space} of dimension $d$ and $G'$ as an embedding of $G$ into $S$. It is easy to see, that by definition every vertex $v\in S$ is a vector $v = (v_1,...,v_d)$ with $v_i\in [n]$, and two vertices $u$ and $v$ are adjacent in $S$ if and only if $v_r\ne u_r$ for every $r\in [d]$. {\it Hyper-rectangle} $R = R(J_1,...,J_d)$ is a subgraph of $S$, that is defined as follows: for every $i=1,...,d$ choose a non-empty subset $J_i \subseteq [n]$, then $R = K_n[J_1]\times\dots\times K_n[J_d]$. The {\it volume} of a hyper-rectangle $R$ is the value $vol(R) = |V(R)| = \prod_{i=1}^d |J_i|$. The family $\mathcal{R} = \{R^1,...,R^m\}$ of hyper-rectangles is a {\it rectangle co-cover} of $G'$, if the subgraphs $R^i$ are pairwise vertex-disjoint, $V(G')\subseteq \bigcup_{i=1}^m V(R^i)$ and $\mathcal{R}$ covers all non-edges of $G'$, i.e. for every $x,y\in V(G')$, $xy\not\in E(G')$ there exists $j\in[m]$ such that $x,y\in V(R^j)$. We define {\it $d$- volume} of a graph $G$ as \begin{equation}\label{grvol} vol^d(G) = \min_{G'}\min_{\mathcal{R}} \sum_{R\in \mathcal{R}} vol(R), \end{equation} where the first minimum is taken over all embeddings $G'$ of $G$ into $d$-dimensional spaces $S$ and the second minimum - over all rectangle co-covers of $G'$ (see Fig. \ref{fig:embedP4}). \begin{figure}[h] \begin{center} \includegraphics[width=80mm,height=80mm,keepaspectratio] {EmbedP4_1.pdf} \caption{\label{fig:embedP4} {\small Embedding of $G=P_4$ into a 2-dimensional space $S=(K_3)^2$ and its rectangle co-cover by a hyper-rectangle of volume $6$}} \end{center} \end{figure} We define a {\it $d$- measure} of a graph $F$ as follows: \begin{equation}\label{neasuregraph} \mathcal{H}^d(F) = \left \{ \begin{array}{lll} vol^d(\overline{F}), \text{ if } \overline{F} \text{ can be represented as (\ref{graphembed})}; \\ +\infty, \text{ otherwise } \end{array} \right. \end{equation} In the remaining part of this section we will prove, that $\mathcal{H}^d$ indeed satisfy the property of a measure, i.e. $\mathcal{H}^d(F^1\cup F^2) = \mathcal{H}^d(F^1) + \mathcal{H}^d(F^2)$, where $F^1\cup F^2$ is a disjoint union of graphs $F^1$ and $F^2$. Let $W^1,W^2\subseteq V(S)$. We write $W_1\sim W_2$, if every vertex from $W_1$ is adjacent to every vertex from $W_2$. Denote by $P_k(W_1)$ {\it the $k$-th projection} of $W_1$, i.e. the set of all $k$-coordinates of vertices of $W_1$: $$J_k(W_1) = \{v_k : v\in W_1\}.$$ In particular, $P_k(R(J_1,...,J_d)) = J_k$. The following proposition follows directly from the definition of $S$ \begin{proposition}\label{padjcoord} $W^1 \sim W^2$ if and only if $P_k(W_1)\cap P_k(W_2) = \emptyset$ for every $k\in [d]$. \end{proposition} Assume that $\mathcal{R} = \{R^1,...,R^m\}$ is a minimal rectangle co-cover of a minimal embedding $G'$, i.e. $vol^d(G) = \sum_{R\in \mathcal{R}} vol(R)$. Further we will demonstrate, that $\mathcal{R}$ has a rather simple structure. Let $J_k^i = P_k(R^i)$. \begin{proposition}\label{dvolumeinterindex} $J^i_k \cap J^j_k = \emptyset$ for every $i,j\in [m]$, $i\ne j$ and every $k\in [d]$ . \end{proposition} \begin{proof} First note, that for every hyper-rectangle $R^i = R(J^i_1,...,J^i_d)$, every coordinate $k\in [d]$ and every $l\in J^i_k$ there exists $v\in V(G')\cap V(R^i)$ such that $v_k = l$. Indeed, suppose that it does not hold for some $l\in J^i_k$. If $|J_k^i| = 1$, then it means that $V(G')\cap V(R^i) = \emptyset$. Thus, $\mathcal{R}' = \mathcal{R}\setminus \{R^i\}$ is a rectangle co-cover, which contradicts the minimality of $\mathcal{R}$. If $|J_k^i| > 1$, consider a hyper-rectangle $(R^i)' = R(J_i^1,...,J_k^i\setminus \{l\},...,J^i_d)$. The set $\mathcal{R}' = \mathcal{R}\setminus \{R^i\}\cup \{(R^i)'\}$ is a rectangle co-cover, and the $d$-volume of $(R^i)'$ is smaller than the $d$-volume of $R^i$. Again it contradicts minimality of $\mathcal{R}$. Now assume that for some distinct $i,j\in [m]$ and $k\in [d]$ we have $J^i_k \cap J^j_k \supseteq \{l\}$. Then there exist $u\in V(G')\cap V(R^i)$ and $v\in V(G')\cap V(R^j)$ such that $u_k = v_k = l$. So, $uv\not\in E(G')$, and therefore by the definition $uv$ is covered by some $R^h\in \mathcal{R}$. The hyper-rectangle $R^h$ intersects both $R^i$ and $R^j$, which contradicts the definition of a rectangle co-cover. \end{proof} \begin{proposition}\label{dvolumecoconcomp} Let $U^i = V(G')\cap V(R^i)$, $i=1,...,m$. Then the set $U=\{U^1,...,U^m\}$ coincides with the set of co-connected components of $G'$. \end{proposition} \begin{proof} Propositions \ref{padjcoord} and \ref{dvolumeinterindex} imply, that vertices of distinct hyper-rectangles from the co-cover $\mathcal{R}$ are pairwise adjacent. So, $U_i \sim U_j$ for every $i,j\in [m]$, $i\ne j$. Let $\mathcal{C}=\{C^1,...,C^r\}$ be the set of co-connected components of $G'$ (thus $V(G') = \bigsqcup_{l=1}^r C^l = \bigsqcup_{i=1}^m U^i$). Consider a component $C^l\in \mathcal{C}$ and the sets $C^l_i = C^l \cap U^i$, $i=1,...,m$. We have $C^l = \bigsqcup_{i=1}^m C^l_i$ and $C^l_i \sim C^l_j$ for all $i\ne j$. Therefore, due to co-connectedness of $C^l$, exactly one of the sets $C^l_i$ is non-empty. So, we have demonstrated, that every co-connected component $C^l$ is contained in some of the sets $U^i$. Now, let some $U_i$ consists of several components, i.e. without loss of generality $U_i = C^1\sqcup\dots\sqcup C^q$, $q\geq 2$. Let $I_k^j = P_k(C^j)$, $j=1,...,q$. By Proposition \ref{padjcoord} we have $I_k^{j_1} \cap I_k^{j_2} = \emptyset$ for all $j_1\ne j_2$, $k=1,...,d$. Consider hyper-rectangles $R^{i,1} = R(I_1^1,...,I_d^1)$,...,$R^{i,q} = R(I_1^q,...,I_d^q)$. Those hyper-rectangles are pairwise vertex-disjoint, and $C^j \subseteq V(R^{i,j})$ for all $j\in [q]$. Since every pair of non-connected vertices of $G'$ is contained in some of its co-connected components, we arrived to the conclusion, that the set $\mathcal{R}' = \mathcal{R}\setminus \{R^i\} \cup \{R^{i,1},...,R^{i,q}\}$ is a rectangle co-cover. Moreover, $V(R^{i,1})\sqcup\dots\sqcup V(R^{i,q}) \subsetneq V(R^i)$, and therefore $\sum_{j=1}^q vol(R^{i,j}) < vol(R^i)$, which contradicts the minimality of $\mathcal{R}$. \end{proof} \begin{corollary}\label{voljoin} Let $\mathcal{C}=\{C^1,...,C^m\}$ be the set of co-connected components of $G'$. Then $R_i = R(P_1(C^i),...,P_d(C^i))\supseteq C^i$. \end{corollary} \begin{proof} By Proposition \ref{dvolumecoconcomp}, $|\mathcal{R}| = |\mathcal{C}|$, and every component $C^i\in \mathcal{C}$ is contained in a unique hyper-rectangle $R^i\in \mathcal{R}$, $i=1,...,m$. Every pair of non-adjacent vertices of $G'$ belong to some of its co-connected components. This fact, together with the minimality of $\mathcal{R}$, implies that $R^i$ is the minimal hyper-rectangle that contains $C^i$. Thus $R_i = R(P_1(C^i),...,P_d(C^i))$. \end{proof} \begin{corollary}\label{cor:minvolspace} If $\overline{G}$ is connected, than $vol^d(G)$ is the minimal volume of $d$-dimensional space $S$, where $G$ can be embedded. \end{corollary} \begin{corollary}\label{volconcomp} Let $\mathcal{D}=\{D^1,...,D^m\}$ be the set of co-connected components of $G$. Then $vol^d(G) = \sum_{i=1}^m vol^d(G[D^i])$. \end{corollary} \begin{proof} Suppose that $\{C^1,...,C^m\}$ is the set of co-connected components of $G'$, and $G[D^i]\cong G'[C^i]$. Proposition \ref{dvolumecoconcomp} and Corollary \ref{voljoin} imply that $\{R_i\}$ is a rectangle co-cover of an embedding of $G[D^i]$. Therefore we have $vol(R^i) \geq vol^d(G[D^i])$ and thus $vol^d(G) = \sum_{i=1}^m vol(R^i) \geq \sum_{i=1}^m vol^d(G[D^i])$. Now let $G'^i$ be a minimal embedding of $G[D^i]$ into $(K_n)^d$. By Proposition \ref{dvolumecoconcomp}, every minimal hyper-rectangle co-cover of $G'^i$ consists of a single hyper-rectangle $R^i = R(J^i_1,...,J^i_d)$ or, in other words, $G[D^i]$ is embedded into $R^i$. Now construct an embedding $G'$ of $G$ into $S=(K_{nm})^d$ and its hyper-rectangle co-cover as follows: let $I^i_k = (i-1)m + J^i_k = \{(i-1)m + l : l\in J^i_k\}$, $k=1,...,d$. Obviously, $Q^i = K_{mn}(I^i_1)\times\dots\times K_{mn}(I^i_d) \cong R^i$. Now embed $G[D^i]$ into $Q^i$. Let $G'^i$ be those embeddings. By Proposition \ref{padjcoord} $V(G'^i) \sim V(G'^j)$, so $G' = G'^1\cup...\cup G'^m$ is indeed embedding of $G$. All $Q^i$ are pairwise disjoint. Since every pair of non-adjacent vertices belong to some $G[D^i]$, we have that $\mathcal{Q} = \{Q^1,...,Q^m\}$ is hyper-rectangle co-cover of $G'$. Therefore $vol^d(G) \leq \sum_{i=1}^m vol(Q^i) = \sum_{i=1}^m vol^d(G[D^i])$. \end{proof} \begin{proposition}\label{measureadditivity} Let $F_1$ and $F_2$ be two graphs. Then \begin{equation}\label{eq:measaddit} \mathcal{H}^d(F^1\cup F^2) = \mathcal{H}^d(F^1) + \mathcal{H}^d(F^2) \end{equation} \end{proposition} \begin{proof} It can be shown \cite{babai1992linear}, that $\overline{F^1\cup F^2}$ can be embedded into a categorical product of $d$ complete graphs if and only if both $\overline{F_1}$ and $\overline{F_2}$ have such embeddings. Therefore the relation (\ref{eq:measaddit}) holds, if one of its members is equal to $+\infty$. If all $\overline{F_1}$,$\overline{F_2}$, $\overline{F^1\cup F^2}$ can be embedded into a product of $d$ complete graphs, then (\ref{eq:measaddit}) follows from Corollary \ref{volconcomp}. \end{proof} Following the analogy with Hausdorff dimension of topological spaces (\ref{hdimtop}), we define a {\it Hausdorff dimension} of a graph $G$ as \begin{equation}\label{hdimgraph} dim_H(G) = \min\{s \geq 0 :\mathcal{H}^s(G) < \infty\}-1. \end{equation} Thus, Hausdorff dimension of a graph can be identified with a Prague dimension of its complement minus 1. \section{Relations with Kolmogorov complexity} Let $\mathbb{B}^*$ be the set of all finite binary strings and $\Phi:\mathbb{B}^* \rightarrow \mathbb{B}^*$ be a computable function. A {\it Kolmogorov complexity} $K_{\Phi}(s)$ of a binary string $s$ with respect to $\Phi$ is defined as a minimal length of a string $s'$ such as $\Phi(s') = s$. Since Kolmogorov complexities with respect to any two functions differ only by an additive constant \cite{li2009Kolmogorov}, it is usually assumed that some canonical function $\Phi$ is fixed, and Kolmogorov complexity is denoted simply by $K(s)$. Thus, informally $K(s)$ could be described as a length of a shortest encoding of the string $s$, that allows to completely reconstruct it. Analogously, for two strings $s,t\in \mathbb{B}^*$, a {\it conditional Kolmogorov complexity} $K(s|t)$ is a a length of a shortest encoding of $s$, if $t$ is known in advance. More information on properties of Kolmogorov complexity can be found in \cite{li2009Kolmogorov}. Every graph $G$ can be naturally encoded using the string representation of an upper triangle of its adjacency matrix. Kolmogorov complexity of a graph could be defined as a Kolmogorov complexity of that string \cite{mowshowitz2012entropy,buhrman1999kolmogorov}. It gives estimations $K(G) = O(n^2)$, $K(G|n) = O(n^2)$. Alternatively, $n$-vertex connected labeled graph can be represented as a list of edges with ends of each edge encoded using their binary representations concatenated with a binary representation of $n$. It gives estimations $K(G) \leq 2m\log(n) + \log(n) = O(m\log(n))$, $K(G|n) \leq 2m\log(n) = O(m\log(n))$ \cite{li2009Kolmogorov,mowshowitz2012entropy}. Further in this section for simplicity we will consider connected graphs (for disconnected graphs all considerations below could be applied to every connected component). Let $dim_H(G) = dim_P(\overline{G}) - 1 = d-1$ and $\mathcal{H}^d(G) = h$. Then by Corollary \ref{cor:minvolspace} $\overline{G}$ is an induced subgraph of a product \begin{equation}\label{graphembedmin} K_{p_1}\times\dots \times K_{p_d}, \end{equation} where $h = p_1\cdot\dots \cdot p_d$. So, by Theorem \ref{thm:pdimcharact}, $G$ and $\overline{G}$ could be encoded using a collection of vectors $\phi(v) = (\phi_1(v),\dots,\phi_d(v))$, $v\in V(G)$, $\phi_j(v)\in [p_j]$. Such encoding could be stored as a string containing binary representations of coordinates $\phi_j(v)$ using $\log(p_j)$ bits concatenated with a binary representations of $n$ and $p_j$, $j=1,...,n$. The length of this string is $(n+1)\sum_{j=1}^d \log(p_j) + \log(n)$. Analogously, if $n$ and $p_j$ are given, then the length of encoding is $n\sum_{j=1}^d \log(p_j)$. Thus, the following estimation is true: \begin{proposition} For any $G$, \begin{equation}\label{eq:KolG} K(G)\leq (n+1)\log(\mathcal{H}^d(G)) + \log(n) \end{equation} \begin{equation}\label{eq:KolcondG} K(G|n,p_1,...,p_d)\leq n\log(\mathcal{H}^d(G)) \end{equation} \end{proposition} Let $p^* = \max_j p_j$. Then we have $K(G)\leq (n+1)d\log(p^*) + \log(n),$\newline $K(G|n,p_1,...,p_d)\leq nd\log(p^*).$ By minimality of the representation (\ref{graphembedmin}), we have $p^* \leq n$. Thus $K(G) = O(dn\log(n))$, $K(G|n,p_1,...,p_d) = O(dn\log(n))$. So, Hausdorff (and Prague) dimension could be considered as a measure of descriptive complexity of a graph. In particular, for graphs with a small Hausdorff dimension (\ref{eq:KolG})-(\ref{eq:KolcondG}) give better estimation of their Kolmogorov complexity than the standard estimations mentioned above. Relations between Hausdorff (Prague) dimension and Kolmogorov complexity could be used to derive lower bound for Hausdorff (and Prague) dimension in a typical case. More rigorously, let $X$ be a graph property and $\mathcal{P}_n(X)$ be the set of labeled n-vertex graphs having $X$. Property $X$ holds for {\it almost all graphs} \cite{erdHos1977chromatic}, if $|\mathcal{P}_n(X)|/2^{\binom{n}{2}} \rightarrow 1$ as $n \rightarrow \infty$. We will use the following lemma: \begin{lemma}\label{lem:kolcomplbgr}\cite{buhrman1999kolmogorov} For every $n > 0$ and $\delta: \mathbb{N}\rightarrow \mathbb{N}$, there are at least $2^{\binom{n}{2}}(1-2^{-\delta(n)})$ $n$-vertex labeled graphs $G$ such that $K(G|n) \geq \frac{n(n-1)}{2} - \delta(n)$. \end{lemma} Then the following theorem is true: \begin{theorem}\label{thm:almostalldim} For every $\epsilon > 0$, almost all graphs have Prague dimension $d$ such that \begin{equation}\label{lowbounddim} d \geq \frac{1}{1+\epsilon}\Big(\frac{n-1}{2\log(n)} - \frac{1}{n}\Big) \end{equation} \end{theorem} \begin{proof} Let $n_{\epsilon} = \ceil{\frac{2}{\epsilon}}$. Consider a graph $G$ with $n\geq n_{\epsilon}$. From (\ref{eq:KolcondG}) we have $K(G)\leq (n+1)d\log(n) + \log(n)$. Using the fact, that $\frac{1}{n} + \frac{1}{nd}\leq \frac{2}{n}\leq \epsilon$, it is straightforward to check that $(n+1)d\log(n) + \log(n) \leq (1+\epsilon)nd\log(n)$. Therefore we have \begin{equation}\label{eq:kolmepsilon} K(G)\leq (1+\epsilon)nd\log(n) \end{equation} Let $X$ be the set of all graphs $G$ such that \begin{equation}\label{eq:kolmcondlog} K(G|n) \geq \frac{n(n-1)}{2} - \log(n) \end{equation} Using Lemma \ref{lem:kolcomplbgr} with $\delta(n) = \log(n)$, we conclude that $|\mathcal{P}_n(X)|/2^{\binom{n}{2}} \geq 1-\frac{1}{n}$, and so almost all graphs have the property $X$. Now it is easy to see that for graphs with the property $X$ and with $n\geq n_{\epsilon}$ the inequality (\ref{lowbounddim}) holds. It follows by combining inequalities (\ref{eq:kolmepsilon})-(\ref{eq:kolmcondlog}) using the fact that $K(G|n)\leq K(G)$. It concludes the proof. \end{proof} Considerations above and proof of Theorem \ref{thm:almostalldim} imply that for every $\epsilon > 0$ and $n$-vertex graph $G$ with sufficiently large $n$, $dim_H(G) \geq \frac{1}{(1+\epsilon)\log(p^*)}\frac{K(G)}{n}$. Interestingly, similar relations between Kolmogorov complexity and analogues of Hausdorff dimension hold for other objects. In particular, for Cantor space $\mathcal{C}$ (the space of all infinite 0-1 sequences) it is proved in \cite{mayordomo2002kolmogorov} that Kolmogorov complexity and effective (or constructive) Hausdorff dimension $dim_H(s)$ of each sequence $s$ are related as follows: $dim_H(s)= \liminf\limits_{n \rightarrow \infty} \frac{K(s_n)}{n}$ (here $s_n$ is the prefix of $s$ of length $n$). Similar estimations are known for other variants of Hausdorff dimension \cite{ryabko1994complexity,staiger1993kolmogorov}. \section{Fractal graphs}\label{fracgraph} Importantly, the relation (\ref{lebvshaus}) between Lebesgue and Hausdorff dimensions of topological spaces remains true for graphs. \begin{proposition}\label{lebvshausgraph} For any graph $G$ \begin{equation} dim_R(G) - 1 = dim_L(G) \leq dim_H(G) = dim_P(\overline{G})-1. \end{equation} \end{proposition} \begin{proof} Let Prague dimension of a graph $\overline{G}$ is equal to $k$. Then by Theorem \ref{thm:pdimcharact} $G$ is an intersection graph of strongly $k$-colorable hypergraph. Since rank of every such hypergraph obviously does not exceed $k$, Theorem \ref{thm:covervsinter} implies, that $dim_R(G)\leq k$. \end{proof} Definitions of dimensions immediately imply, that both Lebesgue and Hausdorff dimensions are monotone with respect to induced subgraphs. Analogously to the definition of fractals for topological spaces, we say, that a graph $G$ is a {\it fractal}, if $dim_L(G) < dim_H(G)$, i.e. $dim_R(G) < dim_P(\overline{G})$. The following proposition provides a first non-trivial example of fractal graphs \begin{proposition} Triangle-free fractals are exactly triangle-free graphs of class 2 \end{proposition} \begin{proof} Note that for triangle-free graphs $dim_R(G) = \Delta(G)$. Moreover, it can be shown (see \cite{hell2004graphs}) that if triangle-free graph $G$ is not a disjoint union of edges $nK_2$, then $dim_P(\overline{G}) = \chi'(G)$, where $\chi'(G)$ is a chromatic index. Therefore triangle-free fractals, that are not disjoint unions of edges, are exactly triangle-free graphs of class 2. On the other hand, all bipartite graphs except $nK_2$ are not fractals, since they all are traingle-free graphs of class 1. \end{proof} As another example, consider Sierpinski gasket graphs $S_n$ \cite{teguia06serp,klavzar2008coloring,teufl2006spanning}. This class of graphs is associated with the Sierpinski gasket - well-known topological fractal with a Hausdorff dimension $\log(3)/\log(2)\approx 1.585$. Edges of $S_n$ are line segments of the $n$-th approximation of the Sierpinski gasket, and vertices are intersection points of these segments (Fig. \ref{fig:serpGasket}). Sierpinski gasket graphs can be defined recursively as follows. Consider tetrads $T_n = (S_n,x_1,x_2,x_3)$, where $x_1,x_2,x_3$ are distinct vertices of $S_n$ called {\it contact vertices}. The first Sierpinski gasket graph $S_1$ is a triangle $K_3$ with vertices $x_1,x_2,x_3$, the first tetrad is defined as $T_1 = (S_1,x_1,x_2,x_3)$. The $(n+1)$-th Sierpinski gasket graph $S_{n+1}$ is constructed from 3 disjoint copies $(S_n,x_1,x_2,x_3)$, $(S'_n,x'_1,x'_2,x'_3)$, $(S''_n,x''_1,x''_2,x''_3)$ of $n$-th tetrad $T_n$ by gluing together $x_2$ with $x'_1$, $x'_3$ with $x''_2$ and $x_3$ with $x''_1$; the corresponding $(n+1)$-th tetrad is $T_{n+1} = (S_{n+1},x_1,x'_2,x''_3)$. \begin{proposition}\label{serpgasket} For every $n\geq 2$ Sierpinski gasket graph $S_n$ is a fractal with $dim_L(S_n) = 1$ and $dim_H(S_n) = 2$ \end{proposition} \begin{proof} First, we will prove that \begin{equation}\label{formula:dimssn} dim_R(S_n)\leq 2, dim_P(\overline{S_n})\leq 3. \end{equation} for every $n\geq 2$. We will show it using an induction by $n$. In fact, we will prove slightly stronger fact: for any $n\geq 2$ there exists a clique cover $\mathcal{C} = \{C_1,...,C_m\}$ such that (i) every non-contact vertex is covered by two cliques from $\mathcal{C}$; (ii) every contact vertex is covered by one clique from $\mathcal{C}$; (iii) cliques from $\mathcal{C}$ can be colored using 3 colors in such a way, that intersecting cliques receive different colors and cliques containing different contact vertices also receive different colors; (iv) every two distinct vertices are separated by some clique from $\mathcal{C}$. For $n=2$ the clique cover $\mathcal{C}$ consisting of 3 cliques, that contain contact vertices, obviously satisfies conditions (i)-(iv) (see Fig. \ref{fig:serpGasket}). Now suppose that $\mathcal{C}$, $\mathcal{C'}$ and $\mathcal{C''}$ are clique covers of $S_n$, $S'_n$ and $S''_n$ with properties (i)-(iv). Assume that $x_i\in C_i$, $x'_i\in C'_i$. $x''_i\in C''_i$ and $C_i,C'_i,C''_i$ have colors $i$, $i=1,...,3$. Then it is straightforward to check, that $\mathcal{C}\cup \mathcal{C'}\cup \mathcal{C''}$ with all cliques keeping their colors is a clique cover of $S_{n+1}$ that satisfies (i)-(iv). So, (\ref{formula:dimssn}) is proved. Finally, note that for every $n\geq 2$ the graph $S_n$ contains graphs $K_{1,2}$ and $K_4 - e$ as induced subgraphs. Since $dim_R(K_{1,2}) = 2$ and $dim_P(\overline{K_4 - e}) = 3$ (the later is easy to see using clique cover formulation), we have equalities in (\ref{formula:dimssn}) \end{proof} \begin{figure}[h] \begin{center} \includegraphics[width=\textwidth,height=\textheight,keepaspectratio] {SierpGasket.pdf} \caption{\label{fig:serpGasket} {\small Sierpinski gasket graphs $S_1-S_4$}} \end{center} \end{figure} It is important to emphasize that Lebesgue and Hausdorff dimensions of Sierpinski gasket graphs agree with the corresponding dimensions of Sierpinski gasket fractal. As a contrast, note that rectangular grid graphs (cartesian products of 2 paths) are not fractals (by K{\"o}nig theorem since they are bipartite), just like rectangles are not fractals in $\mathbb{R}^2$. \begin{comment} So, separation of graphs into fractals and non-fractals in some sense generalizes separation onto class I and class II. For triangle-free graphs the difference between Lebesgue and Hausdorff dimensions does not exceed 1. The natural question to ask is how different can be Lebesgue dimension and Hausdorff dimension of general graphs. We will try to answer this question in the rest of this section. Note first, that equivalence number and complement Prague dimension??? are close to each other. \begin{observation}\label{obs:eqvshaus} For any $n$-vertex graph $G$ $eq(G) \leq dim_P(\overline{G}) \leq eq(G)+1$ \end{observation} \begin{proof} The inequalities follow from second characterization of Theorem \ref{thm:pdimcharact}. Left inequality directly follows from the definition. If $\mathcal{M} = \{M_1,...,M_r\}$ is an equivalent cover of $G$, which is not separating, then $M\cup \{O_n\}$ is a separating equivalent cover of $G$. It proves the second inequality. \end{proof} In light of Observation \ref{obs:eqvshaus} we will further consider relations between weak krausz dimension and equivalence number. Let $dim_R(G) = k$, which, by Theorem \ref{thm:covervsinter}, implies that $G$ is a line graph of a $k$-uniform hypergraph $\mathcal{H}$. Obviously, $eq(G) \leq \chi_s(\mathcal{H})$, where $\chi_s(\mathcal{H})$ is a strong chromatic number of $\mathcal{H}$. This upper bound can be an overestimation. However, we will show that $eq(G)$ can be bounded above by a function of $\chi_s(\mathcal{H})$. To do it, we will generalize a construction proposed in \cite{esperet2010cover}. Obviously, every clique in $G$ corresponds to a family of pairwise intersecting edges of $\mathcal{H}$ ({\it intersecting family} \cite{berge1984hypergraphs}). Therefore, each spanning equivalence subgraph of $G$ corresponds to a partition of $\mathcal{E}(H)$ into intersecting families ({\it intersecting partition}). Equivalent cover of $G$ transforms into a family $\mathcal{M}$ of intersecting partitions of $\mathcal{H}$ such that every pair of intersecting hyperedges belongs to some intersecting family from one of the partitions from $\mathcal{M}$. The family $\mathcal{M}$ will be called an {\it intersecting cover} of $\mathcal{H}$. We will consider special type of intersecting covers, where each intersecting family $S=\{E_1,...,E_p\}$ is required to have {\it Helly property}, i.e. $\bigcap_{j=1}^p E_j \ne \emptyset$ \cite{berge1984hypergraphs}. Formally, let $v\in \mathcal{V}\mathcal{H}$. The family $S(v)=\{E_1,...,E_p\}$ is a {\it Helly family}, if either it is empty or $v\in \bigcap_{j=1}^p E_j$. {\it Helly partition} is a collection $\mathcal{M}^i$ of disjoint Helly families $\mathcal{M}_i = \{S^i_v : v\in \mathcal{V}\mathcal{H}\}$ such that $\bigcup_{v\in \hypvert{H}} S^i(v) = \hypedge{H}$. {\it Helly cover} of size $r$ is a set $\mathcal{M} = \{\mathcal{M}_1,...,\mathcal{M}_r\}$ of Helly partitions such that for every intersecting pair of edges $E_1,E_2\in \hypedge{H}$ there exists $i\in [r]$ and $v\in E_1\cap E_2$ such that $E_1,E_2\in S^i(v)$. Finally, $\sigma(\mathcal{H})$ is the minimal size of Helly cover of $\mathcal{H}$. Obviously, every Helly cover is an intersecting cover, therefore \begin{equation}\label{eqvssigma} eq(G) \leq \sigma(\mathcal{H}) \end{equation} In \cite{esperet2010cover}, the notion of {\it orientation cover} of a simple graph $H$ was introduced. Let $\mathcal{O}=\{\overrightarrow{H}^1,...,\overrightarrow{H}^r\}$ be a set of orientations of $H$. This set is an orientation cover, if for every pair of edges $uv,uw\in E(H)$ there exists an orientation $\overrightarrow{H}^i$ such that $\overrightarrow{uv},\overrightarrow{uw} \in E(\overrightarrow{H}^i)$. It is easy to see that for simple graphs Helly covers and orientation covers are equivalent: every orientation $\overrightarrow{H}^i$ can be transformed into Helly partition $M^i$ by setting $C^i_v = \{vw : \overrightarrow{vw}\in E(\overrightarrow{H}_i)\}$, and vise versa. It was proved in \cite{esperet2010cover}, that for a simple graph $H$ \begin{equation}\label{sigmavschisimpgraph} \sigma(H) \leq 2 \lceil\log_2\log_2 \chi(H)\rceil + 2 \end{equation} \end{comment} \begin{comment} Two vertices $u,v\in V(G)$ are called {\it twins}, if $N[u] = N[v]$, where $N[u]$ is a closed neighborhood of $u$. For such graphs $dim_P(\overline{G}) = eq(G)$. Graphs without twins with $dim_R \leq 2$ are line graphs. For such graphs, equivalence number was studied in \cite{esperet2010cover}. The main result obtained in \cite{esperet2010cover} is the following theorem. \begin{theorem}\cite{esperet2010cover}\label{thm:eqline} Let $G$ is a line graph of a simple graph $G'$. Then \begin{equation} \frac{1}{3}\log_2\log_2 \chi(G') \leq eq(G) \leq 2\log_2\log_2\chi(G') + 2 \end{equation} \end{theorem} So, for any graph without twins with Lebesgue dimension 1 its Hausdorff dimension can be arbitrarily large, although it is bounded above by \end{comment} \section{Acknowlegements} The work of LB was partially supported by the NSF grant DMS-1600568 \bibliographystyle{plain}
b391fb6d99f4e356978b4e0f0a2d74dafb793c77
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{% \@startsection {section}% {1}% {\z@}% {0.8cm \@plus1ex \@minus .2ex}% {0.5cm}% {% \normalfont\small\bfseries \centering }% }% \def\@hangfrom@section#1#2#3{\@hangfrom{#1#2}\MakeTextUppercase{#3}}% \def\subsection{% \@startsection {subsection}% {2}% {\z@}% {.8cm \@plus1ex \@minus .2ex}% {.5cm}% {% \normalfont\small\bfseries \centering }% }% \def\subsubsection{% \@startsection {subsubsection}% {3}% {\z@}% {.8cm \@plus1ex \@minus .2ex}% {.5cm}% {% \normalfont\small\itshape \centering }% }% \def\paragraph{% \@startsection {paragraph}% {4}% {\parindent}% {\z@}% {-1em}% {\normalfont\normalsize\itshape}% }% \def\subparagraph{% \@startsection {subparagraph}% {5}% {\parindent}% {3.25ex \@plus1ex \@minus .2ex}% {-1em}% {\normalfont\normalsize\bfseries}% }% \def\section@preprintsty{% \@startsection {section}% {1}% {\z@}% {0.8cm \@plus1ex \@minus .2ex}% {0.5cm}% {% \normalfont\small\bfseries }% }% \def\subsection@preprintsty{% \@startsection {subsection}% {2}% {\z@}% {.8cm \@plus1ex \@minus .2ex}% {.5cm}% {% \normalfont\small\bfseries }% }% \def\subsubsection@preprintsty{% \@startsection {subsubsection}% {3}% {\z@}% {.8cm \@plus1ex \@minus .2ex}% {.5cm}% {% \normalfont\small\itshape }% }% \@ifxundefined\frontmatter@footnote@produce{% \let\frontmatter@footnote@produce\frontmatter@footnote@produce@endnote }{}% \def\@pnumwidth{1.55em} \def\@tocrmarg {2.55em} \def\@dotsep{4.5pt} \setcounter{tocdepth}{3} \def\tableofcontents{% \addtocontents{toc}{\string\tocdepth@munge}% \print@toc{toc}% \addtocontents{toc}{\string\tocdepth@restore}% }% \def\tocdepth@munge{% \let\l@section@saved\l@section \let\l@section\@gobble@tw@ }% \def\@gobble@tw@#1#2{}% \def\tocdepth@restore{% \let\l@section\l@section@saved }% \def\l@part#1#2{\addpenalty{\@secpenalty}% \begingroup \set@tocdim@pagenum{#2}% \parindent \z@ \rightskip\tocleft@pagenum plus 1fil\relax \skip@\parfillskip\parfillskip\z@ \addvspace{2.25em plus\p@}% \large \bf % \leavevmode\ignorespaces#1\unskip\nobreak\hskip\skip@ \hb@xt@\rightskip{\hfil\unhbox\z@}\hskip-\rightskip\hskip\z@skip \par \nobreak % \endgroup }% \def\tocleft@{\z@}% \def\tocdim@min{5\p@}% \def\l@section{% \l@@sections{}{section }% \def\l@f@section{% \addpenalty{\@secpenalty}% \addvspace{1.0em plus\p@}% \bf }% \def\l@subsection{% \l@@sections{section}{subsection }% \def\l@subsubsection{% \l@@sections{subsection}{subsubsection }% \def\l@paragraph#1#2{}% \def\l@subparagraph#1#2{}% \let\toc@pre\toc@pre@auto \let\toc@post\toc@post@auto \def\listoffigures{\print@toc{lof}}% \def\l@figure{\@dottedtocline{1}{1.5em}{2.3em}} \def\listoftables{\print@toc{lot}}% \let\l@table\l@figure \appdef\class@documenthook{% \@ifxundefined\raggedcolumn@sw{\@booleantrue\raggedcolumn@sw}{}% \raggedcolumn@sw{\raggedbottom}{\flushbottom}% }% \def\tableft@skip@float{\z@ plus\hsize}% \def\tabmid@skip@float{\@flushglue}% \def\tabright@skip@float{\z@ plus\hsize}% \def\array@row@pre@float{\hline\hline\noalign{\vskip\doublerulesep}}% \def\array@row@pst@float{\noalign{\vskip\doublerulesep}\hline\hline}% \def\@makefntext#1{% \def\baselinestretch{1}% \reset@font \footnotesize \leftskip1em \parindent1em \noindent\nobreak\hskip-\leftskip \hb@xt@\leftskip{% \Hy@raisedlink{\hyper@anchorstart{footnote@\the\c@footnote}\hyper@anchorend}% \hss\@makefnmark\ }% #1% \par }% \prepdef \section{INTRODUCTION} \label{sec:introduction} The growth of structure across cosmic time provides an important probe of cosmology. The evolution of the galaxy cluster mass function with redshift places significant constraints on the dark energy equation of state, with the greatest sensitivity found at the high mass end \citep{mass_function1, mass_function2}. Since the baryon distribution in massive galaxy clusters is dominated by hot, X-ray emitting gas tracing the cluster dark matter potential, X-ray observations provide a natural tool to identify clusters. If the cluster is relaxed and the gas in hydrostatic equilibrium, these observational measures allow us to calculate the cluster mass. However, in the current hierarchical model for structure formation, massive clusters grow through mergers between galaxy groups and less massive clusters, that formed along overdense filaments in the cosmic web. Thus mergers are expected to be common. Kinetic energy from the merging partner is converted mainly to thermal energy in the form of shocks and turbulence in the cluster gas, profoundly affecting the evolution of the ICM \citep{Fujitab, Fujitaa, Fujitac, zuhone2010, zuhone2011}. Non-hydostatic gas motions induced by the merger (`sloshing') are long lasting on timescales of order gigayears. Their characteristic signatures in X-ray observations, i.e. multiple surface brightness discontinuities (cold fronts) and/or sweeping spiral features in temperature and surface brightness, have been noted in many galaxy groups and clusters \citep[e.g. see review by][]{markevitch2007, ascasibar2006, paterno-mahler, randall_galaxy, machacek_galaxy}. The study of these merger-induced gas motions and entropy exchanges in massive clusters are key to understanding (1) the properties and evolution of the ICM and (2) how the departures from hydrostatic equilibrium affect uncertainties in the cluster mass determinations used to constrain cosmology. \begin{figure*}[t!] \begin{center} \includegraphics[width=\textwidth]{f1.pdf} \caption{Chandra background subtracted, exposure corrected, combined 0.5-2.5 keV full resolution ($1\,\mbox{pixel}=0 \farcs 492\times 0 \farcs 492$) image of RXJ1347 with log scale. The primary cD galaxy is labeled with a black cross and the subcluster cD galaxy is labeled with a white x. Arrows point to edges forming a spiral pattern.} \label{fig:xray} \end{center} \end{figure*} Outstanding questions include: \begin{itemize} \item{What role do merger shocks play in heating the cluster gas \citep{zuhone2010}?} \item{ Does sloshing inhibit or promote cooling in the cluster core \citep{zuhone2010, zuhone2011}?} \item{ What can the morphology of the cold fronts and spirals tell us about the merger history of the cluster, cluster magnetic fields, or microphysics of the gas \citep[see, e.g. review by][]{markevitch2007, roedigerb, roedigera, roedigerc, roedigerd, roedigere}}? \item{ Does gas sloshing amplify magnetic fields and reaccelerate electrons, creating radio mini-halos found to be correlated with cold front edges in some clusters \citep{gitti2002, gitti2004, mazzotta2008, zuhone2013}?} \end{itemize} Progress on these questions will be found by a comparison of deep X-ray observations of relatively nearby ($z \lesssim 0.5$), massive, merging galaxy clusters, where gas temperatures and densities at cold fronts, shocks, and sloshing spirals can be well studied, with high resolution numerical simulations that connect the observed X-ray features with the orbital history of the merger and the microphysical properties of the surrounding gas. In this paper we use new and archival X-ray observations and archival Sunyaev-Zel'dovich effect (SZE) observations from the literature to characterize the cold fronts, shocks, and sloshing in RXJ$1347.5-1145$ (RXJ1347). With a $2 - 10$\,keV X-ray luminosity of $6 \times 10^{45}$\,{\rm erg\,s$^{-1}$}\,, RXJ1347 is the most X-ray luminous galaxy cluster found in the Rosat All Sky Survey \citep{schindler} and is still one of the more X-ray luminous galaxy clusters known. RXJ1347 is a massive, highly evolved cool core cluster, yet still actively merging with a less massive subcluster. Cool core clusters often show signatures of gas sloshing due to the presence of easily disturbed low entropy gas in their cores \citep{markevitch2007}. This makes RXJ1347 an ideal laboratory to study the on-going evolution of the bright end of the cluster luminosity function. It has been widely studied across many wavebands \citep[see, e.g.][for a review]{johnson2012}. Optical spectroscopic surveys of its galaxy population yield a velocity dispersion and dynamical mass within $r_{ 200}$ for RXJ1347 of $\sigma = 1163 \pm 97$\,\kms and $1.16^{+0.32}_{-0.27} \times 10^{15}\mbox{$M_\odot$}$, respectively \citep{lu2010, cohen}. \citet{lu2010} also found that RXJ1347 resides on a large scale filament, separated from a less massive galaxy cluster (RXJ1347-SW) by $\sim 7$\,Mpc in projection and $4000$\,\kms in radial velocity and with an excess of galaxies in between. Although the large velocity difference between the two clusters makes it unlikely that they are interacting, continued merging activity of RXJ1347 with other subclusters embedded within the same large scale filament would be expected in hierarchical cosmological models. The cluster K-band light is dominated by the two brightest cluster galaxies (BCGs). One, at (RA $13^h47^m30.7^s$, Dec $-11^{\circ}45'10 \farcs 1$ ) near the peak of the X-ray surface brightness, is assumed to be the central dominant galaxy of the primary cluster, and the second, at (RA $13^h47^m31.9^s$, Dec $-11^{\circ}45'10 \farcs 9$) lies $18''$ to the east. Given the second galaxy's relative radial velocity difference of only $\sim 100$\kms with respect to the primary cluster BCG, it has been interpreted as the central dominant galaxy of a merging subcluster. Surface mass density maps from weak and strong lensing show a general elongation in the mass concentration in the direction of the second BCG. However, they do not show a clear mass concentration peak there \citep{bradac2008, schmidt}. Early X-ray observations by {\it Chandra} \citep{allen} and {\it XMM-Newton} \citep{gitti2004} found excess X-ray emission~$\sim20''$~southeast of the primary BCG, which we now interpret as emission from hot gas in a merging subcluster. Deeper {\it Chandra} observations have revealed two sloshing cold fronts in the primary cluster as a result of a recent merger \citep{johnson2012}. Sunyaev-Zel'dovich observations have played a critical role in understanding the merging subcluster's gas dynamics. Observations by \citet{komatsu2001} with the Nobeyama 45 m telescope at~150~GHz first showed a significant SZ decrement~$\sim 20''$~southeast of the primary cluster, suggesting RXJ1347 is a disturbed system that has recently undergone a merger \citep{kitayama2004}. Higher resolution observations with the MUSTANG camera on the Green Bank Telescope at~90~GHz indicated the SZ decrement likely corresponds to shock heated gas from the merging subcluster \citep{mason2010, korngut2011}. A higher temperature and characteristic high pressure gradient support the interpretation that the region between the clusters is a shock front \citep{mason2010}. RXJ1347 also hosts a radio mini-halo centered on its primary cluster. While gas sloshing causes electron re-acceleration in the mini-halo, \citep{mazzotta2008}, \citet{ferrari2011} used observations by the Giant Metrewave Radio Telescope at 237 MHz and 614 MHz to find that excess radio emission in the southeast section of the mini-halo is likely caused by a propagating shock front corresponding to the shock-heated gas and SZ decrement previously found. At a redshift of $z=0.451$, RXJ1347 is close enough that the disturbed X-ray morphology resulting from recent mergers can be observed in detail with the high angular resolution of the {\it Chandra} X-ray Observatory. At this high angular resolution, our combined $186$\,ks {\it Chandra} exposure ( $> 2.5$ times deeper than previous {\it Chandra} observations) allows us to identify features not seen before, to better constrain the ICM gas hydrodynamics and, by comparison with simulations, RXJ1347's merger history. In \S\ref{sec:obs} we discuss our reduction of the observational data and general analysis techniques. In \S\ref{sec:qual} we analyze the mean and asymmetric gas properties in the primary cluster. We discuss the properties of the merging subcluster in \S\ref{sec:subcluster}, and we compare our results to simulations in \S\ref{sec:sim} to discern the cluster merging history. For the standard $\Lambda$ dominated cold dark matter cosmology and assuming $H_0 = 70$\,{\rm km\,s$^{-1}$\,Mpc$^{-1}$}\,, $\Omega_m = 0.3$, $\Omega_\Lambda = 0.7$, the redshift $z=0.451$ for RXJ1347.5-1145 corresponds to a luminosity distance of $2504$\,Mpc and angular scale $1'' = 5.77$\,kpc \citep{cosmocalc}. All WCS coordinates are J2000, and uncertainties are at $90 \%$ CL unless otherwise specified. \section{OBSERVATIONS AND DATA REDUCTION} \label{sec:obs} \begin{deluxetable}{ccccc} \tablewidth{0pc} \tablecaption{{\it Chandra} X-ray Observations \label{tab:obs}} \tablehead{ \colhead{ObsID} & \colhead{Exposure Time} & \colhead{Bkg Norm}& \colhead{Date} & \colhead{PI} \\ & (ks) } \startdata \dataset [ADS/Sa.CXO#obs/3592] {3592} & $49.643$ &$0.870$ & $2003-09-03$ & L. van Speybroeck \\ \dataset [ADS/Sa.CXO#obs/14407] {14407} & $56.076$ &$0.769$ &$2012-03-16$ & C. Jones \\ \dataset [ADS/Sa.CXO#obs/13999] {13999} & $47.817$ &$0.842$ &$2012-05-14$ & C. Jones \\ \dataset [ADS/Sa.CXO#obs/13516] {13516} & $32.629$ &$0.772$ &$2012-12-11$ & S. Murray \enddata \tablecomments{{\it Chandra} observations used in this analysis. All observations were taken in VFAINT mode with ACIS-I at aim point. Column $2$ lists the effective exposure after excluding periods of anomalously high and low count rates (see \S\protect\ref{sec:obs}). Column $3$ lists the additional normalization needed for blank sky backgrounds to match the observation count rate in the $10-12$\,keV energy band.} \end{deluxetable} In Table \ref{tab:obs} we list the data sets used in this analysis. All observations were taken by the Advanced CCD Imaging Spectrometer \citep[ACIS][]{acisa,acisb} on board the {\it Chandra} X-ray Observatory with the ACIS-I array at aim point. Using the {\it Chandra} Interactive Analysis of Observations suite of analysis tools \citep[CIAO 4.6][]{ciao} and CALDB 4.6.1.1, the data were reprocessed with the latest calibrations including updated gains and observation dependent bad pixel files, and corrections for the charge transfer inefficiency on the ACIS CCDs, the time-dependent build-up of contaminant on the optical filter, and the secular drift of the average pulse-height amplitude for photons of fixed energy (tgain). The data were then filtered to reject bad array patterns (grades $1$, $5$, and $7$), and to remove data flagged by the VFAINT mode as having excessive counts in border pixels surrounding event islands. Particle flares were cleaned from the data with the tool {\em lc\_clean}, using a $3\sigma$ clipping algorithm to calculate the mean count rate in the $0.3-12$\,keV energy band, and then rejecting time periods in which the count rate fell more than $20\%$ above or below the mean. These effective exposure times for each observation are listed in Table \ref{tab:obs}, resulting in a combined useful exposure for our analysis of $186,165$\,s. Source free background sets, provided by the Chandra X-ray Center, appropriate for the ACIS-I detector and date of observation, were reprojected onto each observation, and renormalized to match the observation count rate in the $10 - 12$\,keV energy band where particle backgrounds dominate. These additional normalization factors are also listed in Table \ref{tab:obs}. Background subtracted, exposure corrected mosaiced X-ray surface brightness images were then created in various energy bands using the CIAO script {\em flux\_image} and other CIAO tools. We used the broad band ($0.5- 7$\,keV) mosaiced X-ray surface brightness image to identify `by eye' $77$ point sources in the combined field of view. Since we are interested in the properties of the diffuse emission from the cluster gas, these sources were excised from the data. We compared our point source list to that given in the Chandra Source Catalog \citep{evans} for this region, and found that our removal of point sources was more conservative. We also excluded from our analysis a linear artifact (read-out streak), seen in the mosaiced image to extend from the bright X-ray peak at the cluster center to the northeast and southwest (at $110^\circ$ and $290^\circ$, respectively, measured counter-clockwise from west). Since we are interested in the properties of the diffuse emission in the cluster rather than the central bright source, excluding this artifact will not affect our results. \begin{figure}[tp!] \begin{center} \includegraphics[width=0.475\textwidth]{f2.pdf} \caption{We show the flux image from Figure \ref{fig:xray} with a mask covering the region excluded from the mean primary cluster gas distribution fit. The figure has a log scale with the primary (subcluster) cD galaxy labeled as a white cross (black X).} \label{fig:mask_resid} \end{center} \end{figure} \section{Primary Cluster} \label{sec:qual} The $186$~ks~{\it Chandra} background subtracted, exposure corrected, combined $0.5-2.5$~keV full resolution X-ray surface brightness image is displayed in Figure \ref{fig:xray}. We used the~$0.5-2.5$~keV energy band for imaging analysis, even though the cluster is hot, because of lower signal to noise in the higher ($2.5-7.0$~keV) energy bands. The asymmetries in the surface brightness suggest the presence of dynamic, complex gas motions in both the primary cluster and merging subcluster. The subcluster is visible southeast of the primary cD galaxy, with its X-ray peak $\sim 22''$ distant from the primary cluster BCG, as seen in previous X-ray and SZ observations \citep[see, e.g.][]{johnson2012, komatsu2001, kitayama2004, ota}. In this section we focus on the properties of the primary cluster gas, deferring the discussion of the subcluster to Section \ref{sec:subcluster}. Edges in the surface brightness trace a spiral feature in the primary cluster. We have noted these edges, as well as the centers of the primary and subcluster cD galaxies in Figure \ref{fig:xray}. To understand these gas asymmetries in the primary cluster, we begin with an analysis of its mean gas distribution. \subsection{Gas Distribution} \label{sec:mod} We use {\it Sherpa} to fit a $2$ dimensional elliptical beta model of surface brightness to the RXJ$1347$ $0.5-2.5$ keV full resolution ($1$ pixel$=0 \farcs 492$) counts image: \begin{eqnarray} \label{eq:ell_beta} S_x\left(r\right) = A\left[1+\left(\frac{r}{r_0}\right)^2\right]^{-\alpha} \end{eqnarray} with \begin{eqnarray} r\left(x,y\right) = \frac{\sqrt{x_{new}^2(1-\epsilon)^2+y_{new}^2}}{1-\epsilon} \end{eqnarray} \begin{eqnarray} x_{new} = \left(x-x_0\right)\cos\left(\theta \right)+\left(y-y_0\right)\sin\left(\theta \right) \end{eqnarray} \begin{eqnarray} y_{new} = \left(y-y_0\right)\cos\left(\theta \right)-\left(x-x_0\right)\sin\left(\theta \right) \end{eqnarray} and where $x_0$ and $y_0$ are the coordinates of the X-ray peak, A is the amplitude at $x_0$ and $y_0$, $\epsilon$ is ellipticity, $\theta$ is the angle of ellipticity, $r_0$ is the core radius, and $\alpha$ is the power law index. An exposure map for the full resolution $0.5-2.5$ keV image was included in the fit. To obtain parameters descriptive of only the primary cluster rather than the system with both clusters, a sector covering the subcluster centered on the second brightest cluster galaxy (RA $13^h47^m31.9^s$, Dec $-11^{\circ}45'10\farcs 9$) spanning an angle of $227^{\circ}$ to $324^{\circ}$ counterclockwise from west was excluded from the image using CIAO tools (see Figure \ref{fig:mask_resid}). The sector's size was chosen given the cone-like appearance of the subcluster in Figure \ref{fig:xray}. We choose to exclude the subcluster's gas contribution rather than jointly fit both clusters' distributions because there is little diffuse subcluster gas contribution to the primary cluster outside the cone surrounding the subcluster. Parameters for modeling the masked image were found by first using \textit{Sherpa}'s Monte Carlo optimization method with the Cash statistic to roughly locate the global minimum \citep{cash, montecarlo}. The Levenberg-Marquardt minimization technique was then used with the Monte Carlo best-fit as starting points to precisely locate the minimum \citep{levmar}. Core radius $r_0$, $x_0$ and $y_0$ positions, ellipticity $\epsilon$, angle of ellipticity $\theta$, amplitude $A$, and model power law index $\alpha$ were allowed to vary simultaneously during each minimization. Parameters for the best fit are listed in Table \ref{tab:param}. \begin{deluxetable}{cccc} \tablewidth{0pc} \tablecaption{2 Dimensional Elliptical Beta Model Parameters: Primary Cluster\label{tab:param}} \tablehead{\colhead{Parameter} & \colhead{Best-Fit} & \colhead{Lower Bound} & \colhead{Upper Bound} } \startdata $r_0$ & $5.1$ & $-0.1$ & $0.04$ \\ A & $1.6\times10^{-6}$ & $-1.6\times10^{-8}$ & $4.5\times10^{-8}$ \\ $\alpha$ & $1.12$ & $-0.005$ & $0.004$ \\ $x_0$ & $13:47:30.7$ & $-0.01$ & $0.1$ \\ $y_0$ & $-11:45:08.1$ & $-0.03$ & $0.1$ \\ $\epsilon$ & $0.2$ & $-0.005$ & $0.006$ \\ $\theta$ & $88$ & $-1$ & $1$ \enddata \tablecomments{Best-fit {\it Sherpa} parameters with uncertainties for the primary cluster fit. $x_0$ and $y_0$ are given in RA, Dec. Core radius and its uncertainties, as well as $x_0$ and $y_0$ uncertainties, are in arcsec, while amplitude (A) is in photons pix$^{-1}$~cm$^{-2}$~s$^{-1}$. $\alpha$ and $\epsilon$ are unitless, and $\theta$ has units of degrees. Uncertainties for the fit are $90 \%$ CL. } \end{deluxetable} We can find the value of the familiar $\beta$ parameter by equating $\alpha=3\beta-1/2$ \citep{beta}. Given the best-fit $\alpha$ value of $1.12$, $\beta=0.54$ for a core radius of $r_0=29.4^{+0.2}_{-0.6}$~kpc. The average gas density can be modeled using our $\beta$ value: \begin{eqnarray} n_{gas}\left(r\right) = n_{gas}\left(0\right)\left[1+\left(\frac{r}{r_0}\right)^2\right]^{-3\beta/2} \end{eqnarray} Our mean $\beta$-model fit agrees well with that found by \citet{allen} ($\beta = 0.535 \pm 0.003$, $r_0 = 29.2 \pm 0.7$\,kpc), from fitting the $0.3-7$\,keV image excluding the southeast subcluster quadrant. Previous shallower observations by \citet{allen} and \citet{gitti2007b} found that the cluster appears, on average, symmetric and relaxed. Our model shows a small elongation in the N-S direction, consistent with qualitative features seen in the deep X-ray image in Figure \ref{fig:xray}. \begin{figure*}[tp!] \begin{center} \includegraphics[width=0.500\textwidth]{f3a.pdf} \includegraphics[width=0.450\textwidth]{f3b.pdf} \caption{(Left) Temperature map of RXJ1347. The rainbow color scale increases in temperature from black (lowest) to white (highest) in keV. The black circle is a removed point source. Arrows point to the cold fronts labeled in Figure \ref{fig:xray}. The fan region is shown by a dashed arrow. $1\,\sigma$ uncertainties in the fits range from $8\%$ for temperatures $\sim6.5$ keV in the core to $25-30\%$ for temperatures $\gtrsim 20$ keV in the cluster outskirts. (Right) Entropy contours overlaid on the full resolution combined flux image from Figure \ref{fig:xray} matched in WCS coordinates to the temperature map. Low entropy contours trace the cold fronts and fan that form part of the spiral. The +(X) symbol denotes the primary cluster (subcluster) BCG, respectively, in both panels.} \label{fig:cfpos} \end{center} \end{figure*} \begin{deluxetable*}{cccccc} \tablewidth{0pc} \tablecaption{Primary Cluster Cold Front Profile Regions\label{tab:ellreg}} \tablehead{\colhead{Region} & \colhead{Semi-Minor Axis} & \colhead{Semi-Major Axis} & \colhead{Position Angle} & \colhead{Angle A} & \colhead{Angle B} \\ & $\left( \mbox{arcsec}\right)$ & $\left( \mbox{arcsec}\right)$ & $\left( \mbox{deg}\right)$ & $\left( \mbox{deg}\right)$ & $\left( \mbox{deg}\right)$ } \startdata West & $5.1$ & $8.5$ & $74$ & $315$ & $42$\\ Southeast & $6.2$ & $8.4$ & $137$ & $183$ & $275$ \\ East & $6.8$ & $16.3$ & $107$ & $127$ & $152$ \\ Northeast & $6.6$ & $24.0$ & $99$ & $109$ & $127$ \enddata \tablecomments{Bounding ellipses trace sector edges. All sectors were centered at primary cD RA $13:47:30.7$, Dec $-11:45:10.1$ and subtend the angle from A to B, moving counter-clockwise from west. Position angles are defined as the angle between the semi-major axis and west. } \end{deluxetable*} \begin{deluxetable*}{ccccccc} \tablewidth{0pc} \tablecaption{Cold Front Broken Power Model Parameters\label{tab:cfparam}} \tablehead{\colhead{Region} & \colhead{Inner Slope} & \colhead{Outer Slope} & \colhead{Location of Edge} & \colhead{$\frac{\rho_2}{\rho_1}$} & \colhead{$\frac{T_2}{T_1}$} & \colhead{$\frac{P_2}{P_1}$} \\ & & & $\left( \mbox{arcsec}\right)$ & } \startdata West & $-0.65^{+0.10}_{-0.08}$ & $-1.14^{+0.05}_{-0.04}$ & $5.85^{+0.04}_{-0.03}$ & $1.80^{+0.10}_{-0.08}$ & $0.56^{+0.16}_{-0.13}$ & $1.00^{+0.29}_{-0.24}$ \\ Southeast & $-0.78 \pm 0.09$ & $-0.72^{+0.06}_{-0.05}$ & $7.10^{+0.07}_{-0.03}$ & $1.62^{+0.15}_{-0.13}$ & $0.65^{+0.29}_{-0.22}$ & $1.06^{+0.48}_{-0.37}$ \\ East & $-0.53 \pm 0.08$ & $-1.36^{+0.06}_{-0.14}$ & $11.5^{+1.3}_{-1.2}$ & $1.51^{+0.11}_{-0.21}$ & $0.66^{+0.48}_{-0.24}$ & $0.99^{+0.73}_{-0.39}$ \\ Northeast & $-0.76 \pm 0.04$ & $-1.57 \pm 0.06$ & $16.7^{+0.3}_{-0.5}$ & $1.44 \pm 0.09$ & $0.58^{+0.27}_{-0.16}$ & $0.83^{+0.39}_{-0.24}$ \enddata \tablecomments{Parameters correspond to the best-fit density model for the elliptical profiles. $\rho_1$, $T_1$, and $P_1$ refer to the density, temperature, and pressure outside the cold front, and $\rho_2$, $T_2$, and $P_2$ represent the density,temperature, and pressure inside the cold front. Uncertainties are presented at $90 \%$ confidence. Temperatures were fit on each side of the edges (see Table \protect\ref{tab:cftemp}). } \end{deluxetable*} \begin{deluxetable}{ccc} \tablewidth{0pc} \tablecaption{Spectral Local Background Regions\label{tab:lbgreg}} \tablehead{\colhead{ObsID} & \colhead{$x_0$} & \colhead{$y_0$} \\ & RA & dec } \startdata $13516$ & $13:47:41.1$ & $-11:44:05.2$ \\ $3592$ & $13:47:41.1$ & $-11:44:05.2$ \\ $13999$ & $13:47:40.4$ & $-11:44:46.9$ \\ $14407$ & $13:47:39.1$ & $-11:43:59.3$ \enddata \tablecomments{Unrotated rectangles with sides of $39 \farcs 9$ and $38 \farcs 9$ were used as local backgrounds when extracting spectra from regions. } \end{deluxetable} \subsection{Characterization of Gas Asymmetries} \label{sec:asym} Displacement of the X-ray surface brightness peak to the west of the primary cluster cD galaxy indicates the core has most recently sloshed to the west (see Figure \ref{fig:xray}). We show a projected temperature map in the left panel of Figure \ref{fig:cfpos}. The temperature map was created by growing regions around each pixel in the $0.6-9$ keV mosaiced image until the region contained at least 3000 counts. The spectrum was then fit with an absorbed APEC model with Galactic absorption and \citet{anders} abundance fixed at $0.5\,\mbox{Z}_{\odot}$ \citep{ota}. $1\,\sigma$ uncertainties in the fits range from $8\%$ for temperatures $\sim6.5$ keV in the core to $25-30\%$ for temperatures $\gtrsim 20$ keV in the cluster outskirts. The temperature map in Figure \ref{fig:cfpos} shows complicated structures. A temperature dip west of the primary BCG indicates the sloshing core's presence. High temperatures in the southwest and southeast at large radii (the white regions in Figure \ref{fig:cfpos}) relative to temperatures at similar radii to the north and northeast could be due to shock heating from a previous encounter. Wrapping around the primary BCG in the X-ray flux image is a clockwise spiral that is elongated to the north. We denote these spiral edges in Figure \ref{fig:xray}. Both the west and east cold fronts identified by \citet{johnson2012} form part of the spiral. Edges in the projected temperature image align well with those in the X-ray flux image (see arrows in Figures \ref{fig:xray} and \ref{fig:cfpos}). In the right panel of Figure \ref{fig:cfpos}, we overlay pseudo-entropy ($s=TS_X^{-1/3}$) contours on the flux image. The contours show low entropy gas in the primary cluster core that traces the cold fronts. A diffuse, cool fan extends from the northeastern spiral elongation in both the temperature and flux images. The low entropy contours also elongate along this fan feature. However, the temperature map is coarse, giving only a qualitative picture of the complicated temperature structure in the cluster and serving as a guide for more careful spectral modeling of regions of interest. \subsection{Cold Fronts} \label{sec:coldfront} Elliptical surface brightness profiles of the combined, background subtracted, exposure corrected $0.5-2.5$ keV full resolution flux image were taken to characterize edge properties. The profiles were taken in four sectors, each centered on the primary cluster BCG, based on features seen in the X-ray and temperature images described in \S\ref{sec:asym}. The ellipses used to construct the profiles were chosen congruent to a bounding ellipse that traces the edge in each sector. See Table \ref{tab:ellreg} for parameters defining each sector and the bounding ellipses used to trace the edges in each sector. We integrate a broken power law model of density along the line of sight to fit the elliptical surface brightness profiles across each edge. The density model is given by: \begin{align} \label{eq:cf1} n_{gas}\left(r\right)&=n_{e,0}\frac{\rho_2}{\rho_1}\left(\frac{r}{r_s}\right)^{\nu_2} \qquad &\mbox{for}\; r<r_s \\ \label{eq:cf2} n_{gas}\left(r\right)&=n_{e,0}\left(\frac{r}{r_s}\right)^{\nu_1} \qquad &\mbox{for}\; r>r_s \end{align} where $n_{e,0}$ is the overall normalization, $\frac{\rho_2}{\rho_1}$ is the density ratio between the inner and outer power law slopes, $r_s$ is the location of the edge, and $\nu_2$ and $\nu_1$ are the inner and outer power law slopes, respectively. Results of our fits are summarized in Table \ref{tab:cfparam}. \begin{deluxetable*}{cccccc} \tablewidth{0pc} \tablecaption{Cold Front Spectral Regions\label{tab:specreg}} \tablehead{\colhead{Sector} & \colhead{Semi-Minor Axis Out} & \colhead{Semi-Major Axis Out} & \colhead{Semi-Minor Axis In} & \colhead{Semi-Major Axis In} \\ & $\left( \mbox{arcsec} \right)$ & $\left( \mbox{arcsec} \right)$ & $\left( \mbox{arcsec} \right)$ & $\left( \mbox{arcsec} \right)$ } \startdata W3 & $3.50$ & $5.81$ & $2.70$ & $4.49$ \\ W2 & $5.14$ & $8.53$ & $3.50$ & $5.81$ \\ W1 & $8.00$ & $13.38$ & $5.14$ & $8.53$ \\ W0 & $11.00$ & $18.25$ & $8.00$ & $13.38$ \\ SE3 & $3.68$ & $5.01$ & $1.88$ & $2.56$ \\ SE2 & $6.60$ & $8.99$ & $3.68$ & $5.01$ \\ SE1 & $9.50$ & $12.94$ & $6.60$ & $8.99$ \\ SE0 & $12.50$ & $17.03$ & $9.50$ & $12.94$ \\ E3 & $3.69$ & $8.82$ & $0$ & $0$ \\ E2 & $6.82$ & $16.32$ & $3.69$ & $8.82$ \\ E1 & $12.3$ & $29.41$ & $6.82$ & $16.32$ \\ E0 & $19.68$ & $47.06$ & $12.3$ & $29.41$ \\ NE3 & $3.48$ & $12.65$ & $1.72$ & $6.26$ \\ NE2 & $6.61$ & $24.04$ & $3.48$ & $12.65$ \\ NE1 & $14.76$ & $53.68$ & $6.61$ & $24.04$ \enddata \tablecomments{Elliptical segments are restricted to lie within each sector and are congruent to the bounding ellipse that traces the edge within the sector (see Table \protect\ref{tab:ellreg}). } \end{deluxetable*} \begin{figure}[t!] \begin{center} \includegraphics[width=0.5\textwidth]{f4.pdf} \caption{Sectors used to extract the elliptical surface brightness profiles plotted in the top panel of Figure \protect\ref{fig:eastcf} overlaid on the full resolution flux image of RXJ1347 from Figure \protect\ref{fig:xray}. Regions used for spectral analysis are shown as elliptical arcs concentric to the bounding ellipses (long dashed lines) listed in Table \protect\ref{tab:ellreg} that trace the observed surface brightness edges within each sector.} \label{fig:entropy_cont} \end{center} \end{figure} \begin{deluxetable*}{ccccccc} \tablewidth{0pc} \tablecaption{Cold Front Temperatures\label{tab:cftemp}} \tablehead{\colhead{Sector} & \colhead{Temperature} & \colhead{$\chi^2/\left(\mbox{d.o.f.}\right)$}\\ & $\left(\mbox{keV}\right)$ & } \startdata W3$^{\ast}$ & $5.9^{+1.0}_{-0.8}$ & $75.75/84$ \\ W2 & $7.4^{+1.0}_{-0.9}$ & $204.80/143$ \\ W1 & $13.3_{-2.7}^{+3.3}$ & $166.91/145$ \\ W0 & $13.6\pm2.7$ & $130.17/135$ \\ SE3$^{\ast}$ & $12.8_{-3.4}^{+4.9}$ & $81.48/84$ \\ SE2 & $13.8_{-3.0}^{+3.6}$ & $169.14/156$ \\ SE1 & $21^{+8}_{-5}$ & $110.78/130$ \\ SE0 & $20^{+7}_{-4}$ & $173.42/151$ \\ E3$^{\dagger}$ & $11.9^{+5.8}_{-3.2}$ & $62.31/69$ \\ E2$^{\ast}$ & $10.7^{+3.8}_{-2.2}$ & $82.01/83$ \\ E1$^{\ast}$ & $16.2^{+10.3}_{-4.7}$ & $82.22/77$ \\ E0$^{\dagger}$ & $9.7^{+4.4}_{-2.3}$ & $51.67/67$ \\ NE3$^{\dagger}$ & $9.0^{+3.5}_{-1.9}$ & $80.16/66$ \\ NE2$^{\star}$ & $8.9^{+2.4}_{-1.5}$ & $98.93/100$ \\ NE1$^{\star}$ & $15.4^{+5.8}_{-3.5}$ & $97.56/103$ \enddata \tablecomments{Temperatures are presented with $90 \%$~CL after being fit with fixed Galactic absorption and abundance at $0.5\,\mbox{Z}_{\odot}$. Labels of W, SE, E, and NE correspond to temperatures in the West, Southeast, East, and Northeast profiles. Numerical labels of 3, 2, 1, and 0 correspond to the most inward sector, the sector just inside of the cold front, the sector just outside the cold front, and the sector farthest from the primary BCG. We use the following symbols to denote the combined number of net counts: $\dagger$: >1000 counts, $\ast$: >1500 counts, and $\star$: >2000 counts. Unlabeled sectors have more than 2500 net counts. } \end{deluxetable*} We extracted spectra for regions of interest using CIAO script {\em specextract} and chose $39\farcs 9 \times 38\farcs 9 $ rectangular source free regions on the ACIS-I array as local backgrounds for each observation (see Table \ref{tab:lbgreg}). Spectra from each of the four observations were fit together, but extracted separately for each region. Except where noted, we required a minimum of $2000$ net counts total to constrain the expected high gas temperatures in this cluster. Spectra of diffuse gas were modeled in XSpec 12.8.0, using absorbed APEC models \citep{apec} with fixed Galactic absorption ${\rm n_H}= 4.75 \times 10^{20}$\,{\rm cm$^{-2}$}\, \citep[Leiden/Argentine/Bonn collaboration,][]{snowden} and abundance at $0.5\,\mbox{Z}_{\odot}$ \citep{ota}. See Tables \ref{tab:lbgreg} and \ref{tab:specreg} for cold front spectral region parameters and Table \ref{tab:cftemp} for temperature values. Spectral regions are also shown within sectors in Figure \ref{fig:entropy_cont}. \begin{figure*}[tp!] \begin{center} \includegraphics[width=0.497\textwidth]{f5a.pdf} \includegraphics[width=0.497\textwidth]{f5b.pdf} \caption{Surface brightness profiles, best-fit density profiles, temperature profiles, and pressure profiles are displayed for the 4 identified cold fronts. The left panel shows the west and southeast fronts, while the right panel displays the east and northeast fronts. Edge locations for the fronts are denoted by a vertical dashed line. Uncertainties in the edge fit parameters are colored to correspond to the relevant profiles. Best-fit parameters are in Table \protect\ref{tab:cfparam}.} \label{fig:eastcf} \end{center} \end{figure*} Surface brightness, density, temperature, and pressure profiles for the west, southeast, east, and northeast elongated edges are plotted in Figure \ref{fig:eastcf}. We find the west edge is well described by a density jump of $1.80^{+0.10}_{-0.08}$ and temperature jump of $0.56^{+0.16}_{-0.13}$ located $5\farcs 85^{+0.04}_{-0.03}$~west of the primary cluster cD galaxy. These measurements yield a pressure jump of $1.00^{+0.29}_{-0.24}$ across the edge. We find similar density, temperature, and pressure jumps across the southeast, east, and northeast edges located at $7\farcs 10^{+0.07}_{-0.03}$, $11\farcs 5^{+1.3}_{-1.2}$, and $16\farcs 7^{+0.3}_{-0.5}$~from the primary cluster cD galaxy, respectively (see Table \ref{tab:cfparam}). Our density, temperature, and pressure ratios show that all of these edges are consistent with sloshing cold fronts. Our east and west cold fronts agree well with those found by \citet{johnson2012}. Small differences in edge radii and jump values are due partly to our longer exposure, and also to differences in binning, density model, and the geometry and centering used to extract the profiles. The cold fronts we have measured agree with the edges identified visually in the images in \S\ref{sec:asym}. The radial distance between the cold fronts and the primary cluster cD galaxy gradually increases while rotating clockwise from west. This pattern is characteristic of a clockwise gas spiral formed from gas sloshing. Cold fronts radially closer to the primary cluster cD galaxy were formed more recently than those farther away. \begin{figure}[tp!] \begin{center} \includegraphics[width=0.495\textwidth]{f6a.pdf} \par \vspace{0.1cm} \includegraphics[width=0.495\textwidth]{f6b.pdf} \par \vspace{0.1cm} \includegraphics[width=0.495\textwidth]{f6c.pdf} \caption{GGM filtered images using standard deviations $\sigma$ of 1 (top), 2 (middle), and 3 (bottom). The primary and subcluster BCGs are denoted by a cross and X, respectively. Solid arrows in the top and middle panels point to cold fronts labeled in Figure \ref{fig:xray}. In the bottom panel, the black line corresponds to the ellipse edge that traced the eastern shock front. } \label{fig:gradient} \end{center} \end{figure} As an independent check on our edge identifications, we apply the Gaussian gradient magnitude (GGM) filter to the background subtracted, full resolution flux image (Figure \ref{fig:xray}), following the methodology in \citep{sanders2,sanders}. This filter highlights edges by showing high values for large gradients and low values for flat surface brightness profiles. In Figure \ref{fig:gradient} we show filtered images using standard deviations $\sigma$ of 1 pixel ($0 \farcs 492$), 2 pixels ($0 \farcs 984$), and 3 pixels ($1 \farcs 476$) in the top, middle, and bottom panels, respectively. A smaller $\sigma$ will highlight small scale gradients, while a larger $\sigma$ will highlight broad features. In the top and middle panels we have overlaid the 4 solid arrows denoting cold fronts seen in Figure \ref{fig:xray}. These arrows coincide with sharp gradients in the map, confirming our edge identifications and clockwise spiral feature. Filtering with $\sigma=1$, the cold fronts appear to be traced by filamentary structures, possibly similar to those in the Centaurus cluster discussed by \citet{sanders}. The bright central region likely corresponds to the bright, cool core in Figures \ref{fig:xray} and \ref{fig:cfpos}. In the middle panel of Figure \ref{fig:gradient} ($\sigma=2$), the cold front spiral is even more prominent. The elbow-like feature seen to the north in the top two panels is likely associated with the cooler, lower entropy northern fan extension noted in Figure \ref{fig:cfpos}. \section{Subcluster Properties} \label{sec:subcluster} \begin{figure*}[tp!] \begin{center} \includegraphics[width=0.495\textwidth]{f7a.pdf} \includegraphics[width=0.495\textwidth]{f7b.pdf} \caption{(Left) Difference map with log scale resulting from the residuals to the spherical beta model fit to the primary cluster surface brightness in \S\protect\ref{sec:mod}. Note the prominence of the subcluster. We point to the sharp leading edges with arrows. (Right) Difference map with SZ contours overlaid, indicating the maximum SZ decrement over the western subcluster edge seen in the left panel, suggesting the edges represent the Mach cone of a shock front. The primary BCG, subcluster BCG, and subcluster X-ray peak are denoted with a cross, X, and circle, respectively. } \label{fig:sub_resid} \end{center} \end{figure*} In Figure \ref{fig:xray} excess X-ray emission, interpreted as gas from a merging subcluster, is seen to the southeast of the primary cluster. To better isolate the subcluster and study the X-ray morphology of its excess emission, we subtract the mean cluster model obtained in Section \ref{sec:mod} for the primary cluster alone from the merged 0.5-2.5 keV X-ray image of the merging pair. This X-ray surface brightness difference map is given in Figure \ref{fig:sub_resid}. \subsection{Qualitative Features: the Case for Gas Stripping and Core Shredding} \label{sec:sub_qual} Several qualitative features in Figure \ref{fig:sub_resid} are striking. The northern edges of the subcluster appear cone-like with the vertex displaced $\sim 2''$ south of the second BCG (taken to be the subcluster's central dominant galaxy), suggestive of gas displaced from the subcluster gravitational potential by ram pressure as it infalls through the primary cluster ICM from the southwest to the northeast. The western edge of this cone is in close correspondence to the maximum SZ decrement, as shown in the right panel of Figure \ref{fig:sub_resid} \citep{mason2010, korngut2011}, characteristic of a shock, suggesting that the cone-like feature may be the Mach cone of shocks formed by the subcluster's supersonic motion through the primary cluster ICM. Within the Mach cone is a bright region suggestive of the subcluster's gas core. The peak of the X-ray emission, defined as the brightest subcluster pixel in the $0.5-2.5$\,keV X-ray image shown in Figure \ref{fig:xray}, is located at RA~$13^h 47^m 31.5^s$, Dec $-11^{\circ} 45' 23\farcs 7$. The morphology of the excess X-ray emission is elongated towards the southwest with the hint of an extended tail, also qualitatively consistent with that expected for subcluster gas being stripped. \begin{figure*}[tp!] \begin{center} \includegraphics[width=0.475\textwidth]{f8a.pdf} \includegraphics[width=0.475\textwidth]{f8b.pdf} \caption{The left panel shows our difference map from Figure \protect\ref{fig:sub_resid} with entropy contours overlaid while the right panel shows the same contours overlaid on the temperature map for the same region matched in WCS coordinates. The primary BCG, subcluster BCG, and subcluster X-ray peak are denoted with a red cross, blue X, and blue circle, respectively. The black circle in the temperature map is a removed point source. } \label{fig:entropy} \end{center} \end{figure*} In Figure \ref{fig:entropy} we show pseudo-entropy contours overlaid on the surface brightness difference map (left panel) and temperature map (right panel) for the subcluster gas. Low entropy contours coincide with luminous, lower temperature regions in the subcluster and tail. We locate a `ledge' in the southern contours, indicating an irregular distribution of low entropy gas in the elongated region. In a cluster whose core is still intact, we would expect a spherical entropy distribution with regular, monotonically increasing contours of entropy as a function of distance from the subcluster gas peak. While there is a concentration of low entropy gas near the subcluster X-ray surface brightness peak, the elongated, asymmetric entropy contours and lower temperatures to the southwest of the subcluster peak suggest ongoing core shredding as cooler core gas gas is pushed back towards the tail-like structure. These qualitative features are confirmed by the GGM $\sigma=3$ map shown in the bottom panel of Figure \ref{fig:gradient}. The eastern shock edge in the GGM image corresponds to the edge ellipse used for our shock density ratio analysis in \S \ref{sec:shocks} (see Figure \ref{fig:sfell}). The western gradient edge corresponds to the SZ decrement. Together they trace the shock Mach cone. The bright extension to the southwest likely corresponds to the subcluster's stripped gas tail. \begin{deluxetable}{ccc} \tablewidth{0pc} \tablecaption{Circular Subcluster and Background Regions\label{tab:subreg}} \tablehead{\colhead{Region} & \colhead{Center} & \colhead{Flux ($0.5-2.5$\,keV) } \\ & $\left( \mbox{RA}\right)$, $\left( \mbox{Dec}\right)$ & $10^{-4}$\,\,{\rm photons\,s$^{-1}$\,cm$^{-2}$}\,} \startdata SC & $13:47:31.6$, $-11:45:27.4$ & $1.716\pm 0.016$ \\ BE & $13:47:32.1$, $-11:45:02.2$ & $0.775 \pm 0.011$ \\ BN & $13:47:30.4$, $-11:44:48.7$ & $1.491 \pm 0.015$ \\ BW & $13:47:29.2$, $-11:45:06.9$ & $0.634 \pm 0.010$ \enddata \tablecomments{SC includes emission from both the primary cluster and subcluster at $22''$ from the primary cluster central galaxy. BE, BN, and BW denote circular regions at the same radial distance containing primary cluster emission only. All circular regions have radius $11\farcs 7$. Flux uncertainties are $1\sigma$. } \end{deluxetable} \subsection{Subcluster Temperature, Luminosity, and Mass} \label{sec:mass} To measure the temperature and luminosity of the subcluster, we extract the subcluster spectrum from a circular region of radius $11 \farcs 7$ located $22''$ from the primary cluster central galaxy, that covers the majority of the subcluster emission. To minimize the contribution to the spectrum from the primary cluster gas, we choose local background circles of the same size and distance from the primary cluster central galaxy as the subcluster, as in \citet{johnson2012}. However, the sloshing features (spiral and fan) cause the flux to vary with angle by as much as a factor of $2$ at this distance (see Table \ref{tab:subreg} and Figure \ref{fig:sub_spec_reg}). We argue that the X-ray flux in the northern local background circle (BN in Figure \ref{fig:sub_spec_reg}) is anomalously high compared to that expected at the location of the subcluster due to this excess gas displaced to the north by sloshing. These sloshing features found to the north would not be expected at that radius in the southeast at the location of the subcluster. We instead fit the subcluster spectrum using the combined eastern (BE) and western (BW) cluster background regions as more representative of cluster emission at the location of the subcluster. \begin{deluxetable}{ccccccc} \tablewidth{0pc} \tablecaption{Subcluster Temperatures\label{tab:subtemp}} \tablehead{\colhead{Region} & \colhead{Temperature} & \colhead{$\chi^2$/(d.o.f.)}\\ & $\left( \mbox{keV}\right)$ & } \startdata Total Subcluster Gas & $16^{+4}_{-3}$ & $621.66/588$ \\ SZ Decrement & $19^{+6}_{-4}$ & $164.74/153$ \enddata \tablecomments{Temperatures are presented with $90 \%$~CL after being fit with an absorbed APEC model with fixed Galactic absorption and abundance at~$0.3\,Z_\odot$. } \end{deluxetable} Using an absorbed APEC thermal plasma model and fixing the absorbing column at the Galactic value ($4.75\times 10^{-20}\,{\rm cm}^{-2}$), we find a subcluster temperature of $kT=16^{+4}_{-3}$\,keV (see Table \ref{tab:subtemp}) and abundance $A=0.5^{+0.3}_{-0.2}\,Z_\odot$ ($\chi^2/{\rm dof}=629/587$). From directly summing the $0.5-2.5$\,keV flux in the source and subtracting the expected primary cluster contribution using the BE+BW background regions, we measure an absorbed $0.5-2.5$\,keV flux for the subcluster alone of $1.01 \pm 0.02 \times 10^{-4}$\,{\rm photons\,s$^{-1}$\,cm$^{-2}$}\, ($1.96 \pm 0.03 \times 10^{-13}$\,{\rm erg\,s$^{-1}$\,cm$^{-2}$}\, ). While these $1\sigma$ statistical uncertainties are small given the long exposure, we caution the reader that the uncertainties in measuring the subcluster flux are dominated by the choice of model for the primary cluster emission at that location, and may be as high as $\sim 25\%$. Our mean value for the net absorbed $0.5-2.5$~keV flux in the subcluster is higher (but still within their $1\sigma$ uncertainties) of that previously obtained by \citet{johnson2012} ($1.5 \pm 0.9 \times 10^{-13}$\,{\rm erg\,s$^{-1}$\,cm$^{-2}$}\,), because we exclude the northern region (BN) from our estimate of the average primary cluster emission and they did not. Using the spectral model to account for Galactic absorption and modest change in energy band to be consistent with \citet{zhang2011}, we find the unabsorbed $0.5-2$\,keV X-ray luminosity of the subcluster to be $L_{\rm X}= 1.3 \times 10^{44}$\,{\rm erg\,s$^{-1}$}\,. Applying the $M_{gas,500}-L_{0.5-2keV,500}$ relationship from \citet{zhang2011} for disturbed clusters including the core correction, \begin{eqnarray} \log_{10}(Y)=A+B\log_{10}(X) \end{eqnarray} where \begin{eqnarray} Y=\frac{L_{0.5-2keV,500}}{E(z)~\mbox{erg~s$^{-1}$}} \end{eqnarray} \begin{eqnarray} X=\frac{M_{gas,500}E(z)}{10^{14}~\mbox{M$_{\odot}$}} \end{eqnarray} with $A=44.44 \pm 0.61$, $B=1.16\pm 0.04 $~at~$1\sigma$, and \begin{eqnarray} E^2(z)=\Omega_m(1+z)^3+\Omega_{\Lambda}+(1-\Omega_m-\Omega_{\Lambda})(1+z)^2 \end{eqnarray} we calculate the enclosed gas mass to be~$3.3 \times10^{13}\,M_\odot$. Assuming the gas fraction is $10 \%$ of the total mass \citep{vikhlinin2006}, this yields the subcluster's total mass to be~$3.3 \times10^{14}$~M$_\odot$, consistent with the mass estimate found by \citet{johnson2012}. Since we expect that some subcluster gas may already have been stripped due to the encounter, this should be interpreted as a lower bound on the subcluster total mass. \begin{figure*}[bp!] \begin{center} \includegraphics[width=0.475\textwidth]{f9a.pdf} \includegraphics[width=0.475\textwidth]{f9b.pdf} \caption{(Upper Left) The surface brightness profile across the eastern shock front. The black line shows the best fit assuming a broken power law (Equations \ref{eq:cf1} and \ref{eq:cf2}) for the gas density. The potential merger cold front surface brightness jump is seen at~$\sim7-8''$. (Lower Left) Best fit gas density from the upper panel with power law indices~$-1.28^{+0.04}_{-0.03}$~($0.82^{+0.70}_{-0.59}$) upstream (downstream) of the shock. The shock edge location is denoted with a dashed line and $90\%$~CL uncertainties shaded in green. (Right) The sector used to construct the surface brightness profile across the shock is overlaid on the residuals image with log scale. The black line shows the segment of the bounding ellipse that traces the eastern shock edge within the sector. The potential merger cold front surface brightness discontinuity is denoted with a blue arrow. See also the bottom panel of Figure \ref{fig:gradient}.} \label{fig:sfell} \end{center} \end{figure*} \begin{figure}[tp!] \begin{center} \includegraphics[width=0.5\textwidth]{f10.pdf} \caption{Source (subcluster + cluster, solid circle) and local background (cluster only, dashed circles) regions overlaid on the $0.5-2.5$\,keV flux image of RXJ1347. Note the excess emission in background region BN due to the northern extension of the spiral and fan. } \label{fig:sub_spec_reg} \end{center} \end{figure} \begin{deluxetable*}{ccccccccc} \tablewidth{0pc} \tablecaption{Eastern Shock Properties \label{tab:sfparam}} \tablehead{\colhead{Location of Edge} & \colhead{$\frac{\rho_{2,m}}{\rho_{1,m}}$} & \colhead{$\frac{T_2}{T_1}$} & \colhead{$\frac{P_2}{P_1}$} & \colhead{Mach number} & \colhead{Velocity} \\ $\left( \mbox{arcsec}\right)$ & & & & & $\left( \mbox{km\,s}^{-1}\right)$ } \startdata $13.0^{+0.5}_{-0.3}$ & $1.38^{+0.16}_{-0.15}$ & $1.25$ & $1.72$ & $1.25 \pm 0.08$ & $2810^{+210}_{-240}$ \enddata \tablecomments{Shock properties were obtained by fitting a radial profile centered on the subcluster X-ray peak. Subscripts of $2$ and $1$ denote quantities downstream and upstream of the shock, respectively, whereas the subscript~$m$ denotes a measured, rather than derived, quantity. $T_2/T_1$, $P_2/P_1$, and the Mach number were derived from the Rankine-Hugoniot jump conditions using the measured density ratio. The shock velocity assumes an unshocked gas temperature of $19\pm3$\,keV.} \end{deluxetable*} \subsection{Eastern Shock Front and Shock Velocity} \label{sec:shocks} Following the same procedure as in \S\ref{sec:coldfront} with the~$0.5-2.5$~keV full resolution exposure-corrected X-ray flux image, we construct the surface brightness profile across the eastern shock edge using elliptical wedge regions concentric to a bounding ellipse tracing the eastern shock edge and constrained to lie within the sector subtending the angles from~$126^{\circ}$~to~$181^{\circ}$. The bounding ellipse is centered on the subcluster X-ray peak with semi-major (-minor) axes and position angle of~$44''$ ($12''$) and~$51^{\circ}$, respectively (see Figure \ref{fig:sfell}). We model the gas density using a broken power law density model across the subcluster eastern edge (see Equations \ref{eq:cf1} and \ref{eq:cf2}, \S\ref{sec:coldfront}). To fit the surface brightness profile shown in the upper left panel of Figure \ref{fig:sfell}, we integrate the square of the density model along the line of sight by using a multivariate~$\chi^2$~minimization scheme. Our results for the best fit subcluster gas density ratio across the shock are shown in the lower left panel of Figure \ref{fig:sfell}. We find the shock edge is located~$13\farcs 0^{+0.5}_{-0.3}$~northeast of the subcluster X-ray peak with a density ratio of~$1.38^{+0.16}_{-0.15}$. The shock Mach number can be determined from either the density ratio $\rho_2/\rho_1$ or temperature ratio $T_2/T_1$ across the shock using the Rankine-Hugoniot jump conditions \citep{RK,korngut2011} \begin{eqnarray} \label{eq:rk1} \mathcal M_{\rho} = \left[\frac{2\frac{\rho_2}{\rho_1}}{\gamma+1-\left(\gamma-1\right)\frac{\rho_2}{\rho_1}}\right]^{1/2} \end{eqnarray} and \begin{eqnarray} \label{eq:rk2} \mathcal M_T = \left\{\frac{8\frac{T_2}{T_1}-7+\left[\left(8\frac{T_2}{T_1}-7\right)^2+15\right]^{1/2}}{5}\right\}^{1/2} \end{eqnarray} where $\gamma=\frac{5}{3}$, and $\rho_2$ ($T_2$) and $\rho_1$ ($T_1$) are the gas densities (temperatures) inside and outside the shock, respectively. Using our measured density ratio in Equation \ref{eq:rk1}, we find a shock Mach number $M_{\rho} =1.25 \pm 0.08$. We calculate the velocity of the eastern shock by multiplying the Mach number by $c_s$, the speed of sound in the unshocked primary cluster plasma, using \begin{eqnarray} c_s^2=\frac{\gamma P_g}{\rho_g} \end{eqnarray} where \begin{eqnarray} P_g=n_gk_BT_K \end{eqnarray} \begin{eqnarray} \rho_g=n_g\mu m_p \end{eqnarray} and $\gamma=\frac{5}{3}$ and $\mu=0.6$. Thus, \begin{eqnarray} c_s^2=\frac{\gamma k_BT_K}{\mu m_p} & = & \frac{\gamma T_{eV}}{\mu m_p} \end{eqnarray} and \begin{eqnarray} \label{eq:soundspeed} c_s=516~\mbox{km \, s$^{-1}$} T_{keV}^{1/2}. \end{eqnarray} Using the local background regions from Table \ref{tab:lbgreg}, we extract a spectrum for region BE from Figure \ref{fig:sub_spec_reg} and Table \ref{tab:subreg} to the north of the tip of the Mach cone (see Figure \ref{fig:sub_resid}) to measure the temperature and determine the local sound speed in the cluster gas outside the shock. Modeling the spectrum with an absorbed APEC model with fixed Galactic absorption and abundance at $0.3\,Z_\odot$, we find a cluster gas temperature $kT=19\pm3$\,keV ($\chi^2/{\rm dof}=297/305$ for $7386$ net source counts, see Table \ref{tab:subtemp}). From Equation \ref{eq:soundspeed}, we estimate the speed of sound in the unshocked cluster gas as $c_s=2250^{+170}_{-190}$\kms and eastern shock velocity of $2810^{+210}_{-240}$\kms. From Equation \ref{eq:rk2}, we predict a (deprojected) temperature ratio $T_2/T_1 = 1.25$. To test this prediction, one would ideally extract spectra in narrow regions on either side of the shock edge. However, we cannot measure the temperature jump in such regions to within $25 \%$ uncertainty with these data ($T_2=16^{+23}_{-7}$ keV and $T_1=14^{+26}_{-6}$ keV, yielding $T_2/T_1 = 1.2^{+2.7}_{-0.7}$). Although we find no significant difference between the temperatures of the subcluster and the cluster gas just outside the shock, the measured temperature ratio is consistent with the Rankine Hugoniot prediction within $90\%$ CL uncertainties. \subsection{Subcluster Cold Front?} \label{sec:subcoldfront} In the elliptical surface brightness profile across the eastern shock (Figure \ref{fig:sfell}), we find a second jump in the surface brightness $\sim 5''$ closer to the subcluster X-ray peak than the shock front. The surface brightness discontinuity is small and gas temperatures are high and have a complex distribution, so measurements are highly uncertain, causing us to not complete a formal edge analysis. However, qualitatively the double-jump nature of this profile is reminiscent of the Bullet subcluster \citep[see][]{bullet}. For the Bullet, the inner jump was a cold front between the subcluster core and outer gas. The concentration of low entropy gas near the X-ray peak for our subcluster suggests that there is still a remnant core, and that, as in the Bullet cluster, the inner surface brightness edge may correspond to a merger cold front. \begin{figure*}[tp!] \begin{center} \includegraphics[width=0.475\textwidth]{f11a.pdf} \includegraphics[width=0.475\textwidth]{f11b.pdf} \caption{(Left) Residuals in log scale with SZ contours overlaid, indicating the maximum SZ decrement over the western shock. (Right) Projected temperature map in linear scale with the elliptical profile covering the western shock front. High gas temperatures coincide with the western and eastern shocks, while lower gas temperature are found in the subcluster core and tail. The region enclosed in black corresponds to the gas build up at the shock front seen in the surface brightness profile, shown in Figure \protect\ref{fig:west_sfell}.} \label{fig:resid} \end{center} \end{figure*} \begin{figure}[bp!] \begin{center} \includegraphics[width=0.5\textwidth]{f12.pdf} \caption{Elliptical surface brightness profile centered on the primary cD and aligned with the western shock front. The pile up of gas over the front is highlighted in red. The red points correspond to the region enclosed in black in Figure \protect\ref{fig:resid}. } \label{fig:west_sfell} \end{center} \end{figure} \subsection{Western Shock Front} As the subcluster merges with the primary cluster, gas is shock heated. High temperatures southwest of the subcluster cD galaxy suggest the presence of a western shock front (see right panel of Figure \ref{fig:resid}). SZ contours from a MUSTANG SZ map, provided by \citet{mason2010} and smoothed by a $10''$ FWHM Gaussian, were overlaid on the difference map (left panel of Figure \ref{fig:resid}). The most significant SZ decrement over the subcluster occurs from $\sim13''$ to $\sim17''$ southeast of the primary cluster cD galaxy. We choose a rectangular region centered at RA, Dec\,$=13^h47^m31.4^s, -11^\circ45'19\farcs5$ with long (short) sides of $12\farcs 3$ ($4\farcs 5)$, respectively, oriented at position angle $130^\circ$ along the most significant SZ decrement to study this region. Using an absorbed APEC model with fixed Galactic absorption, abundance fixed at $0.3Z_\odot$ and local backgrounds from Table \ref{tab:lbgreg}, we find the temperature of the gas in this region to be $19_{-4}^{+6}$~keV ($\chi^2/{\rm dof} = 165/174$). This region corresponds to the high temperature ridge between the subcluster and primary cluster in the right panel of Figure \ref{fig:resid}. A surface brightness profile centered on the primary cluster cD galaxy taken in the sector subtending the angles from $198^{\circ}$ to $258^{\circ}$ with semi-major (-minor) axes and position angle of~$56''$ ($15''$) and $138^{\circ}$, respectively, traces the shape of the decrement as well as the proposed western shock front. The profile indicates increased surface brightness over the edge. See Figure \ref{fig:west_sfell} for the profile with shock region in red, and right panel of Figure \ref{fig:resid} for the profile's location and shock region enclosed in black. High temperatures, high surface brightness, and a significant SZ decrement in this region suggest that the edge is the western shock front of the subcluster's Mach cone. The opening angle $\mu$ of the Mach cone can be calculated by \citep{markevitch2007} \begin{eqnarray} \label{eq:mach} sin\left(\mu\right)=\mbox{$\mathcal M^{-1}$}. \end{eqnarray} For our calculated eastern Mach number of $1.25 \pm 0.08$, we predict an opening angle of $\mu=53.1 \pm 0.1$\,degrees. Assuming the Mach cone is defined as the angle between the eastern shock edge (see Table \ref{tab:sfparam}) and the SZ decrement, we measure an opening angle $\mu\approx50$ degrees, in good agreement with the predicted value. \begin{figure}[b!] \begin{center} \includegraphics[width=0.475\textwidth]{f13.pdf} \caption{A simulated spectroscopic-like projected temperature map from \citet{zuhonesim} and \citet{johnson2012}. We label the same edges as \citet{johnson2012}. } \label{fig:johnson_orig} \end{center} \end{figure} \begin{figure*}[t!] \begin{center} \includegraphics[width=0.475\textwidth]{f14a.pdf} \includegraphics[width=0.475\textwidth]{f14b.pdf} \caption{In the left panel we show the projected temperature map with cold fronts highlighted. To the right we display the same simulated projected temperature map as in Figure \protect\ref{fig:johnson_orig} flipped about its N-S axis. We label edges that appear to match those found in the data. } \label{fig:temp_sim} \end{center} \end{figure*} \section{Cluster Merger History: A Comparison to Simulations} \label{sec:sim} To constrain the cluster's merger history, it is important to compare our observations with simulations. In Figure \ref{fig:johnson_orig} we present the same simulation and labels shown in \citet{johnson2012}, who argue that the subcluster is on its second pass around the primary cluster. The simulation initial conditions are described in \citet{zuhonesim} and employ a mass ratio of the primary cluster to the subcluster of $10$ and an impact parameter of~$\sim 1$~Mpc. The subcluster is on a clockwise orbit about the primary cluster, such that it first passes the primary cluster in the west. The simulation captures the subcluster's position and the shock fronts well, in comparison with the X-ray data. However, the cold front locations are not consistent with our observations. In the simulations, the eastern edge is closer to the primary BCG, suggesting that it formed later than the western cold front, but our observations definitively show that the western cold front is closer to the primary BCG. Thus our data suggest that the western cold front formed more recently than the eastern cold front. This leads us to consider the possibility that more than 1 subcluster has perturbed RXJ1347. In the right panel of Figure \ref{fig:temp_sim}, we display the same simulation but flipped about the N-S axis such that the subcluster perturber in the simulation is now in a counterclockwise orbit, first passing to the east of the primary cluster. The merger is in the plane of the sky, and we compare our temperature map to this simulation in the left panel. If we ignore the location of the subcluster in the simulation (now in the southwest), we find many similarities between our deep {\it Chandra} observations and the simulation. The four cold fronts forming a clockwise spiral in the observations can be located in the simulation. We denote these fronts with arrows in both the simulation and the temperature map. The fronts also form a clockwise spiral in the simulation. The fan found in the X-ray observations, and most prominently seen in our residuals image, is similar to the cool gas forming part of the spiral that extends to the north and northwest in the simulation. As the subcluster passes the primary cluster, it transfers angular momentum \citep{ascasibar2006}, so we expect the observed clockwise spiral, seen in Figure \ref{fig:xray} and confirmed by our cold front analysis, to have been caused by interaction with a subcluster on a counterclockwise orbit around the primary cluster. The clockwise spiral in the flipped simulation supports this scenario. This discrepancy between the observed subcluster to the southeast moving in a clockwise orbit, and the clockwise sloshing spiral in the primary cluster produced in simulations by multiple passes of a perturber moving in a counterclockwise orbit, suggests that the observed subcluster may be on its first pass through the primary cluster, while the clockwise sloshing spiral was caused by earlier encounters with a different subcluster moving in a counterclockwise orbit. In this scenario, the observed subcluster is being ram-pressure stripped while shock heating the cluster gas. The earlier perturber, whose interaction caused the sloshing spiral, may no longer be visible in X-rays because nearly all of its gas has already been stripped. RXJ1347 is in a $\sim20$~Mpc filament, so it would not be surprising if it had previously interacted with another object in this structure \citep{filament}. We searched the literature for evidence in other wavelengths of a second merging subcluster. \citet{lu2010} find a massive cluster located $\sim 7$ Mpc southwest of RXJ1347 (RXJ1347-SW) with $\sigma=780 \pm 100$ km s$^{-1}$, $M_{200}=3.4^{+1.4}_{-1.1} \times 10^{14}\,M_{\odot}$, and mean redshift $0.4708 \pm 0.0006$. RXJ1347-SW has a large relative radial velocity of $4000$ km s$^{-1}$. \citet{lu2010} argue the likelihood RXJ1347-SW is falling into RXJ1347 is low. Nonetheless, the concentration of galaxies between RXJ1347 and RXJ1347-SW in velocity space implies the two clusters are physically linked by the same large scale filament. The recent strong lensing analysis of \citet{schmidt}, using redshift measurements from the Cluster Lensing and Supernova survey \citep[CLASH,][]{postman}, requires two additional perturbers compared to previous lensing analyses \citep{halkola,bradac2008} to model the observed arc geometry. One is a dark halo aligned with a faint $z_{phot}>0.7$ object and not associated with the primary cluster, and the second is a massive ($\sigma_v\sim763$ km s$^{-1}$) concentration of cluster galaxies to the NW. While the modeled NW mass component is intriguing as a possible subcluster remnant, it does not appear coincident with any bright cluster galaxy, and no redshift is given. Thus, more work is needed to verify the NW component's existence and association with the primary cluster, as well as to resolve discrepancies between the CLASH redshift measurements of the lensed systems and those of previous work \citep{halkola,bradac2008}. Another possible scenario is that the orbit of the observed subcluster is not in the plane of the sky and that the observed clockwise orientation of the edges may be due to projection effects. While we consider this latter scenario unlikely, it cannot be ruled out without better simulation studies. \section{CONCLUSIONS} \label{sec:conclude} With~$2.5$~times the exposure of previous analyses, we are able to study the gas hydrodynamics of RXJ1347 and its merging subcluster in unprecedented detail. We briefly summarize our key results. \begin{itemize} \item{A series of four cold fronts west, southeast, east, and northeast located at~$5\farcs 85^{+0.04}_{-0.03}$, $7\farcs 10^{+0.07}_{-0.03}$, $11\farcs 5^{+1.3}_{-1.2}$, and $16\farcs 7^{+0.3}_{-0.5}$~from the primary cluster's cD galaxy, respectively, forms a clockwise spiral, suggesting the merger is in the plane of the sky. The west and east cold fronts correspond to those found by \citet{johnson2012}.} \item{We measure a subcluster $0.5-2$\,keV luminosity of $L_x= 1.3 \times 10^{44}$\,{\rm erg\,s$^{-1}$}\, and infer a lower bound on the total subcluster mass of $3.3 \times10^{14}$~M$_\odot$.} \item{We identify the shock from the supersonic infall of the subcluster and measure a density ratio across the shock of $1.38^{+0.16}_{-0.15}$~, corresponding to a Mach number of $1.25\pm0.08$ and shock velocity of~$2810^{+210}_{-240}$~km~s$^{-1}$. The measured opening angle of the Mach cone of~$\sim50$~deg is consistent with this shock interpretation.} \item{The subcluster's baryonic gas has been stripped from the dark matter peak and leaves a tail behind the subcluster. Asymmetric entropy contours and cooler temperatures extending through the subcluster and tail indicate the subcluster core is being shredded.} \item{We suggest that the most likely explanation for the observed X-ray features is that the southeast subcluster is on its first passage on a clockwise orbit through the cluster, shock heating the cluster gas as it infalls, while the clockwise sloshing spiral observed in the primary cluster was caused by earlier encounters with a second subcluster moving in a counterclockwise orbit. Better simulations and further observations in other wavelengths are needed to search for remnants of this second perturber and test this scenario.} \end{itemize} \acknowledgements This work was supported in part by NASA CXC grant GO2-13148X and the Smithsonian Institution. CK gratefully acknowledges support from the Smithsonian Institution Minority Awards Program. This research has made use of data obtained from the Chandra Data Archive and the Chandra Source Catalog, and software provided by the Chandra X-ray Center (CXC) in the application packages CIAO and Sherpa. This work has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. We would like to thank Maxim Markevitch for helpful discussions and use of his multivariate~$\chi^2$~minimization code for fitting the surface brightness profiles, Brian Mason for use of his SZ contours, Jeremy Sanders for help with his GGM filtering technique, Sherry Suyu for useful strong lensing discussions, and the anonymous referee for exceptionally helpful comments. \noindent{\it Facilities:} CXO (ACIS-I) \vfill \eject
73fd6623ea7b118a9d7258a7e38a169667b86a82
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }
\section{Introduction} In information theory, the \emph{index coding} problem~\cite{BK98a,BBJK06} is the following: A sender wishes to \emph{broadcast} over a noiseless channel an $n$-symbol string $x\in \mathbb{F}^n$ to a group of $n$ receivers $R_1,\ldots,R_n$, each equipped with some \emph{side information}, namely, a subvector $x_{K_i}$ of $x$ indexed by a subset ${K_i} \subseteq \{x_1,\ldots ,x_n\}$. The index coding problem asks what is the minimum length $m$ of a broadcast message that allows each receiver $R_i$ to retrieve the $i$th symbol $x_i$, given his side-information $x_{K_i}$ and the broadcasted message. The side information of the receivers can be modeled by a directed graph $\mathcal{K}_n$, in which $R_i$ observes % the symbols $K_i := \{ x_j \; : \; (i,j) \in E(\mathcal{K}_n) \}$. $\mathcal{K}_n$ is sometimes called the \emph{knowledge graph}. A canonical example is where $\mathcal{K}_n$ is the complete graph (with no self-loops) on the vertex set $[n]$, i.e., each receiver observes all but his own symbol. In this simple case, broadcasting the sum $\sum_{i=1}^n x_i$ (in $\mathbb{F}$) allows each receiver to retrieve his own symbol, hence $m= 1$. This problem is motivated by applications to distributed storage~\cite{AK15b}, on-demand video streaming (ISCOD, \cite{BirkK06}) and wireless networks (see, e.g., \cite{YZ99a}), where a typical scenario is that clients miss information during transmissions of the network, and the network is interested in minimizing the retransmission length by exploiting the side information clients already possess. In theoretical computer science, index coding is related to some important % communication models and problems in which players have overlapping information, such as the \emph{one-way} communication complexity of the index function~\cite{KNR01} and the more general problem of \emph{network coding}~\cite{Ashwede,ERL15}. Index coding can also be viewed as an interesting special case of nondeterministic computation in the (notoriously difficult to understand) multiparty \emph{Number-On-Forehead} model, which in turn is a promising approach for proving data structure and circuit lower bounds~\cite{Patrascu10, pudlak1997boolean, JuknaS11}. The minimum length of an index code for a given graph has well-known relations to other important graph parameters. For instance, it is bounded from below by the size of the maximum independent set, and it is bounded from above by the clique-cover number ($\chi(\bar{G})$) since for every clique in $G$, it suffices to broadcast a single symbol (recall the example above). The aforementioned connections also led to algorithmic connections (via convex relaxations) between the computational complexity of graph coloring and that of computing the minimum index code length of a graph \cite{ChlamtacH14}. In the context of circuit lower bounds, Riis \cite{R07} observed that a certain index coding problem is equivalent to the so-called \emph{shift conjecture} of Valiant \cite{V92} (see Subsection \ref{subsec_derendomization} below). If true, % this conjecture would resolve a major open problem of proving superlinear size lower bound for logarithmic-depth circuits. When the encoding function of the index code is \emph{linear} in $x$ (as in the example above), the corresponding scheme is called a \emph{linear index code}. In their seminal paper, Bar-Yossef et al. \cite{BBJK06} showed that the minimum length $m$ of a \emph{linear} index code is characterized precisely by a parameter of the knowledge graph $\mathcal{K}_n$, called the \emph{minrank} ($\mathsf{minrk_{\F}}(\mathcal{K}_n)$), first introduced by Haemers \cite{Haemers79} in the context of Shannon capacity of graphs.% \footnote{To be precise, this holds only for graphs without self-loops. We will ignore this minor issue in this paper as it will not affect any of our results.} Namely, $\mathsf{minrk_{\F}}(\mathcal{K}_n)$ is the minimum rank (over $\mathbb{F}$) of an $n\times n$ matrix $M$ that ``represents'' $\mathcal{K}_n$. By ``represents'' we mean a matrix $M$ that contains a zero in all entries corresponding to \emph{non-edges}, and non-zero entries on the diagonal. Entries corresponding to edges are arbitrary. (Over $\mathbb{F}_2$ this is equivalent to being the adjacency matrix of a subgraph of $\mathcal{K}_n$, with diagonal entries set to one.) Note that without the ``diagonal constraint", the above minimum would trivially be $0$, and indeed this constraint is what makes the problem interesting and hard to analyze. While linear index codes are in fact optimal for a large class of knowledge graphs (including directed acyclic graphs, perfect graphs, odd ``holes" and odd ``anti-holes" % \cite{BBJK06}), there are examples where non-linear codes outperform their linear counterparts~\cite{LU07}. In the same paper, Lubetzky and Stav \cite{LU07} posed the following question about \emph{typical} knowledge graphs, namely, \begin{quote} \emph{What is the minimum length of an index code for a random knowledge graph $\mathcal{K}_n=\mathcal{G}_{n,p}$? } % \end{quote} \noindent Here, $\mathcal{G}_{n,p}$ denotes a random Erd{\H o}s-R{\'e}nyi directed graph, i.e., a graph on $n$ vertices in which each arc is taken independently with probability $p$. In this paper, we partially answer this open problem by determining the optimal length of \emph{linear} index codes for such graphs. In other words, we prove a tight lower bound on the minrank of $\mathcal{G}_{n,p}$ for all values of $p\in[0,1]$. In particular, \begin{theorem}[Main theorem, informal] \label{thm_main_informal} For any constant $0<p<1$ and any field $\mathbb{F}$ of cardinality $|\mathbb{F}|<n^{O(1)}$, it holds with high probability that \[ \mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) = \Theta\left( \frac{n}{\log n} \right) \; . % \] \end{theorem} The formal quantitative statement of our result can be found in Corollary \ref{cor_main_thm_const_p} below. We note that our general result (see Theorem \ref{thm_mrk_gnp}) extends beyond the constant regime to \emph{subconstant} values of $p$, and this feature of our lower bound is crucial for potential applications of our result to circuit lower bounds (we elaborate on this in the next subsection). Theorem \ref{thm_main_informal} gives a near quadratic improvement over the previously best lower bound of $\Omega(\sqrt{n})$~\cite{LU07,HL12}, and settles the linear index coding problem for random knowledge graphs, as an $O_p(n/\log n)$ linear index coding scheme is achievable via the clique-covering solution (see Section \ref{sec_proof_overview}). In the following subsection, we propose a concrete (yet admittedly still quite challenging) approach for proving superlinear circuit lower bounds based on a potential ``derandomization" of Theorem \ref{thm_main_informal}. \subsection{Connections to circuit lower bounds for semilinear circuits} \label{subsec_derendomization} \paragraph{General log-depth circuits.} In his seminal line of work, Valiant \cite{V77,V83,V92} proposed a path for proving superlinear lower bounds on the size of % circuits with logarithmic depth, one of the main open questions in circuit complexity. Informally speaking, Valiant's ``depth reduction" method~\cite{V77,V09} allows, for any constant $\varepsilon$, to reduce any circuit of size $O(n)$ and depth $O(\log n)$ (with $n$ inputs and $n$ outputs), to a new circuit % with the same inputs and outputs, where now each output gate is an (arbitrary) Boolean function of $(i)$ at most $n^{\varepsilon}$ inputs which are ``hard-wired" to this output gate, and $(ii)$ an additional fixed set of $m=O_{\varepsilon}(n/\log\log n )$ ``common bits" $b_1(x),\ldots, b_m(x)$ which in general may be arbitrary Boolean functions of the input $x=x_1,\ldots, x_n$. Therefore, if one could exhibit a function that cannot be computed in this model using $O(n/\log\log n)$ common bits, this would imply a superlinear circuit lower bound for logarithmic depth circuits. Valiant~\cite{V92} proposed a concrete candidate hard function for this new model, namely the function whose input is an $n$-bit string $x$ and a number $i \in \{0,\ldots,n-1\}$ and whose output is the $i$th cyclic shift of $x$. Valiant conjectured that no ``pre-wired" circuit as above can realize \emph{all} $n$ cyclic shifts using $m=O(n/\log\log n)$ common bits (in fact, Valiant postulated that $m=\Omega(n)$ common bits are required, and this still seems plausible). This conjecture is sometimes referred to as \emph{Valiant's shift conjecture}. As noted earlier in the introduction, Riis \cite{R07} observed that a certain index coding problem is equivalent to this conjecture. Let $G=(V,A)$ be a directed graph, and $i\in\{0,\ldots,n-1\}$. We denote by $G^i$ the graph with vertex set $V$ and arc set $A^i=\{(u,v+i (\bmod~n)): (u,v)\in A\}$. Riis \cite{R07} showed that the following conjecture is equivalent to Valiant's shift conjecture: \begin{conjecture}\label{conj_riis} There exists $\varepsilon>0$ such that for all sufficiently large $n$ and every graph $G$ on $n$ vertices with max-out-degree at most $n^{\varepsilon}$, there exists a shift $i$ such that the minimum length of an index coding scheme for $G^i$ (over $\mathbb{F}_2$) is $\omega(n/\log\log n)$. \end{conjecture} \paragraph{Semilinear log-depth circuits.} Let us consider a function $f(x,p)$ whose input is partitioned into two parts, $x\in\{0,1\}^k$ and $p\in\{0,1\}^t$. We say that the function $f$ is \emph{semilinear} if for every fixed value of $p=p_0$, the function $f(x,p_0)$ is a linear function (over $\mathbb{F}_2$) of $x$. The class of semilinear functions is quite rich, and includes for instance bilinear functions in $x$ and $p$ (such as matrix multiplication) and permutations $\pi_p(x)$ of $x$ that may depend arbitrarily on $p$. A circuit $G$ is called \emph{semilinear} if for every fixed value of $p=p_0$, one can assign linear functions to the gates of $G$, so that $G$ computes $f(x,p_0)$. So it is only the circuit's topology that is fixed, and the linear functions computed by the gates may depend arbitrarily on $p$. It is easy to see that a semilinear function with a one-bit output can always be computed by a linear-size log-depth semilinear circuit (namely, the full binary tree). However, if we consider semilinear functions with $O(n)$ output bits, then the semilinear circuit complexity of a random function is $\Omega(n^2/\log{n})$ with high probability. It is an open problem to prove a superlinear lower bound against log-depth semilinear circuits~\cite{pudlak1997boolean}. This would follow from the semilinear variant of Valiant's shift conjecture, which is equivalent to the following slight modification of Conjecture~\ref{conj_riis}~\cite{pudlak1997boolean,R07}. \begin{conjecture}\label{conj_riis_semilinear} There exists $\varepsilon>0$ such that for all sufficiently large $n$ and every graph $G$ on $n$ vertices with max-out-degree at most $n^{\varepsilon}$, there exists a shift $i$ such that the minimum length of a \emph{linear} index coding scheme for $G^i$ (over $\mathbb{F}_2$) is $\omega(n/\log\log n)$. Equivalently, \[ \forall \; G \text{ of out-degrees at most $n^{\varepsilon}$} \;\; \exists \;\; i\in [n] \;\; \mathsf{minrk_{2}}(G^i) = \omega(n/\log\log n) \; .\] \end{conjecture} Theorem~\ref{thm_main_informal} (and the more precise concentration bound we prove in Theorem \ref{thm_mrk_gnp}) asserts that % with high probability, a graph chosen from $\mathcal{G}_{n,p}$ (with $p=n^{\varepsilon-1}$ for the expected degree of each vertex to be $n^{\varepsilon}$) has minrank $\Omega(n)$. Conjecture~\ref{conj_riis_semilinear} would follow from a ``derandomization'' of Theorem~\ref{thm_main_informal} in which we replace the distribution $\mathcal{G}_{n,p}$ with a random shift of an arbitrary given graph of the right degree. In fact, for the purpose of circuit lower bounds, one could replace cyclic shifts with any (efficiently computable) set of at most $\exp(O(n))$ permutations. (Since the permutation itself is part of the input, its description size must be linear in $n$.) \paragraph{Semilinear series-parallel circuits.} Finally, we mention one last circuit class for which the above ``derandomization" approach might be easier. Here we replace the depth restriction by another restriction on the topology of the circuit. Namely, a circuit $G=(V,A)$ is called \emph{Valiant series-parallel (VSP)}, if there is a labeling of its vertices $l\colon V \to \mathbb{R}$, such that for every arc $(u,v)\in A$, $l(u)<l(v)$, but there is no pair of arcs $(u,v),(u',v')\in A$, such that $l(u)<l(u')<l(v)<l(v')$. Most of the known circuit constructions (i.e., circuit upper bounds) are VSP circuits. Thus, it is also a big open question in circuit complexity to prove a superlinear lower bound on the size of semilinear VSP circuits (of arbitrary depth). Valiant~\cite{V77}, Calabro~\cite{C08}, and Riis~\cite{R07} show that in order to prove a superlinear lower bound for semilinear VSP circuits, it suffices to show that for a sufficiently large \emph{constant} $d$, for every graph $G$ of max-out-degree at most $d$, the minrank of one of its shifts is at least $n/100$. We note that Theorem~\ref{thm_main_informal} for this regime of $p=d/n$ gives a lower bound of $n/20$. Thus, derandomization of the theorem in this regime would imply a superlinear lower bound. Note that in the case of $p=O(n^{-1})$, the entropy of a random graph is only $O(n\log{n})$ bits, hence, information-theoretically it seems easier to derandomize than the case of $p=n^{\varepsilon-1}$. \subsection{Proof overview of Theorem \ref{thm_main_informal}} \label{sec_proof_overview} In \cite{LU07}, Lubetzky and Stav showed that for any field $\mathbb{F}$ and a directed graph $G$, $$\mathsf{minrk_{\F}}{(G)}\cdot\mathsf{minrk_{\F}}{(\bar{G})} \geq n \; .$$ This inequality gives a lower bound of $\Omega(\sqrt{n})$ on the expected value of the minrank of $\mathcal{G}_{n,1/2}$. (Indeed, the random variables $\mathcal{G}_{n,1/2}$ and $\bar{\mathcal{G}}_{n,1/2}$ have identical distributions). Since $\mathsf{minrk_{\F}}(\mathcal{G}_{n,p})$ is monotonically non-increasing in $p$, the same bound holds for any $p\leq 1/2$. Haviv and Langberg~\cite{HL12} improved this result by proving a lower bound of $\Omega(\sqrt{n})$ for all constant $p$ (and not just $p \le 1/2$), and also by showing that the bound holds with high probability. We now outline the main ideas of our proof. For simplicity we assume that $\mathbb{F}=\mathbb{F}_2$ and $p=1/2$. To prove that $\mathsf{minrk}_2(\mathcal{G}_{n,p}) \geq k $, we need to show that with high probability, $\mathcal{G}_{n,p}$ has no representing matrix (in the sense of Definition \ref{def_mrk}) whose rank is less than $k$. As a first attempt, we can show that any \emph{fixed} matrix $M$ with $1$s on the diagonal of rank less than $k$ has very low probability of representing a random graph in $\mathcal{G}_{n,p}$, and then apply a union bound over all such matrices $M$. Notice that this probability is simply $2^{-s+n}$, where $s$ is the sparsity of $M$ (i.e., the number of non-zero entries) and the $n$ is to account for the diagonal entries. Moreover, we observe that the sparsity $s$ of any rank-$k$ matrix with $1$s on its main diagonal must be\footnote{To see why, notice that any maximal linearly independent set of columns must ``cover'' all coordinates, i.e., there must not be any coordinate that is zero in all vectors, as otherwise we could take the column vector corresponding to that coordinate and it would be linearly independent of our set (due to the nonzero diagonal) in contradiction to maximality. Assuming all columns have roughly the same number of 1s, we obtain that each column has at least $n/k$ 1s, leading to the claimed bound. See Lemma~\ref{lem_rank_vs_sparsity_diagonal} for the full proof.} at least $\approx n^2/k$. Finally, since the number of $n\times n$ matrices of rank $k$ is $\approx 2^{2nk}$ (as a rank-$k$ matrix can be written as a product of $n\times k$ by $k\times n$ matrices, which requires $2nk$ bits to specify), by a union bound, the probability that $\mathcal{G}_{n,p}$ contains a subgraph of rank $<k$ is bounded from above by (roughly) $2^{2nk}\cdot (1/2)^{n^2/k}$, which is $\ll 1$ for $k=O(\sqrt{n})$. This recovers the previous $\Omega(\sqrt{n})$ lower bound of \cite{HL12} (for all constant $p$, albeit with a much weaker concentration bound). To see why this argument is ``stuck'' at $\sqrt{n}$, we observe that we are not overcounting and indeed, there are $2^{n^{3/2}}$ matrices of rank $k \approx n^{1/2}$ and sparsity $s \approx n^{3/2}$. For instance, we can take the rank $n^{1/2}$ matrix that consists of $n^{1/2}$ diagonal $n^{1/2} \times n^{1/2}$ blocks of $1$s (a disjoint union of $n^{1/2}$ equal-sized cliques), and replace the first $n^{1/2}$ columns with arbitrary values. Each such matrix has probability $2^{-n^{3/2}}$ of representing $\mathcal{G}_{n,p}$ (because of its sparsity) and there are $2^{n^{3/2}}$ of them, so the union bound breaks for $k=\Omega(\sqrt{n})$. In order to go beyond $\sqrt{n}$, we need two main ideas. To illustrate the first idea, notice that in the above example, even though individually each matrix has probability $2^{-n^{3/2}}$ of representing $\mathcal{G}_{n,p}$, these ``bad events'' are highly correlated. In particular, each of these events implies that $\mathcal{G}_{n,p}$ must contain $n^{1/2}-1$ disjoint cliques, an event that happens with roughly the same probability $2^{-n^{3/2}}$. Therefore, we see that the probability that the \emph{union} of these bad events happens is only $2^{-n^{3/2}}$, greatly improving on the naive union bound argument. (We remark that this idea of ``bunching together related events'' is reminiscent of the chaining technique as used, e.g., in analyzing Gaussian processes.) More generally, the first idea (and also centerpiece) of our proof is Lemma~\ref{lem_k'_n'_n_k}, which shows that every matrix must % contain a ``nice'' submatrix (in a sense to be defined below). The second and final idea, described in the next paragraph, will be to bound the number of ``nice'' submatrices, from which the proof would follow by a union bound over all such submatrices. Before defining what we mean by ``nice'', we mention the following elementary yet crucial fact in our proof: Every rank $k$ matrix is uniquely determined by specifying some $k$ linearly independent rows, and some $k$ linearly independent columns (i.e., a row basis and a column basis) including the indices of these rows and columns (see Lemma~\ref{lem_basis_determines_matrix}). This lemma implies that we can encode a matrix using only $\approx s_{basis}\cdot \log n$ bits, where $s_{basis}$ is the minimal sparsity of a pair of row and column bases that are guaranteed to exist. This in turn implies that there are only $\approx 2^{s_{basis} \log n}$ such matrices. Now, since the average number of $1$s in a row or in a column of a matrix of sparsity $s$ is $s/n$, one might hope that such a matrix contains a pair of row and column bases of sparsity $k\cdot (s/n)$, and this is precisely our definition of a ``nice'' matrix. (Obviously, not all matrices are nice, and as the previous example shows, there are lots of ``unbalanced'' matrices where the nonzero entries are all concentrated on a small number of columns, hence they have no sparse column basis even though the average sparsity of a column is very low; this is exactly why we need to go to submatrices.) To complete this overview, notice that using the bound on the number of ``nice'' matrices, the union bound yields \[ 2^{ks\log(n)/n}\cdot (1/2)^{s}, \] so one could set the rank parameter $k$ to be as large as $\Theta(n/\log n)$ and the above expression would still be $\ll 1$. A similar bound holds for nice submatrices, completing the proof. \section{Preliminaries} For an integer $n$, we denote the set $\{1,\ldots,n\}$ by $[n]$. For an integer $n$ and $0\le p\le 1$, we denote by $\mathcal{G}_{n,p}$ the probability space over the directed graphs on $n$ vertices where each arc is taken independently with probability $p$. For a directed graph $G$, we denote by $\chi(G)$ the chromatic number of the undirected graph that has the same set of vertices as $G$, and an edge in place of every arc of $G$. By $\bar{G}$ we mean a directed graph on the same set of vertices as $G$ that contains an arc if and only if $G$ does not contain it.\footnote{Throughout the paper we assume that graphs under consideration do not contain self-loops. In particular, neither $G$ nor $\bar{G}$ has self-loops.} Let $\mathbb{F}$ be a finite field. For a vector $v\in \mathbb{F}^n$, we denote by $v^{j}$ the $j$th entry of $v$, and by $v^{\leq j} \in \mathbb{F}^j$ the vector $v$ truncated to its first $j$ coordinates. For a matrix $M\in\mathbb{F}^{n\times n}$ and indices $i,j\in[n]$, let $M_{i,j}$ be the entry in the $i$th row and $j$th column of $M, \mathrm{Col}_i(M)$ be the $i$th column of $M$, $\mathrm{Row}_i(M)$ be the $i$th row of $M$, and $\mathsf{rk}(M)$ be the rank of $M$ over $\mathbb{F}$. By a \emph{principal submatrix} we mean a submatrix whose set of row indices is the same as the set of column indices. By the \emph{leading principal submatrix} of size $k$ we mean a principal submatrix that contains the first $k$ columns and rows. For a matrix $M\in\mathbb{F}^{n \times n}$, the sparsity $s(M)$ is the number of non-zero entries in $M$. We say that a matrix $M\in \mathbb{F}^{n\times n}$ of rank $k$ \emph{contains} an \emph{$s$-sparse column (row) basis}, if $M$ \emph{contains} a column (row) basis (i.e., a set of $k$ linearly independent columns (rows)) with a total of at most $s$ non-zero entries. \begin{definition}[Minrank~\cite{BBJK06, LU07}]\footnote{In this paper we consider the directed version of minrank. Since the minrank of a directed graph does not exceed the minrank of its undirected counterpart, a lower bound for a directed random graph implies the same lower bound for an undirected random graph. The bound is tight for both directed and undirected random graphs (see Theorem~\ref{thm:tight}). }\label{def_mrk} Let $G=(V,A)$ be a graph on $n=|V|$ vertices with the set of directed arcs $A$. A matrix $M \in \mathbb{F}^{n\times n}$ \emph{represents} $G$ if $M_{i,i}\neq0$ for every $i\in[n]$, and $M_{i,j}=0$ whenever $(i,j)\notin A$ and $i\neq j$. The minrank of $G$ over $\mathbb{F}$ is \[ \mathsf{minrk_{\F}}(G)= \min_{M \text{ represents }G} \mathsf{rk}(M) \; . \] \end{definition} \ We say that two graphs \emph{differ at only one vertex} if they differ only in arcs leaving one vertex. Following~\cite{HHMS10,HL12}, to amplify the probability in Theorem~\ref{thm_mrk_gnp}, we shall use the following form of Azuma's inequality for the vertex exposure martingale. \begin{lemma}[Corollary 7.2.2 and Theorem 7.2.3 in~\cite{AS15}]\label{thm:azuma} Let $f(\cdot )$ be a function that maps directed graphs to $\mathbb{R}$. If $f$ satisfies the inequality $|f(H)-f(H')|\le1$ whenever the graphs $H$ and $H'$ differ at only one vertex, then \[ \Pr[\left|f(\mathcal{G}_{n,p})-\mathbb{E}[f(\mathcal{G}_{n,p})]\right|>\lambda\sqrt{n-1}]<2e^{-\lambda^2/2} \; . \] \end{lemma} \section{The Minrank of a Random Graph} \label{sec_minkank_gnp} The following elementary linear-algebraic lemma shows that a matrix $M \in \mathbb{F}^{n\times n}$ of rank $k$ is fully specified by $k$ linearly independent rows, $k$ linearly independent columns, and their $2k$ indices. In what follows, we denote by $\mathcal{M}_{n,k}$ the set of matrices from $ \mathbb{F}^{n\times n}$ of rank $k$. \begin{lemma}[Row and column bases encode the entire matrix] \label{lem_basis_determines_matrix} Let $M \in \mathcal{M}_{n,k}$, and let $R = (\mathrm{Row}_{i_1}(M),\ldots , \mathrm{Row}_{i_k}(M)), C = (\mathrm{Col}_{j_1}(M),\ldots , \mathrm{Col}_{j_k}(M))$ be, respectively, a row basis and a column basis of $M$. Then the mapping $\phi\colon \mathcal{M}_{n,k} \to (\mathbb{F}^{1\times n})^k \times (\mathbb{F}^{n\times 1})^k \times [n]^{2k}$ defined as \[ \phi(M)=(R,C,i_1,\ldots,i_k,j_1,\ldots,j_k) \; , \] is a one-to-one mapping. \end{lemma} \begin{proof} We first claim that the intersection of $R$ and $C$ has full rank, i.e., that the submatrix $M'\in\mathbb{F}^{k\times k}$ obtained by taking rows $i_1,\ldots,i_k$ and columns $j_1,\ldots,j_k$ has rank $k$. This is a standard fact, see, e.g.,~\cite[p20, Section 0.7.6]{HornJohnson}. We include a proof for completeness. Assume for convenience that $(i_1,\ldots,i_k)=(1,\ldots, k)$ % and $(j_1,\ldots,j_k)=(1,\ldots, k)$. % Next, assume towards contradiction that $\mathsf{rk}(M')=\mathsf{rk}(\{\mathrm{Col}_1(M'),\ldots , \mathrm{Col}_{k}(M')\}) = k' < k$. Since $C$ is a column basis of $M$, every column $\mathrm{Col}_i(M)$ is a linear combination of vectors from $C$, and in particular, every $\mathrm{Col}_i(M')$ is a linear combination of $\{\mathrm{Col}_1(M'),\ldots , \mathrm{Col}_{k}(M')\}$. Therefore, the $k\times n$ submatrix $M'' \coloneqq (\mathrm{Col}^{\leq k}_1(M),\ldots , \mathrm{Col}^{\leq k}_{n}(M))$ has rank $k'$. On the other hand, the $k$ rows of $M''\colon \mathrm{Row}_1(M),\ldots , \mathrm{Row}_k(M)$ were chosen to be linearly independent by construction. Thus, $\mathsf{rk}(M'')=k>k'$, which leads to a contradiction. In order to show that $\phi$ is one-to-one, we show that $R$ and $C$ (together with their indices) uniquely determine the remaining entries of $M$. We again assume for convenience that $(i_1,\ldots,i_k)=(1,\ldots, k)$ % and $(j_1,\ldots,j_k)=(1,\ldots, k)$. Consider any column vector $\mathrm{Col}_i(M)$, $i\in [n]\setminus [k]$. By definition, $\mathrm{Col}_i(M) = \sum_{t=1}^k \alpha_{i,t} \cdot \mathrm{Col}_t(M)$ for some coefficient vector $\alpha_i\coloneqq (\alpha_{i,1},\ldots , \alpha_{i,k})\in\mathbb{F}^{k\times1}$. Thus, in order to completely specify all the entries of $\mathrm{Col}_i(M)$, it suffices to determine the coefficient vector $\alpha_i$. But $M'$ has full rank, hence the equation $$ M' \alpha_i^T = \mathrm{Col}^{\leq k}_i(M) $$ has a \emph{unique} solution. Therefore, the coefficient vector $\alpha_i$ is fully determined by $M'$ and $\mathrm{Col}^{\leq k}_i(M)$. Thus, the matrix $M$ can be uniquely recovered from $R, C$ and the indices $\{i_1,\ldots, i_k\}, \{j_1,\ldots,j_k\}$. \end{proof} The following corollary gives us an upper bound on the number of low-rank matrices that contain sparse column and row bases. In what follows, we denote by $\mathcal{M}_{n,k,s}$ the set of matrices over $\mathbb{F}^{n \times n}$ of rank $k$ that contain an $s$-sparse row basis and an $s$-sparse column basis. \begin{corollary}[Efficient encoding of sparse-base matrices]\label{cor_matrix_encoding} \[ \left| \mathcal{M}_{n,k,s} \right| \leq (n\cdot|\mathbb{F}|)^{6s} \; . \] \end{corollary} \begin{proof} Throughout the proof, we assume without loss of generality that $s \geq k$, as otherwise $\left| \mathcal{M}_{n,k,s} \right|=0$ hence the inequality trivially holds. The function $\phi$ from Lemma~\ref{lem_basis_determines_matrix} maps matrices from $\mathcal{M}_{n,k,s}$ to $(R,C,i_1,\ldots, i_k, j_1,\ldots, j_k)$, where $R$ and $C$ are $s$-sparse bases. Therefore, the total number of matrices in $\mathcal{M}_{n,k,s}$ is bounded from above by \[ \left({\binom{kn}{s}} \cdot |\mathbb{F}|^{s}\right)^2 \cdot n^{2k} \leq \left( (n^2)^{s} \cdot |\mathbb{F}|^{s} \right)^2 \cdot n^{2k} \leq (n\cdot|\mathbb{F}|)^{6s} \; , \] where the last inequality follows from $k\leq s$. \end{proof} Now we show that a matrix of low rank with nonzero entries on the main diagonal must contain many nonzero entries. To get some intuition on this, notice that a rank $1$ matrix with nonzero entries on the diagonal must be nonzero everywhere. Also notice that the assumption on the diagonal is crucial -- low rank matrices in general can be very sparse. \begin{lemma}[Sparsity vs.\ Rank for matrices with non-zero diagonal] \label{lem_rank_vs_sparsity_diagonal} For any matrix $M \in \mathbb{F}^{n\times n}$ with non-zero entries on the main diagonal (i.e., $M_{i,i}\neq0$ for all $i\in[n]$), it holds that \[ s(M) \geq \frac{n^2}{4 \mathsf{rk}(M)} \; . \] \end{lemma} \begin{proof} Let $s$ denote $s(M)$. The average number of nonzero entries in a column of $M$ is $s/n$. Therefore, Markov's inequality implies that there are at least $n/2$ columns in $M$ \emph{each of which} has sparsity at most $2s/n$. Assume without loss of generality that these are the first $n/2$ columns of $M$. Now pick a maximal set of linearly independent columns among these columns. We claim that the cardinality of this set is at least $n^2/(4s)$. Indeed, in any set of less than $n^2/(4s)$ columns, the number of coordinates that are nonzero in \emph{any} of the columns is less than \[ \frac{n^2}{4s} \cdot \frac{2s}{n} = \frac{n}{2} \] and therefore there exists a coordinate $i \in \{1,\ldots,n/2\}$ that is zero in all those columns. As a result, the $i$th column, which by assumption has a nonzero $i$th coordinate, must be linearly independent of all those columns, in contradiction to the maximality of the set. We therefore get that \[ \mathsf{rk}(M) \geq n^2/(4s) \; , \] as desired. \end{proof} The last lemma we need is also the least trivial. In order to use Corollary~\ref{cor_matrix_encoding}, we would like to show that any $n \times n$ matrix of rank $k$ has sparse row and column bases, where by sparse we mean that their sparsity is roughly $k/n$ times that of the entire matrix. If the number of nonzero entries in each row and column was roughly the same, then this would be trivial, as we can take any maximal set of linearly independent columns or rows. However, in general, this might be impossible to achieve. E.g., consider the $n \times n$ matrix whose first $k$ columns are chosen uniformly and the remaining $n-k$ columns are all zero. Then any column basis would have to contain all first $k$ columns (since they are linearly independent with high probability) and hence its sparsity is equal to that of the entire matrix. Instead, what the lemma shows is that one can always choose a \emph{principal submatrix} with the desired property, i.e., that it contains sparse row and column bases, while at the same time having relative rank that is at most that of the original matrix. \begin{lemma}[Every matrix contains a principal submatrix of low relative-rank and sparse bases] \label{lem_k'_n'_n_k} Let $M\in \mathcal{M}_{n,k}$ be a matrix. % There exists a principal submatrix $M'\in \mathcal{M}_{n',k'}$ of $M$, such that $k'/n' \leq k/n$, and $M'$ contains a column basis and a row basis of sparsity at most \[ s(M')\cdot\frac{2k'}{n'}\; . % \] \end{lemma} Note that if $M$ contains a zero entry on the main diagonal, the lemma becomes trivial. Indeed, we can take $M'$ to be a $1\times 1$ principal submatrix formed by this zero entry. Thus, the lemma is only interesting for matrices $M$ without zero elements on the main diagonal (i.e., when every principal submatrix has rank greater than $0$). \begin{proof} We prove the statement of the lemma by induction on $n$. The base case $n=1$ holds trivially. Now let $n>1$, and assume that the statement of the lemma is proven for every $m\times m$ matrix for $1\leq m < n$. Let $s(i)$ be the number of nonzero entries in the $i$th column plus the number of non-zero entries in the $i$th row (note that a nonzero entry on the diagonal is counted twice). Let also $s_{\max}=\max_{i}{s(i)}$. By applying the same permutation to the columns and rows of $M$ we can assume that $s(1)\leq s(2)\leq \cdots \leq s(n)$ holds. If for some $1\leq n'<n$, the leading principal submatrix $M'$ of dimensions $n'\times n'$ has rank at most $k'\leq n'k/n$, then we use the induction hypothesis for $M'$. This gives us a principal submatrix $M''$ of dimensions $n''\times n''$ and rank $k''$, such that $M''$ contains a column basis and a row basis of sparsity at most $s(M'')\cdot\frac{2k''}{n''}$. Also, by induction hypothesis $k''/n''\leq k'/n'\leq k/n$, which proves the lemma statement in this case. Now we assume that for all $n'<n$, the rank of the leading principal submatrix of dimension $n'\times n'$ is greater than $n'k/n$. We prove that the lemma statement holds for $M'=M$ for a column basis, and an analogous proof gives the same result for a row basis. For every $0\leq i\leq s_{\max}$, let $a_i=|\{j \,:\, s(j)=i\}|$. Note that \begin{equation}\label{eq:suma} \sum_{i=0}^{s_{\max}} a_i=n \; . \end{equation} Let us select a column basis of cardinality $k$ by greedily adding linearly independent vectors to the basis in non-decreasing order of $s(i)$. Let $k_i$ be the number of selected vectors $j$ with $s(j)=i$. Then \begin{equation}\label{eq:sumk} \sum_{i=0}^{s_{\max}} k_i=k. \end{equation} Next, for any $0 \leq t<s_{\max}$, consider the leading principal submatrix given by indices $i$ with $s(i)\leq t$. The rank of this matrix is at most $k'=\sum_{i=0}^t k_i$, and its dimensions are $n'\times n'$, where $n'=\sum_{i=0}^t a_i < n$. Thus by our assumption $k'/n'\geq k/n$, or equivalently, \begin{equation}\label{eq:ineq_on_t} \sum_{i=0}^tk_i\geq\frac{k}{n}\cdot \sum_{i=0}^ta_i \; . \end{equation} From~\eqref{eq:suma} and \eqref{eq:sumk}, \begin{equation}\label{eq:totalsum} \sum_{i=0}^{s_{\max}}{k_i}=\frac{k}{n}\cdot\sum_{i=0}^{s_{\max}}{a_i} \; . \end{equation} Now, \eqref{eq:ineq_on_t} and~\eqref{eq:totalsum} imply that for all $0\leq t\leq s_{\max}$:\begin{equation}\label{eq:suffix} \sum_{i=t}^{s_{\max}}{k_i} \leq \frac{k}{n}\cdot \sum_{i=t}^{s_{\max}}{a_i} \; . \end{equation} To finish the proof, notice that the sparsity of the constructed basis of $M$ is at most \begin{align*} \sum_{i=1}^{s_{\max}}i\cdot k_i =\sum_{t=1}^{s_{\max}}\sum_{i=t}^{s_{\max}}k_i \stackrel{\eqref{eq:suffix}}{\leq} \frac{k}{n}\cdot\sum_{t=1}^{s_{\max}}\sum_{i=t}^{s_{\max}}a_i =\frac{k}{n}\cdot\sum_{i=1}^{s_{\max}}i\cdot a_i=s(M)\cdot \frac{2k}{n} \; . \end{align*} \end{proof} Now we are ready to prove our main result -- a lower bound on the minrank of a random graph. \begin{theorem} \label{thm_mrk_gnp} \[ \Pr\left[\, \mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) \geq \Omega\left(\frac{n\log(1/p)}{\log{\left(n|\mathbb{F}|/p\right)}}\right) \, \right ] \; \geq \; 1-e^{-\Omega\left(\frac{n\log^2{(1/p)}}{\log^2{\left(n|\mathbb{F}|/p\right)}}\right)} \; . \] \end{theorem} \begin{proof} Let us bound from above probability that a random graph $\mathcal{G}_{n,p}$ has minrank at most $$k := \frac{n\log(1/p)}{C\log{\left(n|\mathbb{F}|/p\right)}},$$ for some constant $C$ to be chosen below. Recall that by Lemma~\ref{lem_k'_n'_n_k}, every matrix of rank at most $k$ contains a principal submatrix $M'\in\mathcal{M}_{n',k'}$ of sparsity $s' = s(M')$ with column and row bases of sparsity at most $$s'\cdot\frac{2k}{n},$$ where $k'/n'\leq k/n$. % By Corollary~\ref{cor_matrix_encoding}, there are at most $(n'\cdot|\mathbb{F}|)^{6(2s'k/n)}$ such matrices $M'$, and (for any $s'$) there are $\binom{n}{n'}$ ways to choose a principal submatrix of size $n'$ in a matrix of size $n\times n$. Furthermore, recall that Lemma~\ref{lem_rank_vs_sparsity_diagonal} asserts that for every $n',k'$, \begin{equation}\label{eq_density_M_prime} s'\geq \frac{n'^2}{4k'}. \end{equation} Finally, since $M'$ contains at least $s'-n'$ off-diagonal non-zero entries, $\mathcal{G}_{n,p}$ contains it with probability at most $p^{s'-n'}$. We therefore have \begin{align} \Pr\left[\mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) \leq k\right]\nonumber\\ &\leq\sum_{k',n',s'}\Pr\left[\mathcal{G}_{n,p} \text{ contains $M'\in\mathcal{M}_{n',k'},$ } s(M')=s', s(\text {bases of }M') \leq s'\cdot\frac{2k}{n}\right] \nonumber\\ &\leq\sum_{k',n',s'} \binom{n}{n'} \cdot p^{s'-n'} \cdot \left(n'\cdot|\mathbb{F}| \right)^{12s'k/n}\nonumber\\ &\leq\sum_{k',n',s'} 2^{n'\log{n}-s'\log(1/p)+n'\log(1/p)+(12s'k/n)\log{(n'|\mathbb{F}|)}} \label{mainbound} \; , \end{align} where all the summations are taken over $n', k'$, s.t. $k'/n' \leq k/n$ and $s'\geq \frac{n'^2}{4k'}$, and the first inequality is again by Lemma~\ref{lem_k'_n'_n_k}. We now argue that for sufficiently large constant $C$, all positive terms in the exponent of \eqref{mainbound} are dominated by the magnitude of the negative term ($s'\log(1/p)$). Indeed: \begin{align*} n'\log{n}+n'\log(1/p)+(12s'k/n)\log{(n'|\mathbb{F}|)} &=n'\log{(n/p)}+(12s'k/n)\log{(n'|\mathbb{F}|)}\\ \leq(4s'k'/n')\log{(n/p)}+(12s'k/n)\log{(n|\mathbb{F}|)} &\leq(16s'k/n)\log{(n|\mathbb{F}|/p)} =(16s'/C)\log{(1/p)} \; , \end{align*} where the first inequality follows from~\eqref{eq_density_M_prime}, and the second one follows from $k'/n' \leq k/n$. Thus, for $C>16$, \begin{align*} \Pr\left[\mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) \leq \frac{n\log(1/p)}{C\log{\left(n|\mathbb{F}|/p\right)}}\right] \leq n^4\cdot 2^{-\Omega(s'\log(1/p))}\leq 2^{-\Omega(\log(n))}. \end{align*} In particular, $\mathbb{E}\left[\mathsf{minrk_{\F}}(\mathcal{G}_{n,p})\right] \geq \frac{n\log(1/p)}{2C\log{\left(n|\mathbb{F}|/p\right)}}$. Furthermore, note that changing a single row (or column) of a matrix can change its minrank by at most $1$, hence the minrank of two graphs that differ in one vertex differs by at most $1$. We may thus apply Lemma~\ref{thm:azuma} with $\lambda=\Theta\left(\frac{\sqrt{n}\log(1/p)}{\log{\left(n|\mathbb{F}|/p\right)}} \right)$ to obtain \[ \Pr\left[\, \mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) \geq \Omega\left(\frac{n\log(1/p)}{\log{\left(n|\mathbb{F}|/p\right)}}\right) \, \right ] \; \geq \; 1-e^{-\Omega\left(\frac{n\log^2{(1/p)}}{\log^2{\left(n|\mathbb{F}|/p\right)}}\right)} \; . \] as desired. \end{proof} \begin{corollary}\label{cor_main_thm_const_p} For a constant $0<p<1$ and a field $\mathbb{F}$ of size $|\mathbb{F}|<n^{O(1)}$, \[ \Pr\left[\, \mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) \geq \Omega(n/\log{n}) \, \right ] \; \geq \; 1-e^{-\Omega\left(n/\log^2{n}\right)} \; . \] \end{corollary} \subsection{Tightness of Theorem~\ref{thm_mrk_gnp}} In this section, we show that Theorem~\ref{thm_mrk_gnp} provides a tight bound for all values of $p$ bounded away from~$1$ (i.e., $p \leq 1-\Omega(1)$). (See also the end of the section for the regime of $p$ close to $1$.) \begin{theorem}\label{thm:tight} For any $p$ bounded away from~$1$, \[ \Pr\left[\, \mathsf{minrk_{\F}}(\mathcal{G}_{n,p}) = O\left(\frac{n\log(1/p)}{{\log{n}+\log(1/p)}}\right) \, \right ] \; \geq \; 1- e^{-\Omega\left(n\right)}\; . \] \end{theorem} \begin{proof} We can assume that $p > n^{-1/8}$ as otherwise the statement is trivial. As we saw in the introduction, in the case of a clique (a graph with an arc between every pair of distinct vertices) it is enough to broadcast only one bit. This simple observation leads to the ``clique-covering'' upper bound: If a directed graph $G$ can be covered by $m$ cliques, then $\mathsf{minrk_{\F}}(G)\leq m$~\cite{haemers1978upper, BBJK06, HL12}. Note that the minimal number of cliques needed to cover $G$ is exactly $\chi(\bar{G})$. Thus, we have the following upper bound: For any field $\mathbb{F}$ and any directed graph $G$, \begin{align}\label{thm:upperbound} \mathsf{minrk_{\F}}(G)\leq \chi(\bar{G}) \; . \end{align} Since the complement of $\mathcal{G}_{n,p}$ is $\mathcal{G}_{n,1-p}$, it follows from~\eqref{thm:upperbound} that an upper bound on $\chi(\mathcal{G}_{n,1-p})$ implies an upper bound on $\mathsf{minrk_{\F}}(\mathcal{G}_{n,p})$. Let $\mathcal{G}^-_{n,p}$ denote a random Erd{\H o}s-R{\'e}nyi % \emph{undirected} graph on $n$ vertices, where each edge is drawn independently with probability $p$. For constant $p$, the classical result of Bollob{\'a}s~\cite{bollobas1988chromatic} asserts that the chromatic number of an undirected random graph satisfies \begin{equation}\label{eq:chiconst} \Pr\left[\chi(\mathcal{G}^-_{n,1-p})\leq \frac{n\log{(1/p)}}{2\log{n}}\left(1+o(1)\right)\right] > 1-e^{-\Omega(n)} \; . \end{equation} In fact, Pudl{\'a}k, R{\"o}dl, and Sgall~\cite{pudlak1997boolean} showed that~\eqref{eq:chiconst} holds for any $p>n^{-1/4}$. Since we define the chromatic number of a directed graph to be the chromatic number of its undirected counterpart, $\chi(\mathcal{G}_{n,1-p})=\chi(\mathcal{G}^-_{n,1-p^2})$. The bound~\eqref{eq:chiconst} depends on $p$ only logarithmically ($\log{(1/p)}$), thus, asymptotically the same bounds hold for the chromatic number of a random directed graph. \end{proof} The lower bound of Theorem~\ref{thm_mrk_gnp} is also almost tight for the other extreme regime of $p=1-\varepsilon$, where $\varepsilon=o(1)$. {\L}uczak~\cite{luczak1991chromatic} proved that for $p=1-\Omega(1/n)$, \begin{equation}\label{eq:chilarge} \Pr\left[\chi(\mathcal{G}^-_{n,1-p})\leq \frac{n(1-p)}{2\log{n(1-p)}}\left(1+o(1)\right)\right] > 1- \left(n(1-p)\right)^{-\Omega(1)} \; . \end{equation} When $p=1-\varepsilon$, the upper bound~\eqref{eq:chilarge} matches the lower bound of Theorem~\ref{thm_mrk_gnp} for $\varepsilon\geq n^{-1+\Omega(1)}$. For $\varepsilon=O(n^{-1})$,~\eqref{eq:chilarge} gives an asymptotically tight upper bound of $O(1)$. Thus, we only have a gap between the lower bound of Theorem~\ref{thm_mrk_gnp} and known upper bounds when $p=1-\varepsilon$ and $\omega(1) \leq n\varepsilon \leq n^{o(1)}$. \section*{Acknowledgements} We would like to thank Ishay Haviv for his valuable comments on an earlier version of this work. \bibliographystyle{alpha}
84ad2fde300ec2cceb1a6cbc39d9e36d82ab64ff
{ "file_path": "/home/ubuntu/dolma-v1_7/arxiv-0000.json.gz" }